ArticlePDF Available

Orientation and Navigation of Migrating Birds

Authors:

Abstract and Figures

The question of how migrating birds find their way to winter quarters and back has fascinated humans since the beginning of scientific research into avian biology. Migrating birds have been shown to possess compass systems that allow them to select and maintain certain compass directions. Three such systems are known, solar, stellar and magnetic. Their details are not quite clear and need further research. Hierarchy and interaction of compass systems of migrating birds are poorly studied; different species may vary in this respect. During migration, birds learn to use maps that make true navigation possible, i.e. to detect their position relatively to the goal of movement. The physical nature of navigational maps is an object of intensive research; currently the most promising concepts are the geomagnetic and possibly olfactory maps. A significant contribution to the study of formation of navigational maps was made by Soviet/Russian researchers, whose work was published in Zoologicheskii Zhurnal (Sokolov et al., 1984). Migrating birds have no innate map, and first-autumn individuals reach their species-specific wintering areas by using compass sense and counting time that should be spent moving in certain genetically fixed directions. However, in recent years more and more data surface that suggest that juveniles (maybe not of all species) do have some mechanism of controlling their position on the migratory route that allows them to compensate for errors of the spatio-temporal programme of migration.
Content may be subject to copyright.
788
ISSN 1062-3590, Biology Bulletin, 2016, Vol. 43, No. 8, pp. 788–803. © Pleiades Publishing, Inc., 2016.
Original Russian Text © N.S. Chernetsov, 2016, published in Zoologicheskii Zhurnal, 2016, Vol. 95, No. 2, pp. 128–146.
100 Years of Zoologicheskii Zhurnal
Orientation and Navigation of Migrating Birds1
N. S. Chernetsov
Biological Station Rybachy, Zoological Institute RAS, Rybachy, Kaliningrad region, 238535 Russia
Department of Vertebrate Zoology, St. Petersburg State University, St. Petersburg, 199034 Russia
e-mail: nikita.chernetsov@gmail.com
Received January 12, 2015
AbstractThe question of how migrating birds find their way to winter quarters and back has fascinated
humans since the beginning of scientific research into avian biology. Migrating birds have been shown to pos-
sess compass systems that allow them to select and maintain certain compass directions. Three such systems
are known, solar, stellar and magnetic. Their details are not quite clear and need further research. Hierarchy
and interaction of compass systems of migrating birds are poorly studied; different species may vary in this
respect. During migration, birds learn to use maps that make true navigation possible, i.e. to detect their posi-
tion relatively to the goal of movement. The physical nature of navigational maps is an object of intensive
research; currently the most promising concepts are the geomagnetic and possibly olfactory maps. A signifi-
cant contribution to the study of formation of navigational maps was made by Soviet/Russian researchers,
whose work was published in Zoologicheskii Zhurnal (Sokolov et al., 1984). Migrating birds have no innate
map, and first-autumn individuals reach their species-specific wintering areas by using compass sense and
counting time that should be spent moving in certain genetically fixed directions. However, in recent years
more and more data surface that suggest that juveniles (maybe not of all species) do have some mechanism of
controlling their position on the migratory route that allows them to compensate for errors of the spatio-tem-
poral programme of migration.
Keywords: birds, migration, orientation, navigation, compass systems, navigational map
DOI: 10.1134/S1062359016080069
INTRODUCTION
Migrating birds annually perform movements for
hundreds and thousands of kilometres, which allows
them to efficiently exploit resources in different cli-
matic zones. Since the first years of scientific research
into avian migration it became apparent that many
birds show breeding site fidelity, i.e. they are able to
return to an area with a radius of several kilometres
after spending winter in hundreds and thousands of
kilometres away (Sokolov, 1991). Considering the dis-
tance of migratory movements, recaptures of tens of
percent of birds at the sites of their previous breeding
obviously cannot result from random processes. If the
birds returned to the immediate area of their hatching
or previous breeding by chance only, recaptures of
individuals marked earlier would have been very rare
events, with just isolated cases over more than a hun-
dred years of bird ringing. The observed return rates in
long-distance migratory songbirds, up to 13% after the
first winter (Bushuev et al., 2012) and 30–40% in
adults (Sokolov, 1991), can only be explained by true
navigation, i.e. finding the goal of migration without
direct sensory contact with it. It has also been shown
that many migratory birds show winter site fidelity at a
level comparable to breeding site fidelity rates
(Mewaldt, 1976; Salewski et al., 2000). Similarly, only
true navigation can explain the ability of many sea-
birds, in particular procellariiforms, to locate small
islands where they breed, after long-distance flights in
the ocean, often involving circumnavigation of the
Earth and lasting for many years (Jouventin and
Weimerskirch, 1990; Åkesson and Weimerskirch,
2014).
In the middle of the 20th century the map and
compass concept was suggested, which assumed that a
migrating (or homing) bird should first detect where it
is located in respect to the goal (or, equivalently, where
the goal is located in respect to it; map step) and then
select and maintain the direction of movement
towards the goal (compass step, Kramer, 1953, 1957,
1961). The ability to use compass, i.e. to select and
maintain a certain compass direction, is called orien-
tation; the ability to use map, i.e. to detect the position
of the goal of movement without direct sensory con-
1The article was translated by the author.
BIOLOGY BULLETIN Vol. 43 No. 8 2016
ORIENTATION AND NAVIGATION OF MIGRATING BIRDS 789
tact with the goal, is called navigation. Generally
speaking, compass systems are necessary not only to
perform migratory movements (Wiltschko, R. and
Wiltschko, W., 2015). Any animal that needs to move
along a straight line for a long period of time (even if
the scale of movements is not very large), needs a
compass system independent of local landmarks. Nav-
igational map that makes it possible to detect the posi-
tion of the goal with which the animal has no sensory
contact (visual, acoustic, olfactory, etc.) is only neces-
sary for the animals that regularly leave their usual
home range and move far beyond it.
It should be emphasized that the map and compass
concept, which has been suggested in the mid 20th cen-
tury and remains a most important theoretical frame-
work of long-distance animal orientation and naviga-
tion research (Wiltschko, R. and Wiltschko, W., 2015),
is not linked to certain mechanisms of orientation and
navigation. The compass system and the navigation
system (map) may be based on different physical prin-
ciples and different sensory modalities. In other
words, the common question “How do migrating
birds find their way?” actually consists of two different
questions (the nature of compass and the nature of
map), and answers to these questions may be different.
METHODS OF ORIENTATION
AND NAVIGATION RESEARCH
The main method of avian orientation and naviga-
tion research are behavioural experiments. As soon as
it became possible to mark birds individually (ringing)
studies of homing success of experimentally displaced
birds started (usually towards the nest). Researchers
had good reasons to assume that at least during the
breeding period the motivation to home to the nest is
high, and interpreted homing success rate as a proxy
for navigation capabilities of a given species or popula-
tion (Rüppel, 1934, 1935; Wallraff, 2005). However,
radio tracking studies showed that the situation is
more equivocal: apart from motivation (which is
impossible to measure directly), there are other phys-
iological factors (e.g. the ability to rapidly change the
diel rhythm of activity) which may influence homing
performance and speed (Mukhin et al., 2009; Mukhin
and Kobylkov, 2014).
Many homing studies have been done in domestic
pigeons (Columba livia f. domestica), because these
birds show a high motivation to return to their loft
regardless of the season. It should be kept in mind
however that wild ancestors of homing pigeons, rock
pigeons, are sedentary birds that perform foraging
flights within several kilometres, up to maximum of
15–20 km. It was questionable whether compass and
especially map cues they use are identical to those used
by migrants for thousands of kilometres.
Significant progress in avian orientation research
was achieved when Gustav Kramer showed that spon-
taneous nocturnal migratory restlessness (Zugunruhe),
whose existence had been known long before the onset
of the scientific bird migration research (Berthold,
1996), is directed (Kramer, 1949). A migratory bird
put in a round cage during the migration season jumps
significantly more often towards the direction to
which it would have moved, were it free-flying. This
discovery made it possible to study spatial orientation
of birds under the controlled lab conditions and
resulted in a significant progress in avian orientation
and navigation research. It would be no exaggeration
to say that birds are the best studied animal long-dis-
tance migrants just because of the behavioural para-
digm that allows lab experiments.
Currently the progress in this field is reached by a
combination of behavioural experiments, both in the
lab in round arena and in free-flying radio-tagged
birds (Cochran et al., 2004; Thorup et al., 2007; Cher-
netsov et al., 2011; Schmaljohann et al., 2013); neuro-
anatomical studies (Heyers et al., 2010; Lefeldt et al.,
2014) and the combination of these methods (Zapka
et al., 2009; Kishkinev et al., 2013). Besides biophysi-
cal research, theoretical work combined with
behavioural experiments make a great contribution to
orientation research, especially to the study of the
magnetic compass (review: Kishkinev and Cher-
netsov, 2015). Research into orientation and naviga-
tion of birds and other animals is without exaggeration
one of the most integrative fields in zoology.
COMPASS SYSTEMS
The current consensus is that avian orientation sys-
tem incorporates redundancy. Usually the existence of
at least three different compass systems is acknowl-
edged: solar compass, based on the apparent move-
ment of the Sun on the sky (Kramer, 1951, 1953;
Schmidt-Koenig, 1990), stellar compass, based on the
pattern of constellations (Emlen, 1967, 1967a; Wiltsc-
hko, W. et al., 1987; Wiltschko, R. and Wiltschko, W.,
2009a), and magnetic compass, based on the lines of
the geomagnetic field (Wiltschko, W. and Wiltschko, R.,
1972). The proposed use of other cues for selecting
and maintaining compass directions, e.g. the moon,
has not been proven and is questionable (Wiltschko, R.
and Wiltschko, W., 1999).
Solar and stellar compasses are sometimes lumped
under the name of astronomical cues, as both of them
are based on the apparent rotation of the celestial bod-
ies and on the visual system. They are contasted with
the magnetic compass which is based on radically dif-
ferent principles and utilises an own sensory system
(magnetoreception; Mouritsen, 2013, 2015). How-
ever, a detailed analysis shows that solar and stellar
compasses differ significantly, and lumping them
together is probably not justified.
790
BIOLOGY BULLETIN Vol. 43 No. 8 2016
CHERNETSOV
Solar Compass
Solar compass is based on the regular apparent
movement of the Sun across the sky from the east to
the west via south (in the Northern Hemisphere) or via
north (in the Southern Hemisphere). Using this regu-
lar movement for detecting compass direction is only
possible when taking the time of the day into account,
i.e. solar compass is time-dependent. In the Northern
Hemisphere at solar noon the sun is exactly in the
south—but to use this information, one should know
when the solar noon is.
Furthermore, for the efficient use of the solar com-
pass it is necessary to account for the non-uniform
rate of solar azimuth change: just after sunrise, the Sun
rapidly climbs, and just before sunrise, it rapidly
descends, with a low rate of horizontal movement
(azimuth change). Conversely, around solar noon, the
elevation of the Sun above the horizon changes slowly,
and the azimuth changes more rapidly than in the
morning and in the evening. The form of the solar arch
depends on the geographic latitude and on season.
As it is not possible to experimentally manipulate
the position of the Sun, and a realistic imitation in a
planetarium is challenging, most studies of the solar
compass have been performed by clock shifting. The
birds are for a sufficient period of time kept in an arti-
ficial photoperiod which is shifted by several hours as
compared with the natural one. When the internal
clock of the birds is synchronised with this artificial
photoperiod, the birds are tested outdoors in sunny
weather. It is not possible to move the Sun, but one
can shift the animal’s internal clock, which gives an
equivalent result in the case of a time-dependent com-
pass mechanism. When interpreting the results of such
tests, one must take the non-uniform movement of the
Sun into account. Many authors assume that shifting
the clock by 1 hour results in the solar compass being
offset by 15°. This assumption is, generally speaking,
not correct: in summer at 50° N, a six-hour clock shift
may offset the solar compass by as much as 130°
(Wiltschko, R. and Wiltschko, W., 2015). The experi-
ments in homing pigeons showed that these birds
account for that (Wiltschko, R. et al., 2000). Moreover,
pigeons when using the solar compass, consider only
the azimuth of the Sun and ignore its elevation above
the horizon (Schmidt-Koenig, 1958; Keeton, 1979).
Most research into solar compass of birds has been
done in homing pigeons (Schmidt-Koenig, 1958;
Keeton, 1979; Wiltschko, W. et al., 1976, 1984). The
reason is that clock shifting experiments are easier to
perform in the behavioural context of homing, and
homing is easier to study in pigeons. To clock shift a
songbird, one has to take it into captivity for several
days (Mukhin et al., 2009), which may significantly
impact its motivation to return to its nest. Therefore
most studies of solar compass in passerines have been
done in a non-migratory behavioural context (Wiltsc-
hko, W. and Balda, 1989; Duff et al., 1998; Wiltschko, W.
et al., 1999). It has been even suggested that solar com-
pass exists but is not used by the birds for migratory
orientation (Wiltschko, R. and Wiltschko, W., 2015).
This viewpoint is difficult to support: at least some
migrating birds, e.g. grey-cheeked thrushes (Catharus
minimus) and Swainson’s thrushes (C. ustulatus),
select the direction of migratory flights on the basis of
the information derived from sunset cues (i.e. basically
from the sun), but maintain the direction by magnetic
compass (Cochran et al., 2004). The existence of such
a mechanism has been also proposed for other song-
bird species, e.g. for American sparrows (Muheim
et al., 2006a, 2007, 2009), but for thrushes it has been
proven beyond reasonable doubt.
Apart from the time-dependent solar compass,
another method of determining cardinal directions
from the movement of the Sun has been proposed
(Muheim et al., 2006a). This hypothetical method is
based on the idea that if a bird detects (directly or from
the band of maximum polarisation on the sky) the
sunset point and remembers its position, and then,
without changing its position, detects the sunrise
point, it would be able to find the bisector of the angle
between the sunset and sunrise directions. This bisec-
tor is always, in any season and everywhere on earth
(except of the geographic poles, but they are not very
relevant for migrating birds), the north—south axis.
This method of orientation with the help of the Sun is
very simple and elegant, but the claim that it is sup-
ported by the experimental data (Muheim et al.,
2006a) is erroneous (Liu and Chernetsov, 2012). The
existence of this time-independent form of the solar
compass is still pending confirmation.
Most probably, the classic time-dependent solar
compass is used by the birds, primarily by diurnal
migrants, for orientation. The paucity of supporting
data is probably due to the methodological difficulties
of studying solar compass in the migratory context and
to the difficulty of using round arenas for the study of
directedness of diurnal migratory activity.
Stellar Compass
The second celestial compass system used by
migratory birds is the star compass. The ability of
migratory birds to orient by the pattern of the starry
sky was demonstrated soon after the discovery of the
sun compass (Sauer, F., 1956, 1957) and subsequently
confirmed by various authors (Katz and Mihelson,
1978; Bingman, 1984, 1987; Mouritsen and Larsen,
2001). It was initially assumed that birds derive posi-
tional information from the stars (Sauer, F., 1957;
Sauer, E.G.F. and Sauer, E.M., 1960), but soon it was
shown that this was not the case: stellar information is
used for compass orientation, but is not a component
of map (Wallraff, 1960).
It was shown in the 1960s that the star compass of
migrating birds undergoes a complex ontogenetic
BIOLOGY BULLETIN Vol. 43 No. 8 2016
ORIENTATION AND NAVIGATION OF MIGRATING BIRDS 791
development (Emlen, 1967a). During the pre-migra-
tory period, juvenile indigo buntings (Passerina cya-
nea) learn to find the centre of rotation of the starry
sky located near the Polaris, which in the northern
hemisphere shows the direction towards geographic
North. By the moment of starting autumn migration,
indigo buntings in the experiments of Stephen Emlen
had learned to use the pattern of constellations to deter-
mine the geographic North, and no longer needed the
rotation. Birds raised in the planetarium under the sky
that was rotated around Betelgeuse (αOrion), under
the natural autumnal sky considered the direction
towards Betelgeuse to be the North and offset the inher-
ited population-specific autumn migratory direction
from this false “North” (Emlen, 1967a, 1970). They did
not care that the sky was rotating in a different fashion
as compared to their pre-migratory periods, i.e. stellar
rotation was no longer used as a reference.
Stephen Emlen performed his experiments in a
planetarium which provided a rather realistic imita-
tion of the starry sky (Emlen, 1970). Further studies
showed that a rather rough imitation (just 16 light
dots) is already sufficient for juvenile migrants and
allows them to find the rotation centre and to learn to
offset the directions from it, “believing” that it is the
North (Wiltschko, W. et al., 1987; Able, K.P. and
Able, M.A., 1990; Weindler et al., 1997). The birds
learn to find the North without observing the slow
motion of the stars, but just on the basis of the learned
geometric pattern of the light dots (real stars or their
imitation). Stellar compass developed this way no lon-
ger needs celestial rotation to operate (under the sta-
tionary sky in the planetarium the birds orient exactly
the same as under the rotating one) and is time-inde-
pendent (Emlen, 1970; Wiltschko, W. et al., 1987;
Weindler et al., 1997), which makes it significantly
distinct from the solar compass.
The ontogenetic development of the stellar com-
pass in indigo buntings described by Stephen Emlen is
astonishingly complex. Its use demands well devel-
oped cognitive abilities, much more sophisticated
than the use of the magnetic compass, which usually
amazes humans most. The magnetic compass is based
on the insufficiently studied sensory modality absent
in humans, which is on top of that probably based on
rather subtle quantum mechanical processes (Krylov
et al., 2014; Kishkinev and Chernetsov, 2015). How-
ever, stellar compass cannot function without devel-
oped cognitive abilities (Chernetsov, 2015). In partic-
ular, the birds need to detect a very slow rotational
motion of the sky (15° per hour) and to identify the
centre of rotation, which is a non-trivial cognitive task
(Mukhin, 1999; Alert et al., 2015). The ontogenetic
development of this compass system has been shown
in garden warblers (Sylvia borin) as well as indigo bun-
tings (Wiltschko, W. et al., 1987; Weindler et al., 1997).
In a number of cases researchers failed to show a work-
ing star compass formed during observations of the
rotating sky, possibly due to phototactic artefacts
which are difficult to avoid in captivity (Michalik
et al., 2014).
Besides, the time-independent nature of the devel-
oped stellar compass actually remains to be formally
shown. This feature of the star compass, which makes it
distinct from the sun compass, has been shown in indigo
buntings (Emlen, 1967a) and confirmed in the experi-
ments in pied flycatchers (Ficedula hypoleuca) and black-
caps (Sylvia atricapilla) (Mouritsen and Larsen, 2001).
However, both these studied were performed in the mag-
netic field which was disturbed as compared with the nat-
ural one, but could provide directional information.
Probably the developed star compass of migrating birds
indeed does not need a correctly synchronised internal
clock to function properly, but this remains to be demon-
strated.
Magnetic Compass
Magnetic compass is the third independent com-
pass systems of migrating birds. This compass is based
on a sensory modality which is not present in humans;
therefore, its use amazes researchers. This was proba-
bly the reason for the scepticism which met the
authors of the first reports on the use of geomagnetic
information by migrating birds (Merkel and Wiltsc-
hko, W., 1965; Wiltschko, W., 1968). Authorities on
bird migration wrote that “studies in the non-celestial
orientation of birds are not more successful than the
studies of human telepathy. Fading and again reviving
interest to this field is fuelled by the interest to the
enigmatic and mystical” (Dolnik, 1973). The situation
changed when a key paper was published in Science,
which reported the existence and the inclination
nature of the magnetic compass in European robins
(Erithacus rubecula) (Wiltschko, W. and Wiltschko, R.,
1972), and when the proponents and the opponents of
the magnetic orientation performed a joint study and
published results that convincingly showed that the
former group had been right (Emlen et al., 1976).
Currently the existence of the magnetic compass
systems in birds is not questioned by serious research-
ers. There are also reports on using magnetic compass
by mammals (Burda et al., 1990; Deutschlander et al.,
2003; Holland et al., 2006), sea turtles (Lohmann, K.J.,
1991; Lohmann, K.J. and Lohmann, C.M.F., 1993),
amphibians (Deutschlander et al., 1999; Diego-
Rasilla et al., 2010, 2013) and bony fishes (Quinn
et al., 1981). The sensory mechanism of the magnetic
compass of animals remained unknown for a long time
after its discovery, but significant progress has been
achieved in this field in the last 15–20 years (Wiltsc-
hko, R. and Wiltschko, W., 2015).
The most widely held theory is that the detection of
geomagnetic compass information is based on the per-
ception of the magnetic field through light-dependent
biradical chemical reactions, the receptor molecule is
a cryptochrome protein, and magnetoreception takes
792
BIOLOGY BULLETIN Vol. 43 No. 8 2016
CHERNETSOV
place in the retina. As a result, geomagnetic field is
perceived as a visual image, i.e. birds literally “see” the
magnetic field (Solov’yov et al., 2010; Mouritsen, 2013,
2015). For detailed reviews of the working mechanism of
the avian magnetic compass, the readers are referred to
earlier publications (Kishkinev and Chernetsov, 2015;
Krylov et al., 2014). However, it should also be men-
tioned that it has recently been shown that the biophysi-
cal model of biradical magnetoreception suggested earlier
(Schulten et al., 1978; Ritz et al., 2000) cannot explain
some experimental effects (Kavokin, 2009; Kavokin et
al., 2014; Engels et al., 2014) and needs to be modified.
Most probably, the general working principles of the
magnetic compass are correctly described by the current
model, but some important points need clarification and
possibly correction.
The magnetic compass, unlike solar and stellar
compasses, is believed to be innate and not learned. As
it does not involve the movement of celestial bodies,
the magnetic compass can be used 24 hours a day and
does not depend on the animal’s internal clock. Light-
dependence of the magnetic compass means that it
cannot function in complete darkness, but illumina-
tion characteristic of a starry moonless night, is suffi-
cient. One may think that it is a perfect compass
mechanism which makes celestial cues obsolete.
However, it should be kept in mind that magnetic
poles do not coincide with geographic ones, so the
direction towards the magnetic North is generally
speaking not the same as the direction toward the geo-
graphic North. The angle between these two direc-
tions, known as magnetic declination, may vary
between –20°…+20° in the areas regularly frequented
by migrating birds (Fig. 1). As a result some migrants,
e.g. a Swainson’s thrush population breeding in Alaska
(declination +20°) and wintering in northern South
America (declination –10°; Delmore et al., 2012;
Ruegg et al., 2014), visit areas where the true geo-
graphic azimuth of the magnetic North varies by 30°.
Such an error of the compass systems is unacceptable.
A bird from such a population would not be able to
successfully solve orientation tasks during its annual
cycle if it only uses a magnetic compass.
Hierarchy and Integration of Compass Systems
The redundancy of information provided by com-
pass systems apparently demands a hierarchy and inte-
gration rules. It has been repeatedly shown that even
though the magnetic compass is believed to be innate,
its successful use during migration in many bird spe-
cies depends on the ability to observe celestial rota-
tion, diurnal (Able, K.P. and Able, M.A., 1990a, 1993;
Weindler et al., 1998) or nocturnal (Emlen, 1970; Shu-
makov and Zelenova, 1988; Able, K.P. and Able, M.A.,
1990; Weindler et al., 1995; Michalik et al., 2014).
Birds that had not been able to observe celestial rota-
tion during their growth phase before the onset of first
Fig. 1. Map of the isolines the magnetic declination (isogones). Based on the data of U.S. NOAA National Geophysical Data
Center and Cooperative Institute for Research in Environmental Sciences; source: http://ngdc.noaa.gov/geomag/WMM/data/
WMM2015/WMM2015_D_MERC.pdf. Isogones are shown after every 2° of declination.
0
70°
15°
30°
30°
45°
60°
70°
45°
60°
15°
180°
0°
135°13 5 °180°90°90°45°45°0°
20
10
20
10
−10
−10
−30
−20
10
0
−20
−10
−10
10
20
−20
20
30
40
50
0
0
−30
−60
−70
−80
−110
−120
−130
−100
90
0
20
30
40
50
60
70
80
10
70
60
100
130
110
80
90
50
40
30
20
10
10
10
0
N
W
BIOLOGY BULLETIN Vol. 43 No. 8 2016
ORIENTATION AND NAVIGATION OF MIGRATING BIRDS 793
migration, when tested in the natural magnetic field
without access to astronomical cues either were disori-
ented, or showed a bimodal orientation along the
migratory axis, i.e. were not able to select the correct
end of the axis. Juvenile European robins that had
been allowed to observe the imitated rotation of the
sky for three weeks, proved able to calibrate their mag-
netic compass, whereas the birds that had observed the
rotation for one week or less, were disoriented like
those that did not see the rotating sky at all (Michalik
et al., 2014). Not only the duration of exposure under
the rotating celestial sphere in the post-juvenile
period, but also exposure in the certain sensitive
period may have played a role; similarly to the forma-
tion of the navigational map of “home,” to which a
bird returns after its first non-breeding season (Löhrl,
1959; Sokolov et al., 1984, 1987). The general pattern
seems to be that before the onset of first migration, the
magnetic compass of songbirds is calibrated by the
celestial rotation. It is noteworthy that in this particu-
lar case the solar compass (based on the movement of
the Sun), and the rotation of the stars behave as a sin-
gle celestial compass system.
During migration the interaction of compass sys-
tems looks different. Many authors have reported the
results of their cue conflict experiments, during which
they manipulated either magnetic or celestial informa-
tion (by mirrors or polarising filters). The results were
conflicting (see reviews Muheim et al., 2006; Liu and
Chernetsov, 2012). In some cases migrants calibrated
the celestial (stellar) compass from the magnetic one
(Wiltschko et al., 1998, 1999; Sandberg et al., 2000), in
others, conversely, the direction was selected from
sunset cues and maintained during nocturnal flight,
when neither the Sun nor the polarisation pattern are
visible, by the magnetic cues (Moore et al., 1985;
Cochran et al., 2004; Muheim et al., 2006a, 2007,
2009). In some experiments compass systems did not
interact at all, and simple domination of one of the
compasses was observed, i.e. the wrong readings of the
experimentally manipulated compass systems were
simply ignored (Katz and Mihelson, 1978; Wiltschko, R.
et al., 2008; Chernetsov et al., 2011; Schmaljohann
et al., 2013; Åkesson et al., 2015). In some cases the
same team of researchers in the same model species
obtained contradicting results by different methods
(Gaggini et al., 2010; Guinchi et al., 2015).
It should be emphasized that this complex and
intricate pattern should not be interpreted to mean
that nothing is certain or that many, if not most, stud-
ies are erroneous and should be discarded. Even
though many authors have criticised the approaches of
their colleagues, possibly sometimes with reason
(Wiltschko, R. et al., 2008, 2008a; Muheim et al.,
2008), at least some studies are performed very rigor-
ously so that their results are highly unlikely to be
caused by artefacts. For example, it is difficult to find
any flaws in the study by Cochran et al. (2004) which
found that grey-cheeked and Swainson’s thrushes
select the direction of their nocturnal migratory flights
by sunset polarisation information, and maintain it by
the magnetic compass (i.e., that they calibrate the
magnetic compass from solar cues). No less convinc-
ing are the studies that demonstrated that song
thrushes (Turdus philomelos) and Northern wheatears
(Oenanthe oenanthe) do not do that (Chernetsov et al.,
2011; Schmaljohann et al., 2013) but show a simple
domination of one of the compass systems, most prob-
ably the magnetic one.
The most parsimonious explanation of these
diverse results is, in our opinion, the suggestion that
compass system hierarchy differs between the species
of migrants, or even between populations within spe-
cies. Some birds, for instance those that perform long-
distance migrations and cross the areas with very dif-
ferent magnetic declination values, have to regularly
check their magnetic compass from the celestial one.
Other species that migrate shorter distances or even
longer flights which remain within similar declination
values (noteworthy, declination does not need to be
close to zero; it suffices if it is relatively uniform
throughout the route), do not need this cognitively
challenging system and get by with the magnetic (or
stellar) compass alone. Furthermore, mortality risk of
an orientation error is much different for a bird that
migrates for short or medium distance over land, with-
out crossing major ecological barriers (e.g., a Euro-
pean robin or a chaffinch (Fringilla coelebs) from a
Baltic population; Payevsky, 1971), than for a bird that
crosses deserts or oceans and winters in relatively small
islands (e.g. for a bar-tailed godwit (Limosa lapponica
baueri) flying from Alaska to New Zealand (Gill et al.,
2009)). In the former case a directional error of 5° or
even 10° may have no negative consequences, whereas
in the latter case an error of 30 is fatal. If natural selec-
tion forces different species and populations of
migrants to solve navigational and orientation tasks of
varying complexity, it is to be expected that the real-
ised systems will differ, even if sensory basis of orien-
tation is similar or identical. Selection will not shape
“the perfect navigator” (just like it does not support
“the perfect flying machine”; Dolnik, 1995), but will
support the accuracy of orientation and navigation
that is sufficient for population survival.
It may also be the case that some migratory birds
may secondarily lose some orientation systems, or, to
put it more cautiously, these systems may function sig-
nificantly worse in some species than in others (Cher-
netsov, 2015). Some data suggest that not all species of
migrants may successfully use all three compass sys-
tems (solar, stellar and magnetic): garden warbler
devoid of access to geomagnetic information at sunset
(i.e., unable to calibrate the magnetic compass from
sunset cues), were disoriented during the deep night,
even if they had access to the correct stellar informa-
tion (Pakhomov and Chernetsov, 2014). This certainly
makes the pattern even more complex.
794
BIOLOGY BULLETIN Vol. 43 No. 8 2016
CHERNETSOV
NAVIGATIONAL MAP
Unlike compass systems, about which relatively
much is known (even though, as we have just shown,
many questions remain), the physical nature of the
navigation system (the map) is much less clear. There
are two main concepts of how the map could be organ-
ised. One of them is the mosaic map concept (Wallraff,
1974), which assumes that an animal remembers a
network of local cues around the area it needs to return
to in the future. These local cues, or beacons, may be
landscape ones (visual), or perhaps olfactory or mag-
netic ones. Most humans use the mosaic map when
moving across a familiar terrain: we know how to walk
or to dri ve from a s ubway sto p to ou r home, from hom e
to the nearest supermarket or to school, etc.
Mosaic map is not identical to a cognitive map sensu
Tolman (1948) and subsequent authors (O’Keefe and
Nadel, 1978). A cognitive map concept assumes that
an animal has a mental representation of the geomet-
ric parameters of an area and thus can take novel
routes (make shortcuts) between localities in this area.
A mosaic map suitable for navigation may not have
these qualities.
Another variant is the gradient map concept (Wall-
raff, 1974). This concept assumes that birds (or other
animals) measure a gradient that regularly and pre-
dictably varies along a spatial scale comparable to the
scale of the intended movement, or ideally, slightly
beyond the scale of the movement. The gradient may
be of geomagnetic field parameters (total intensity,
inclination and declination), or possibly of concentra-
tions of certain volatile compounds in the atmosphere.
Unlike a mosaic map, which probably cannot be used
at distances exceeding several tens of kilometres, a gra-
dient map can potentially be used for navigation across
hundreds and thousands of kilometres. To do that,
migrating birds need a bicoordinate grid of two inter-
secting gradients, roughly equivalent to geographic
latitude and longitude.
The use of migrating birds of a gradient map, at
least at the scale of several kilometres or tens of kilo-
metres, has been demonstrated in an important field
experiment performed by Sokolov and colleagues on
the Courish Spit on the Baltic coast (Sokolov et al.,
1984). The authors showed that in order to be able to
return to their natal area in the subsequent year, juve-
nile chaffinches must be able to move freely across
that area. The birds that were kept in a semi-open
pavilion until 50–70 days old, and then moved beyond
their natal area and released, proved unable to return.
It seems that to form bonds with the natal area (i.e., to
form a map), chaffinches need to move at least at the
scale of several hundred metres or several kilometres
and to sample some physical gradients.
As for the physical basis of navigation map, the
most popular and actively discussed possibilities are
currently the olfactory and the magnetic map (Kishki-
nev, 2015). Apart from these two hypotheses, infra-
sound and gravitational anomalies are also suggested
as the source of navigational information. However,
these concepts are seriously discussed by a smaller
number of researchers.
Olfactory Map
Olfactory map as a source of map information was
first suggested by Papi and co-authors (Papi et al.,
1971, 1972) and has been actively discussed over the
last 40 years (Wallraff, 1999, 2005; Gagliardo, 2013).
Most experimental evidence supporting this theory
has been obtained in homing pigeons. In a nutshell,
birds deprived of olfactory information (e.g., by plug-
ging their nostrils, destroying the olfactory epithelium
by washing the nostrils with zinc sulphate ZnSO4 solu-
tion, application of local anaesthesia, or, most radi-
cally, by ablation of the olfactory nerve) perform much
poorer during homing to their loft than control birds
(for a review, see Wallraff, 2005). This affects all of the
main parameters that are commonly used for estimat-
ing homing efficiency in pigeons: scatter of vanishing
directions, returning time and the proportion of
returning individuals. Many early experiments were
performed without correct controls (Wallraff, 1980,
1981), which provided reasons to criticise the olfactory
map concept. In recent years proponents of the olfac-
tory map published the results of new experiments
with all the necessary controls (Gagliardo et al., 2006,
2008, 2009). These experiments showed that only
pigeons with intact olfactory nerves were able to home
successfully, whereas the pigeons with the ablated
nerve proved unable to solve this task. Ablation of the
ophthalmic branch of the trigeminal nerve (V1),
which is believed to transmit the magnetic information
to the brain (Mora et al., 2004; Heyers et al., 2010;
Lefeldt et al., 2014), did not influence homing perfor-
mance in pigeons. Similar results were obtained in
Cory’s shearwaters (Calonectris borealis): these birds
also use information transmitted via the olfactory
nerve for homing, but can return to the colony without
V1-transmitted information (Gagliardo et al., 2013).
The ability of procellariiforms to home successfully
without geomagnetic information had been earlier
independently shown by various authors (Benhamou
et al., 2003; Bonadonna et al., 2003; Mouritsen et al.,
2003).
It should be mentioned however that the olfactory
hypothesis is still being criticised. The main issue is
that it is difficult to understand which volatile com-
pounds could enable the existence of sufficiently sta-
ble gradients across tens, hundreds and even thou-
sands of kilometres. Modelling has shown that over
the ocean, where atmospheric flows are more stable
and predictable, such gradients are more or less imag-
inable, whereas in continental areas the olfactory
landscape is too dynamic to be used to navigation
(Waldvogel, 1987). The proponents of the olfactory
concept object and maintain that the current data
BIOLOGY BULLETIN Vol. 43 No. 8 2016
ORIENTATION AND NAVIGATION OF MIGRATING BIRDS 795
show that the concept is realistic from the viewpoint of
atmospheric physics (Wallraff and Andreae, 2000;
Wallraff, 2013), but this issue cannot be considered
resolved. The existing experimental evidence of olfac-
tory map use mainly refers to seabirds and homing
pigeons. It is conceivable that over the ocean, smell
gradients indeed play an important role in navigation;
in the case of homing pigeons, one may assume that
the olfactory navigational abilities of their sedentary
wild ancestors were hypertrophied by artificial selec-
tion. The experiment with grey catbirds (Dumetella
carolinensis) stands apart. Grey catbirds displaced by
1000 km from their migratory route were unable to
compensate for the displacement if their olfactory epi-
thelium had been degraded by zinc sulphate solution
(Holland et al., 2009). The lead author himself agreed
that it is difficult to imagine stable olfactory gradients
enabling navigation at this spatial scale, and called
these results surprising (Holland, 2014).
One should also mention the so-called activation
hypothesis, which suggest that smells per se are not
used by homing pigeons for navigation, but instead
activate some “independent non-olfactory naviga-
tional system” (Jorge et al., 2009, 2010). According to
these authors, any strong unusual olfactory stimuli,
including obviously artificial ones and not carrying
any navigational information (e.g. lavender or euca-
lyptus smells), may “activate” the navigational system
of pigeons. This hypothesis triggered a discussion
(Gould, 2009; Gagliardo et al., 2011; for a review see
Wallraff, 2014); in our opinion, its weaknesses have
been analysed in sufficient detail in the aforemen-
tioned review, to which an interested reader is referred.
Magnetic Map
The magnetic map hypothesis is, in our opinion, a
well-founded one. It should however be emphasised
that even though this hypothesis has been repeatedly
discussed by different authors (Gould, 1989; Phillips,
1996; Walker et al., 2002; Freake et al., 2006; Kishki-
nev, 2015) and is believed by many non-specialists in
animal navigation to be an established fact, at least in
respect to migrating birds it is still (as of yet) a hypoth-
esis, and not a theory proven beyond reasonable
doubt. A crucial piece of evidence could be data of an
experiment with virtual magnetic displacement, in
which the birds kept in the artificial magnetic field corre-
sponding to the natural field parameters in a remote
location show a predictable orientation response to the
magnetic stimulus, e.g. re-orientation consistent with
the compensation of virtual displacement. The mag-
netic field should be the only parameter altered, and
all other potential navigational cues (smells, photope-
riod, access to celestial cues, etc.) should be the same.
Such data have been recently obtained by us (Kishki-
nev et al., 2015). Disorientation of birds in response to
magnetic displacement, obtained in some experiments
(Fischer et al., 2003; Deutschlander et al., 2012), can be
claimed to be a non-specific response and cannot be
treated as hard evidence in favour of a magnetic map.
The geomagnetic map hypothesis assumes that
migrating birds use a gradient map which is based on
the regular variation of the geomagnetic field parame-
ters: total intensity and magnetic inclination. Both
these parameters generally vary from the poles to the
equator: total field intensity is the highest (ca. 60000 nT)
at the geomagnetic poles and the lowest (ca. 30000 nT,
locally down to 24000 nT) at the magnetic equator;
inclination is +90° at the northern magnetic pole, 0°
at the magnetic equator (i.e., field lines are parallel to
the surface of the Earth; which is the definition of the
magnetic equator) and –90° at the southern magnetic
pole (Skiles, 1989). However this is a very broad
description; in reality the isolines of total intensity and
inclination do not run exactly parallel to the geo-
graphic parallels and to each other (Fig. 2). In some
regions (e.g. in the Indian Ocean) isolines of both
parameters form a grid usable for bicoordinate purely
magnetic navigation (north—south and east—east);
however, in other large areas, e.g. in southern North
America and in Australia, they run parallel to each
other and provide no east—west information
(Boström et al., 2012).
Therefore, a geomagnetic map based on total
intensity of the field and on inclination, may provide a
bicoordinate navigation grid in some regions but not
everywhere on Earth. If the birds can detect not only
total intensity and inclination, but also declination,
the longitude problem (east—west axis) will be com-
pletely solved, and bicoordinate navigation provided
(Skiles, 1989; Fig. 1). In principle, migratory birds
have a sensory basis for measuring declination, and
they should use a magnetic and a celestial (solar or
stellar) compass and compare the readings. In this
case the map is not purely geomagnetic, but magnetic-
celestial, as one of the gradients is measured by an
astronomic compass. However, the ability to detect
declination and to use it as a map element remains to
be demonstrated in migratory birds.
Unlike the magnetic declination, the ability of
birds to detect total intensity and inclination has been
shown. Operant conditioning experiments showed
that pigeons can respond to simultaneous strong
change in intensity and inclination (Mora et al.,
2004), when this information is transmitted to the
brain via V1, and also to the change in inclination
alone (Mora et al., 2014). Similar results have been
obtained in Pekin ducks (Anas platyrhynchos f. domes-
tica; Freire et al., 2012). We were able to show that the
ability of Eurasian reed warblers (Acrocephalus scirpa-
ceus) to compensate for the physical displacement by
1000 km along longitude is critically dependent of
receiving V1-transmitted information (Kishkinev et al.,
2013). Intact reed warblers are able compensate for the
displacement (Chernetsov et al., 2008), and also
sham-operated birds, whereas V1-ablated individuals
796
BIOLOGY BULLETIN Vol. 43 No. 8 2016
CHERNETSOV
continued to show the same orientation as they had
done before the displacement (Kishkinev et al., 2013).
Even though we did not show the nature of the infor-
mation transmitted through V1, indirect evidence sug-
gests that this information is magnetic. Taken
together, these data in our opinion strongly suggest
that (1) migrating birds can perceive information on
intensity and direction of the geomagnetic field;
(2) this information is transmitted to the brain via V1;
(3) this information is used for position finding at the
scale of hundreds and thousands of kilometres. Clear
evidence of the existence of a magnetic (or magnetic-
celestial, see above) map in migratory birds would be
shown by successful and repeatable experiments with
virtual magnetic displacement with a predictable ori-
entation response, similar to the response to the real
displacement (Kishkinev et al., 2015).
It should be noted that the structure of the putative
V1-innervated magnetoreceptors remains unknown
(a review: Kishkinev and Chernetsov, 2015). However,
there is indirect evidence that such receptors exist
(Mouritsen and Hore, 2012) and transmit information
on intensity and inclination of the magnetic field
(Mora et al., 2004; Heyers et al., 2010; Lefeldt et al.,
2014), which might be used for long-distance naviga-
tion (Kishkinev et al., 2013). However, the structure
and the exact location of these receptors remain
unknown; previously suggested opinions on that topic
appeared to be erroneous (Treiber et al., 2012).
If the geomagnetic map exists, the question of its
spatial accuracy inevitably follows. Difference
between total field intensity at the poles and at the
equator is ca. 30000 nT; the distance between the pole
and the equator is ca. 10000 km. Thus, the mean gra-
dient of field intensity is 3 nT/km. At the same time,
daily variation of the field is tens of nT, and magnetic
storms can alter the field by hundreds and in excep-
tional cases up to 1000 nT (Skiles, 1989). This means
that the spatial accuracy of the magnetic map is ca.
25–50 km (at smaller scale noise becomes stronger
than signal), whereas the accuracy of avian navigation
is considerably higher, roughly by an order of magni-
tude (Sokolov, 1991). Either migrating birds use rather
sophisticated algorithms of isolation of signal from
noise, or avian position finding is a hierarchical pro-
cess using a magnetic map at large scale and some
other mechanism at smaller scale.
Other Navigation Systems
Apart from the putative olfactory and magnetic (or
magnetic-celestial) maps, other positioning cues have
been suggested as sources of information for avian
maps. At an early stage, the proposal that birds use the
Coriolis force was forwarded (Yeagley and Whitemore,
1947). This suggestion has not been treated seriously
for a while and now belongs in history of science.
Anther exotic hypothesis, namely that pigeons use
gravitational anomalies for navigation, is still being
Fig. 2. Map with isolines of total intensity of the geomagnetic field (isodynamic lines, light grey, in μT) and of magnetic inclina-
tion (isoclines, dark grey, in degrees). Mercator map projection, revised from Boström et al. (2012). Isodynamic lines and iso-
clines may form a good grid (e.g. in Siberia or in the Indian Ocean), or may run nearly parallel to each other (North America,
Australia).
10
55
60
80
40
40
60
20
25
20
30
35
40
40
80
60
55
50
35
45
50
55
60
−80
65
0
BIOLOGY BULLETIN Vol. 43 No. 8 2016
ORIENTATION AND NAVIGATION OF MIGRATING BIRDS 797
debated (Larkin and Keeton, 1978; Kanevskyi et al.,
1985; Blaser et al., 2013, 2014). The idea is that birds
can detect heterogeneities in the gravitational field of
the Earth generated by varying density of the rocks,
and use it as a source of map-related information (or
perceive it as a disturbance). The main issue with this
hypothesis is not that all evidence invoked in its sup-
port is purely correlational (considering the nature of
gravitational force, a controlled experiment is hardly
conceivable), and not even that using a gravitational
map would demand that pigeons should discriminate
the angles of the gravitational force direction differing
by several angular seconds. It is that to detect gravita-
tional anomalies, one should be able to sense the
downward direction (i.e., towards the centre of the
Earth’s gravitational mass) in a given point, and also
simultaneously compare it with the downward direc-
tion in a remote location. This ability is not provided
by the vestibular system, either avian or human; other-
wise we would have detected that “downward” direc-
tions in St. Petersburg and Vladivostok are much dif-
ferent, and upon disembarking a f light from Europe to
New Zealand, we would have felt that we walk upside
down. Lacking ability to detect the frozen direction
“downward at home,” independent from the local
downward direction in any animal, forces one to treat
the gravitational hypothesis as a curiosity, in spite of
recent publications in respectable journals (Blaser
et al., 2013, 2014).
Among the “non-traditional” navigation hypothe-
ses the most credible one looks at the use of infrasound
(Hagstrum, 2001, 2003, 2013). Some landscape struc-
tures, for instance mountain ranges and especially sea
coasts, generate the constant infrasound with frequen-
cies between 0.1–10 Hz, which might be used as a
source of spatial information by animals, including
birds that are able to perceive it. There are correlative
data that homing efficiency of pigeons is changed
during natural or anthropogenic infrasound distur-
bances (Hagstrum, 2000). The infrasound hypothesis
looks much less exotic than using Coriolis force or
gravitational anomalies. Infrasound can hardly be the
main source of positional information for long-dis-
tance migrants, because it does not form a grid, but its
use by some birds as one of local landscape cues is pos-
sible and does not violate the known laws of nature.
Finally, one of the first position-finding hypothe-
ses ever suggested was the astronomical navigation
concept, based on observations of apparent move-
ments of the Sun and the stars (Matthews, 1953; Sauer, F.,
1957; Sauer, E.G.F. and Sauer, E.M., 1960). These
hypotheses did not find support in subsequent studies.
It is currently accepted that both the Sun (Schmidt-
Koenig, 1961, 1990) and the stars (Mouritsen and
Larsen, 2001) provide migrating birds only with com-
pass information. In particular, to use the Sun or the
stars for true navigation the birds would need to detect
their position in the sky with a high precision, and it is
not clear whether they are capable of that (Hodos,
1993). Furthermore, experimental evidence does not
support the celestial navigation hypothesis.
Spatio-Temporal Programmes and Controlling
the Position on the Migratory Route
Since the classic experiments of Perdeck on star-
lings (Sturnus vulgaris) (Perdeck, 1958) it is usually
assumed that experienced adult migrants performing
migration not for the first time use a navigation map,
whereas juvenile birds migrating for the first time have
no map (in the Northern Hemisphere it usually hap-
pens in autumn). It is believed that first-autumn
migrants flying towards their winter quarters where
they have never been before, do not use a map, which
in birds is not innate, but experience-based, but follow
an innate spatio-temporal programme (Gwinner and
Wiltschko, 1978). This programme in the simplest way
may be formulated as “start migration at a certain age,
fly towards the direction α during n days, then fly
towards β during m days, etc.; in the end you have
reached your destination (winter area).” First-autumn
migrants are believed to have no idea in which part of
their route they currently are. This idea is called “clock
and compass” concept, even though a more accurate
name would be “calendar and compass.” The main
point is that it does not include an innate map.
This concept is the most influential and widely
held view currently. A number of facts agree with it:
experiments with white-crowned sparrows (Zonotri-
chia leucophrys) displaced from the northwestern to
the northeastern U.S. confirmed the ability of adult
experienced migrants to detect and compensate for a
longitudinal displacement and the inability of juve-
niles to do it (Thorup et al., 2007). The geographic
distribution pattern of long-distance recoveries of
ringed pied flycatchers is exactly the same as should be
expected if during their first autumn these birds use a
clock and a compass and do not compensate for the
inevitable orientation errors, wind drift, etc. (Mourit-
sen, 1998). Spatio-temporal programmes appear to be
inherited, and F1 hybrids show intermediate patterns
both temporally (duration of Zugunruhe; Berthold
and Querner, 1981; Berthold, 1988) and directionally
(Helbig, 1991, 1996; Berthold et al., 1992). Recently
these results obtained in lab tests in Emlen funnels
have been confirmed by tracking birds of hybrid origin
in the wild (Delmore and Irwin, 2014).
However, soon after E. Gwinner and W. Wiltschko
demonstrated that first-autumn garden warblers show
a preferred direction of spontaneous locomotor activ-
ity in round arenas corresponding to the migratory
direction in free-living conspecifics without external
information, on the basis of their inherited pro-
gramme alone (Gwinner and Wiltschko, 1978), Beck
and Wiltschko (1988) showed that juvenile pied fly-
catchers are not able to do that. The birds of the latter
species need the geomagnetic information corre-
sponding to their migratory route to successfully
798
BIOLOGY BULLETIN Vol. 43 No. 8 2016
CHERNETSOV
develop their innate spatio-temporal programme.
Later juvenile thrush nightingales (Luscinia luscinia)
from Sweden were shown to sharply increase their
fuelling rate if placed in the magnetic conditions of
northern Egypt, i.e., before a migratory flight across
the Sahara. The birds that remained in the magnetic
conditions of Sweden did not gain much fuel (Frans-
son et al., 2001). Importantly, these data refer to first-
autumn birds that had not yet experienced magnetic or
other conditions of Egypt.
It has been shown that first-autumn long-tailed tits
(Aegithalos caudatus) respond to decreasing day length
by increasing their locomotor activity, which in the
wild means a higher migratory speed in birds delayed
in northern areas (Bojarinova and Babushkina, 2015).
These and many other data suggest that apart from
clock (or calendar) and compass, naïve migrants have
some innate idea of the magnetic and/or photoperi-
odic conditions that they should experience during the
normal, according to schedule, movement along their
autumn migratory route, and can correct their migra-
tory behaviour if the schedule is not followed. It might
not be a full map, but rather a system of “beacons” or
“signposts” that allow first-autumn migrants to con-
trol their position on the migratory route. Satellite
tracks of juvenile birds, e.g. Eleonora falcons (Falco ele-
onorae) (Gschweng et al., 2008) and wandering alba-
trosses (Diomedea exulans) (Åkesson and Weimerskirch,
2014), are very difficult to explain by clock and com-
pass alone, without a map sense and at least some ele-
ments of an innate map.
In our opinion, sufficient evidence has been made
available to revise the clock-and-compass concept.
First-autumn migrants (possibly not of all species)
seem to be able to control their position to a varying
degree of accuracy on the basis of inherited informa-
tion. Which cues they use and the accuracy of their
control should be subjects of further studies. It is pos-
sible that in this respect, like in hierarchy of compass
systems, significant between-species variation will be
found, governed primarily by length and complexity of
migratory route. A chaffinch from the Baltic area
might be able to reach its species-specific wintering
area in southwestern Europe without an innate map,
whereas an Eleonora falcon hatched in the Mediterra-
nean may need such information to reach its non-
breeding quarters in Madagascar. It is even more diffi-
cult to imagine a Pacific golden plover (Pluvialis fulva)
breeding in Alaska and reaching its winter range in the
Hawaii without a map sense (Johnson et al., 1997).
The aforementioned considerations can be brief ly
summarised as follows. Migrating birds possess com-
pass systems that allow them to select and maintain
compass directions. At least three such systems are
currently known (solar, stellar and magnetic), even
though it cannot be ruled out that not all species and
populations are equally proficient in using all of them.
Details of each of these systems are not completely
clear and require further research, but their existence
is a rather well established fact. Hierarchy and interac-
tion of compass systems of migrants are poorly stud-
ied; quite possibly, avian species differ in this respect.
During migration birds learn how to use a map that
makes it possible to perform true navigation (i.e., to
detect their position in respect to the goal of migra-
tion) at the scale of their usual movements. The phys-
ical nature of the map is currently being actively stud-
ied; currently the most promising are the geomagnetic
and olfactory map hypotheses.
It is usually assumed that map is experience-based
and formed during migratory movements. This means
that the map exists only for the areas visited during the
life of an individuals, and possibly areas in their more
or less immediate vicinity. Migrating birds are believed
to have no innate map of the areas not visited earlier,
so that juveniles reach their species-specific wintering
areas for the first time by using their compass sense
and counting time, thus moving in a certain geneti-
cally inherited direction. However, in recent years evi-
dence becomes available that juveniles (possibly not of
all species of migrants) do have some mechanism
(possibly a crude one) of controlling their position on
migratory route, which allows them to compensate for
errors in their spatio-temporal programme to some
extent. This mechanism might be based on the param-
eters of the geomagnetic field.
Avian orientation and navigation research devel-
oped in the 1950s to early 1980s, after two main meth-
ods of behavioural experiments had been made avail-
able: lab tests in round arenas (Kramer, 1949) and
pigeon releases (homing; Matthews, 1951). Soviet
researchers participated in that research, in particular
in the studies of navigational map. An important study
that showed that in order to be able to return to their
natal area, migrating birds need to move across it (i.e.,
that they sample some physical gradients), was per-
formed on the Courish Spit by the members of the
Biological Station Rybachy of the Zoological Insti-
tute. Its results were published in 1984 in Zoologich-
eskii Zhurnal (Sokolov et al., 1984). In the late 1980s
and until the end of the 20th century, orientation
research declined somewhat, only to become more
active again since about 2000. The current phase is
characterised by an active collaboration between zool-
ogists, neurobiologists and biophysicists. Russian
researchers do participate in this development (Cher-
netsov et al., 2004, 2008, 2011; Kavokin, 2009;
Solov’yov et al., 2010; Kishkinev et al., 2010, 2013,
2015; Kavokin et al., 2014), but their contribution may
and should be more active.
ACKNOWLEDGMENTS
The author is sincerely grateful to A.L. Mukhin,
J.G. Bojarinova, L.V. Sokolov and an anonymous
reviewer whose comments contributed to improving
BIOLOGY BULLETIN Vol. 43 No. 8 2016
ORIENTATION AND NAVIGATION OF MIGRATING BIRDS 799
earlier drafts. This study was performed with support
from Russian Foundation for Basic Research (15-04-
05386) and with participation of the Zoological Insti-
tute of Russian Academy of Sciences (01201351182).
The study has also been supported by St. Petersburg
State University (1.37.149.2014).
REFERENCES
Able, K.P. and Able, M.A., Calibration of the magnetic
compass of a migratory bird by celestial rotation, Nature,
1990, vol. 347, no. 6291, pp. 378–380.
Able, K.P. and Able, M.A., Ontogeny of migratory orienta-
tion in the savannah sparrow, Passerculus sandwichensis:
calibration of the magnetic compass, Anim. Behav., 1990a,
vol. 39, no. 5, pp. 905–913.
Able, K.P. and Able, M.A., Daytime calibration of magnetic
orientation in a migratory bird requires a view of skylight
polarization, Nature, 1993, vol. 364, no. 6437, pp. 523–525.
Åkesson, S. and Weimerskirch, H., Evidence for sex-segre-
gated ocean distributions of first-winter wandering alba-
trosses at Crozet Islands, PLoS One, 2014, vol. 9, no. 2,
p. e86779.
Åkesson, S., Odin, C., Hegedüs, R., Ilieva, M., Sjöholm, C.,
Farkas, A., and Horváth, G., Testing avian compass calibra-
tion: comparative experiments with diurnal and nocturnal
passerine migrants in South Sweden, Biol. Open, 2015,
vol. 4, no. 1, pp. 35–47.
Alert, B., Michalik, A., Helduser, S., Mouritsen, H., and
Güntürkün, O., Perceptual strategies of pigeons to detect a
rotational centre—a hint for star compass learning?, PLoS
One, 2015, vol. 10, no. 3, p. e0119919.
Beck, W. and Wiltschko, W., Magnetic factors control the
migratory direction of pied flycatchers (Ficedula hypoleuca
Pallas), in Proceedings of 19th International Ornithological
Congress, 1988, pp. 1955–1962.
Benhamou, S., Bonadonna, F., and Jouventin, P., Success-
ful homing of magnet-carrying white-chinned petrels
released in the open sea, Anim. Behav., 2003, vol. 65, no. 4,
pp. 729–734.
Berthold, P., Evolutionary aspects of migratory behavior in
European warblers, J. Evol. Biol., 1988, vol. 1, no. 3,
pp. 195–209.
Berthold, P., Control of Bird Migration, London: Chapman
and Hall, 1996.
Berthold, P. and Querner, U., Genetic basis of migratory
behavior in European warblers, Science, 1981, vol. 212,
no. 4490, pp. 77–79.
Berthold, P., Helbig, A.J., Mohr, G., and Querner, U.,
Rapid microevolution of migratory behaviour in a wild bird
species, Nature, 1992, vol. 360, no. 6405, pp. 668–670.
Bingman, V.P., Night sky orientation of migratory pied f ly-
catchers raised in different magnetic fields, Behav. Ecol.
Sociobiol., 1984, vol. 15, no. 1, pp. 77–80.
Bingman, V.P., Earth’s magnetism and the nocturnal ori-
entation of migratory European robins, Auk, 1987, vol. 104,
no. 3, pp. 523–525.
Blaser, N., Guskov, S.I., Meskenaite, V., Kanevskyi, V.A.,
and Lipp, H.-P., Altered orientation and flight parts of
pigeons reared on gravity anomalies: a GPS tracking study,
PLoS One, 2013, vol. 8, no. 10, p. e77102.
Blaser, N., Guskov, S.I., Entin, V.A., Wolfer, D.P.,
Kanevskyi, V.A., and Lipp, H.-P., Gravity anomalies with-
out geomagnetic disturbances interfere with pigeon hom-
ing—a GPS tracking study, J. Exp. Biol., 2014, vol. 217,
no. 22, pp. 4057–4067.
Bojarinova, J. and Babushkina, O., Photoperiodic condi-
tions affect the level of locomotory activity during migration
in the long-tailed tit (Aegithalos C. caudatus), Auk, 2015,
vol. 132, no. 2, pp. 370–379.
Bonadonna, F., Chamaille-Jammes, S., Pinaud, D., and
Weimerskirch, H., Magnetic cues: are they important in
black-browed albatross Diomedea malanophris orientation?,
Ibis, 2003, vol. 145, no. 1, pp. 152–155.
Boström, J.E., Åkesson, S., and Alerstam, T., Where on
earth can animals use a geomagnetic bi-coordinate map for
navigation?, Ecography, 2012, vol. 35, no. 11, pp. 1039–
1047.
Burda, H., Marhold, S., Westenberger, T., Wiltschko, R.,
and Wiltschko, W., Magnetic compass orientation in the
subterranean rodent Cryptomys hottentotus (Bathyergidae,
Rodentia), Experientia, 1990, vol. 46, no. 5, pp. 528–530.
Bushuev, A.V., Husby, A., Sternberg, H., and Grinkov, V.G.,
Quantitative genetics of basal metabolic rate and body mass
in free-living pied flycatchers, J. Zool., 2012, vol. 288, no. 4,
pp. 245–251.
Chernetsov, N., Avian compass systems: do all migratory
species possess all three?, J. Avian Biol., 2015, vol. 46, no. 4,
pp. 342–343.
Chernetsov, N., Berthold, P., and Querner, U., Migratory
orientation of first-year white storks (Ciconia ciconia):
inherited information and social interactions, J. Exp. Biol.,
2004, vol. 207, no. 6, pp. 937–943.
Chernetsov, N., Kishkinev, D., and Mouritsen, H., A long-
distance avian migrant compensates for longitudinal dis-
placement during spring migration, Curr. Biol., 2008,
vol. 18, no. 3, pp. 188–190.
Chernetsov, N., Kishkinev, D., Kosarev, V., and Bolsha-
kov, C.V., Not all songbirds calibrate their magnetic com-
pass from twilight cues: a telemetry study, J. Exp. Biol.,
2011, vol. 214, no. 15, pp. 2540–2543.
Cochran, W.W., Mouritsen, H., and Wikelski, M., Migrat-
ing songbirds recalibrate their magnetic compass daily from
twilight cues, Science, 2004, vol. 304, no. 5669, pp. 405–
408.
Delmore, K.E. and Irwin, D.E., Hybrid songbirds employ
intermediate routes in a migratory divide, Ecol. Lett., 2014,
vol. 17, no. 10, pp. 1211–1218.
Delmore, K.E., Fox, J.W., and Irwin, D.E., Dramatic
intraspecific differences in migratory routes, stopover sites
and wintering areas, revealed using light-level geolocators,
Proc. Roy. Soc., 2012, vol. 279, no. 1747, pp. 4582–4589.
Deutschlander, M.E., Borland, S.C., and Phillips, J.B.,
Extraocular magnetic compass in newts, Nature, 1999,
vol. 400, no. 6742, pp. 324–325.
Deutschlander, M.E., Freake, M.J., Borland, S.C., Phil-
lips, J.B., Madden, R.C., Anderson, L.E., and Wilson, B.W.,
Learned magnetic compass orientation by the Siberian
hamster, Phodopus sungorus, Anim. Behav., 2003, vol. 65,
no. 4, pp. 779–786.
Deutschlander, M.E., Phillips, J.B., and Munro, U., Age-
dependent orientation to magnetically-simulated magnetic
displacements in migratory Australian silvereyes (Zosterops
l. lateralis), Wilson J. Ornithol., 2012, vol. 124, no. 3,
pp. 467–477.
Diego-Rasilla, F.J., Luengo, R.M., and Phillips, J.B.,
Light-dependent magnetic compass in Iberian green frog
tadpoles, Naturwissenschaften, 2010, vol. 97, no. 1,
pp. 1077–1088.
800
BIOLOGY BULLETIN Vol. 43 No. 8 2016
CHERNETSOV
Diego-Rasilla, F.J., Luengo, R.M., and Phillips, J.B., Use
of a light-dependent magnetic compass for y-axis orienta-
tion in European common frog (Rana temporaria) tadpoles,
J. Comp. Physiol. A, 2013, vol. 199, no. 7, pp. 619–628.
Dolnik, V.R., Astronomical orientation of birds, in Orien-
tatsiya ptits i territorial’nye svyazi populyatsii ptits (Orienta-
tion of Birds and Territorial Relationships of Bird Popula-
tions), Riga: Zinatne, 1973, pp. 14–61.
Dolnik, V.R., Resursy energii i vremeni u ptits v prirode (Energy
and Time Resources of Birds in Nature), St. Petersburg:
Nauka, 1995.
Duff, S.J., Brownlie, L.A., Sherry, D.F., and Sangster, M.,
Sun compass and landmark orientation by black-capped
chickadees (Parus atricapillus), J. Exp. Phychol.: Anim.
Behav. Proc., 1998, vol. 24, no. 3, pp. 243–253.
Emlen, S.T., Migratory orientation in the indigo bunting,
Passerina cyanea. Part I: Evidence of use of celestial cues,
Auk, 1967, vol. 84, no. 3, pp. 309–341.
Emlen, S.T., Migratory orientation in the indigo bunting,
Passerina cyanea. Part II: Mechanism of celestial orienta-
tion, Auk, 1967a, vol. 84, no. 4, pp. 463–489.
Emlen, S.T., Celestial rotation: its importance in the devel-
opment of migratory orientation, Science, 1970, vol. 170,
no. 3963, pp. 1198–1201.
Emlen, S.T., Wiltschko, W., Demong, N.J., Wiltschko, R.,
and Bergman, S., Magnetic direction finding: evidence for
its use in migratory indigo buntings, Science, 1976, vol. 193,
no. 4252, pp. 505–508.
Engels, S., Schneider, N.-L., Lefeldt, N., Hein, C.M.,
Zapka, M., et al., Anthropogenic electromagnetic noise
disrupts magnetic compass orientation in a migratory bird,
Nature, 2014, vol. 509, no. 7500, pp. 353–356.
Fischer, J.H., Munro, U., and Phillips, J.B., Magnetic nav-
igation by an avian migrant?, in Avian Migration, Berlin:
Springer, 2003, pp. 423–432.
Fransson, T., Jakobsson, S., Johansson, P., Kullberg, C.,
Lind, J., and Vallin, A., Magnetic cues trigger extensive
refuelling, Nature, 2001, vol. 414, no. 6859, pp. 35–36.
Freake, M.J., Muheim, R., and Phillips, J.B., Magnetic
maps in animals: a theory comes of age?, Quart. Rev. Biol.,
2006, vol. 81, no. 4, pp. 327–347.
Freire, R., Dunston, E., Fowler, E.M., McKenzie, G.L.,
Quinn, C.T., and Michelsen, J., Conditioned response to a
magnetic anomaly in the Pekin duck (Anas platyrhynchos
domestica) involves the trigeminal nerve, J. Exp. Biol., 2012,
vol. 215, no. 14, pp. 2399–2404.
Gaggini, V., Baldaccini, N., Spina, F., and Giunchi, D.,
Orientation of the pied flycatcher Ficedula hypoleuca: cue-
conflict experiments during spring migration, Behav. Ecol.
Sociobiol., 2010, vol. 64, no. 8, pp. 1333–1342.
Gagliardo, A., Forty years of olfactory navigation in birds,
J. Exp. Biol., 2013, vol. 216, no. 12, pp. 2165–2171.
Gagliardo, A., Ioalè, P., Savini, M., and Wild, J.M., Having
the nerve to home: trigeminal magnetoreceptor versus
olfactory mediation of homing in pigeons, J. Exp. Biol.,
2006, vol. 209, no. 15, pp. 2888–2892.
Gagliardo, A., Ioalè, P., Savini, M., and Wild, M., Naviga-
tional abilities of homing pigeons deprived of olfactory or
trigeminally mediated magnetic information when young,
J. Exp. Biol., 2008, vol. 211, no. 13, pp. 2046–2051.
Gagliardo, A., Ioalè, P., Savini, M., and Wild, M., Naviga-
tional abilities of adult and experienced homing pigeons
deprived of olfactory or trigeminally mediated magnetic
information, J. Exp. Biol., 2009, vol. 212, no. 19, pp. 3119–
3124.
Gagliardo, A., Ioalè, P., Filannino, C., and Wikelski, M.,
Homing pigeons only navigate in air with intact environ-
mental odours: a test of the olfactory activation hypothesis
with GPS data loggers, PLoS One, 2011, vol. 6, no. 8,
p. e22385.
Gagliardo, A., Bried, J., Lambardi, P., Luschi, P., Wikelski, M.,
and Bonadonna, F., Oceanic navigation in Cory’s shearwa-
ters: evidence for a crucial role of olfactory cues for homing
after displacement, J. Exp. Biol., 2013, vol. 216, no. 15,
pp. 2798–2805.
Gill, R.E., Jr., Tibbits, T.L., Douglas, D.C., Handel, C.M.,
Mulcahy, D.M., et al., Extreme endurance flights by land-
birds crossing the Pacific Ocean: ecological corridor rather
than barrier?, Proc. Roy. Soc., 2009, no. 1656, pp. 447–457.
Gould, J.L., Do animals have magnetic maps?, in Biogennyi
magnetit i magnitoretseptsiya. Novoe o biomagnetizme (Bio-
genic Magnetite and Magnetoreception. New in Biomagne-
tism), Moscow: Mir, 1989, vol. 1, pp. 334–349.
Gould, J.L., Animal navigation: a wake-up call for homing,
Curr. Biol., 2009, vol. 19, no. 8, pp. R338–R339.
Gschweng, M., Kalko, E.K.V., Querner, U., Fielder, W.,
and Berthold, P., All across Africa: highly individual migra-
tion routes of Eleonora’s falcon, Proc. Roy. Soc., 2008,
no. 1653, pp. 2887–2896.
Guinchi, D., Vanni, L., Baldaccini, N.E., Spina, F., and
Biondi, F., New cue-conflict experiments suggest a leading
role of visual cues in the migratory orientation of pied flycatch-
ers Ficedula hypoleuca, J. Ornithol., 2015, vol. 156, no. 1,
pp. 113–121.
Gwinner, E. and Wiltschko, W., Endogenously controlled
changes in migratory direction of the garden warbler, Sylvia
borin, J. Comp. Physiol. A, 1978, vol. 125, no. 3, pp. 267–
273.
Hagstrum, J.T., Infrasound and the avian navigational
map, J. Exp. Biol., 2 000 , v o l. 2 0 3 , n o . 7, p p . 11031111.
Hagstrum, J.T., Infrasound and the avian navigational
map, J. Navigat., 2001, vol. 54, no. 3, pp. 377–391.
Hagstrum, J.T., Atmospheric propagation modeling indi-
cates homing pigeons use loft-specific infrasonic ‘map’
cues, J. Exp. Biol., 2013, vol. 216, no. 4, pp. 687–699.
Helbig, A.J., Inheritance of a migratory direction in a bird
species: a cross-breeding experiment with SE- and SW-
migrating blackcaps (Sylvia atricapilla), Behav. Ecol. Socio-
biol., 1991, vol. 28, no. 1, pp. 9–12.
Helbig, A.J., Genetic basis, mode of inheritance and evolu-
tionary changes of migratory directions in palearctic war-
blers (Aves: Sylviidae), J. Exp. Biol., 1996, vol. 199, no. 1,
pp. 49–55.
Heyers, D., Zapka, M., Hoffmeister, M., Wild, J.M., and
Mouritsen, H., Magnetic field changes activate the trigem-
inal brainstem complex in a migratory bird, Proc. Natl.
Acad. Sci. U. S. A., 2010, vol. 107, no. 20, pp. 9394–9399.
Hodos, W., The visual capabilities of birds, in Vision, Brain,
and Behavior in Birds, Cambridge, Mass.: MIT Press, 1993,
pp. 63–76.
Holland, R.A., True navigation in birds: from quantum
physics to global migration, J. Zool., 2014, vol. 293, no. 1,
pp. 1–15.
Holland, R.A., Thorup, K., Vonhof, M.J., Cochran, W.W.,
and Wikelski, M., Bat orientation using Earth’s magnetic
field, Nature, 2006, vol. 444, no. 7120, p. 702.
BIOLOGY BULLETIN Vol. 43 No. 8 2016
ORIENTATION AND NAVIGATION OF MIGRATING BIRDS 801
Holland, R.A., Thorup, K., Gagliardo, A., Bisson, I.,
Knecht, E., et al., Testing the role of sensory systems in the
migratory heading of a songbird, J. Exp. Biol., 2009,
vol. 212, no. 24, pp. 4065–4071.
Johnson, O.W., Warnock, N., Bishop, M.A., Bennett, A.J.,
Johnson, P.M., and Kienholz, R.J., Migration by radio-
tagged Pacific golden-plovers from Hawaii to Alaska, and
their subsequent survival, Auk, 1997. vol. 114, no. 3, pp.
521–524.
Jorge, P.E., Marques, P.A.M., and Phillips, J.B., Activa-
tional effects of odours on avian navigation, Proc. Roy. Soc.,
2010, no. 1678, pp. 45–49.
Jorge, P.E., Marques, A.E., and Phillips, J.B., Activational
rather than navigational effects of odors on homing of
young pigeons, Curr. Biol., 2009, vol. 19, no. 8, pp. 650–
654.
Kanevskii, V.A., Sytnik, K.M., Shelyag-Sosonko, Yu.R.,
Melnikov, D.I., Dyma, A.G., et al., Application of biote-
lemetry in remote geophysical studies, Dokl. Akad. Nauk
SSSR, 1985, vol. 282, no. 2, pp. 291–294.
Katz, E.B. and Mihelson, Kh.A., Effect of changes in the
direction of stellar and magnetic meridians on the orienta-
tion of robins in round cells in a planetarium, in Orientatsiya
ptits (Orientation of Birds), Riga: Zinatne, 1978, pp. 180–
193.
Kavokin, K.V., The puzzle of magnetic resonance effect on
the magnetic compass of migratory birds, Bioelectromagnet-
ics, 2009, vol. 30, no. 5, pp. 402–410.
Kavokin, K., Chernetsov, N., Pakhomov, A., Bojarinova, J.,
Kobylkov, D., and Namozov, B., Magnetic orientation of
garden warblers (Sylvia borin) under 1.4 MHz radiofre-
quency magnetic field, J. Roy. Soc. Interface, 2014, vol. 11,
no. 97, p. 20140451.
O’Keefe, J. and Nadel, L., The Hippocampus As a Cognitive
Map, Oxford: Oxford University Press, 1978.
Kishkinev D., Sensory mechanisms of long-distance navi-
gation in birds: a recent advance in the context of previous
studies, J. Ornithol., 2015, vol. 156 (Suppl. 1), pp. S145–
S161. doi 10.1007/s10336-015-1215-4
Kishkinev, D.A. and Chernetsov, N.S., Magnetoreception
systems in birds: a review of current research, Biol. Bull.
Rev., 2015, vol. 5, no. 1, pp. 46–62.
Kishkinev, D., Chernetsov, N., and Mouritsen, H., A dou-
ble-clock or jetlag mechanism is unlikely to be involved in
detection of east-west displacements in a long-distance
avian migrant, Auk, 2010, vol. 127, no. 4, pp. 773–780.
Kishkinev, D., Chernetsov, N., Heyers, D., and Mourit-
sen, H., Migratory reed warblers need intact trigeminal
nerves to correct for a 1,000 km eastward displacement,
PLoS One, 2013, vol. 8, no. 6, p. e65847.
Kishkinev, D., Chernetsov, N., Pakhomov, A., Heyers, D.,
and Mour itsen, H., Eurasian reed warblers compensate for vir-
tual magnetic displacement, Curr. Biol., 2015, vol. 25, no. 19,
pp. R822–R824.
Kramer, G., Über Richtungstendenzen bei der nächtlichen
Zugunruhe gekäfigter Vögel, in Ornithologie als biologische
Wissenschaft, Heidelberg: Carl Winter Verlag, 1949, pp.
269–283.
Kramer, G., Eine neue Methode zur Erforschung der
Zugorientierung und die bisher damit erzielten Ergebnisse,
in Proc. X Ornithol. Congr., Uppsala, 1951, pp. 269–280.
Kramer, G., Wird die Sonnenhöhe bei der Heimfindeori-
entierung verwertet?, J. Ornithol., 1953, vol. 94, nos. 3/4,
pp. 201–219.
Kramer, G., Experiments in bird orientation and their
interpretation, Ibis, 1957, vol. 99, no. 2, pp. 196–227.
Kramer, G., Long-distance orientation, in Biology and
Comparative Physiology of Birds, New York: Academic
Press, 1961, pp. 341–371.
Krylov, V.V., Osipova, E.A., and Izyumov, Yu.G., Orienta-
tion behavior of animals in the geomagnetic field and mech-
anisms of magnetoreception, Geofiz. Prots. Biosfera, 2014,
vol. 13, no. 4, pp. 60–82.
Larkin, T. and Keeton, W.T., An apparent lunar rhythm in
the day-today variations in initial bearings of homing
pigeons, in Animal Migration, Navigation and Homing, Ber-
lin: Springer, 1978, pp. 92–106.
Lefeldt, N., Heyers, D., Schneider, N.-L., Engels, S.,
Elbers, D., and Mouritsen, H., Magnetic field-driven
induction of ZENK in the trigeminal system of pigeons
(Columba livia), J. Roy. Soc. Interface, 2014, vol. 11, no. 100 ,
p. 20140777.
Liu, X. and Chernetsov, N., Avian orientation: multi-cue
integration and calibration of compass systems, Chinese
Birds, 2012, vol. 3, no. 1, pp. 1–8.
Lohmann, K.J., Magnetic orientation by hatchling logger-
head sea turtles (Caretta caretta), J. Exp. Biol., 1991,
vol. 155, no. 1, pp. 37–49.
Lohmann, K.J. and Lohmann, C.M.F., A light-indepen-
dent magnetic compass in the leatherback sea turtle, Biol.
Bull., 1993, vol. 185, no. 1, pp. 149–151.
Löhrl, H., Zur Frage des Zeitpunktes einer Prägung auf die
Heimatregion beim Halsbandschnäpper (Ficedula albicol-
lis), J. Ornithologie, 1959, vol. 100, no. 2, pp. 132–140.
Matthews, G.V.T., The experimental investigation of navi-
gation in homing pigeons, J. Exp. Biol., 1951, vol. 28, no. 4,
pp. 508–536.
Matthews, G.V.T., Sun navigation in homing pigeons, J.
Exp. Biol., 1953, vol. 30, no. 2, pp. 243–267.
Merkel, F.W. and Wiltschko, W., Magnetismus und Rich-
tungsfinden zugunruhiger Rotkehlchen (Erithacus rubec-
ula), Vogelwarte, 1965, vol. 23, no. 1, pp. 71–77.
Mewaldt, L.R., Winter philopatry in white-crowned spar-
rows (Zonotrichia leucophrys), North American Bird Bander,
1976, vol. 1, no. 1, pp. 14–20.
Michalik, A., Alert, B., Engels, S., Lefeldt, N., and Mou-
ritsen, H., Star compass learning: how long does it take?, J.
Ornithol., 2014, vol. 155, no. 1, pp. 225–234.
Moore, F.R., Integration of environmental stimuli in the
migratory orientation of the savannah sparrow, Passerculu s
sandwichensis, Anim. Behav., 1985, vol. 33, no. 2, pp. 654–
663.
Mora, C.V., Davison, M., Wild, J.M., and Walker, M.M.,
Magnetoreception and its trigeminal mediation in the hom-
ing pigeon, Nature, 2004, vol. 432, no. 7016, pp. 508–511.
Mora, C.V., Acerbi, M.L., and Bingman, V.P., Conditioned
discrimination of magnetic inclination in a spatial-orienta-
tion arena task by homing pigeons (Columba livia), J. Exp.
Biol., 2014, vol. 217, no. 23, pp. 4123–4131.
Mouritsen, H., Modelling migration: the clock-and-com-
pass model can explain the distribution of ringing recover-
ies, Anim. Behav., 1998, vol. 56, no. 4, pp. 899–907.
Mouritsen, H., The magnetic senses, in Neurosciences—
From Molecule to Behavior. A University Textbook, Berlin:
Springer, 2013, pp. 427–443.
Mouritsen, H., Magnetoreception in birds and its use for
long-distance migration, in Sturkie’s Avian Physiology, 6th ed.,
London: Academic Press, 2015, pp. 113–133.
802
BIOLOGY BULLETIN Vol. 43 No. 8 2016
CHERNETSOV
Mouritsen, H. and Hore, P.J., The magnetic retina: light-
dependent and trigeminal magnetoreception in migratory
birds, Curr. Opin. Neurobiol., 2012, vol. 22, no. 2, pp. 343–
352.
Mouritsen, H. and Larsen, O.N., Migrating songbirds
tested in computer-controlled Emlen funnels use stellar
cues for a time-independent compass, J. Exp. Biol., 2001,
vol. 204, no. 22, pp. 3855–3865.
Mouritsen, H., Huyvaert, K.P., Frost, B.J., and Anders on, D.J.,
Waved albatrosses can navigate with strong magnets
attached to their head, J. Exp. Biol., 2003, vol. 206, no. 22,
pp. 4155–4166.
Muheim, R., Moore, F.R., and Phillips, J.B., Calibration of
magnetic and celestial compass cues in migratory birds—a
review of cue-conflict experiments, J. Exp. Biol., 2006,
vol. 209, no. 1, pp. 2–17.
Muheim, R., Phillips, J.B., and Åkesson, S., Polarized light
cues underlie compass calibration in migratory songbirds,
Science, 2006a, vol. 313, no. 5788, pp. 837–839.
Muheim, R., Åkesson, S., and Phillips, J.B., Magnetic
compass of migratory savannah sparrows is calibrated by
skylight polarization at sunrise and sunset, J. Ornithol.,
2007, vol. 148, suppl. 2, pp. 485–494.
Muheim, R., Åkesson, S., and Phillips, J.B., Response to
R. Wiltschko et al. (J. Ornithol.): contradictory results on
the role of polarized light in compass calibration in migra-
tory songbirds, J. Ornithol., 2008, vol. 149, no. 4, pp. 659–
662.
Mukhin, A., Nocturnal restlessness in caged juvenile reed
warblers Acrocephalus scirpaceus, Avian Ecol. Behav., 1999,
vol. 3, pp. 91–97.
Mukhin, A.L. and Kobylkov, D.S., Homing by night—it is
better to lose the day and then to fly for half an hour, in Ori-
entatsiya i navigatsiya zhivotnykh. Tezisy nauch. konf. (Ori-
entation and Navigation of Animals. Abstracts of Scientific
Conference), Moscow: Tovar. Nauch. Izd. KMK, 2014,
p. 42.
Mukhin, A., Grinkevich, V., and Helm, B., Under cover of
darkness: nocturnal life of diurnal birds, J. Biol. Rhythms,
2009, vol. 24, no. 3, pp. 225–231.
Payevsky, V.A., Atlas of bird migration according to band-
ing data on the Curonian Spit, in Ekologicheskie i fiziolog-
icheskie aspekty pereletov ptits (Ecological and Physiological
Aspects of Bird Migrations), Leningrad: Nauka, 1971,
pp. 3–110.
Pakhomov, A. and Chernetsov, N., Early evening activity of
migratory garden warbler Sylvia borin: compass calibration
activity?, J. Ornithol., 2014, vol. 155, no. 3, pp. 621–630.
Papi, F., Fiore, L., Fiaschi, V., and Benvenuti, S., The
influence of olfactory nerve section on the homing capacity
of carrier pigeons, Monitore Zoologico Italiano N.S., 1971,
vol. 5, pp. 265–267.
Papi, F., Fiore, L., Fiaschi, V., and Benvenuti, S., Olfaction
and homing in pigeons, Monitore Zoologico Italiano N.S.,
1972, vol. 6, pp. 85–95.
Perdeck, A.C., Two types of orientation in migrating star-
lings, Sturnus vulgaris L., and chaffinches, Fringilla coelebs L.,
as revealed by displacement experiments, Ardea, 1958,
vol. 46, no. 1, pp. 1–37.
Phillips, J.B., Magnetic navigation, J. Theor. Biol., 1996,
vol. 180, no. 4, pp. 309–319.
Quinn, T.P., Merrill, R.T., and Brannon, E.L., Magnetic
field detection in sockeye salmon, J. Exp. Zool., 1981,
vol. 217, no. 1, pp. 137–142.
Ritz, T., Adem, S., and Schulten, K., A model for photore-
ceptor-based magnetoreception in birds, Biophys. J., 2000,
vol. 78, no. 2, pp. 707–718.
Ruegg, K., Anderson, E.C., Boone, J., Pouls, J., and
Smith, T.B., A role for migration-linked genes and genomic
islands in divergence of a songbird, Mol. Ecol., 2014, vol. 23,
no. 19, pp. 4757–4769.
Rüppell, W., Versuche zur Ortstreue und Fernorientierung
der Vögel. III. Heimfindeversuche mit Rauchschwalben
(Hirundo rustica) und Mehlschwalben (Delichon urbica) von
H. Warnat (Berlin-Charlottenburg), Vogelzug, 1934, vol. 5,
pp. 161–166.
Rüppell, W., Heimfindeversuche mit Staren 1934, J. Orni-
thologie, 1935, vol. 83, no. 3, pp. 462–524.
Salewski, V., Bairlein, F., and Leisler, B., Recurrence of
some palaearctic migrant passerine species in West Africa,
Ring. Migr., 2000, vol. 20, no. 1, pp. 29–30.
Sandberg, R., Bäckman, J., Moore, F.R., and Lõhmus, M.,
Magnetic information calibrates celestial cues during
migration, Anim. Behav., 2000, vol. 60, no. 4, pp. 453–462.
Sauer, F., Die Sternenorientierung nächtlich Ziehender
Grasmücken (Sylvia atricapilla, borin und curruca),
Zeitschrift fur Tierpsychologie, 1957, vol. 14, no. 1, pp. 29–
70.
Sauer, F., Zugorientierung einer Mönchsgrasmücke (Sylvia
a. atricapilla, L.) unter künstlichem Sternenhimmel, Natur-
wissenschaften, 1956, vol. 43, no. 10, pp. 231–232.
Sauer, E.G.F. and Sauer, E.M., Star navigation of noctur-
nal migrating birds. The 1958 planetarium experiments,
Cold Spring Harbor Symp. Quant. Biol., 1960, vol. 25,
pp. 463–473.
Schmaljohann, H., Rautenberg, T., Muheim, R., Naef-
Daenzer, B., and Bairlein, F., Response of a free-f lying
songbird to an experimental shift of the light polarization
pattern around sunset, J. Exp. Biol., 2013, vol. 216, no. 8,
pp. 1381–1387.
Schmidt-Koenig, K., Experimentelle Einflußname auf die
24-Stunden-Periodik bei Brieftauben und deren Auswirkun-
gen unter besonderer Berücksichtigung des Heimfindever-
mögens, Zeitschrift fur Tierpsychologie, 1958, vol. 15, no. 3,
pp. 301–331.
Schmidt-Koenig, K., Die Sonne als Kompaß im Heimori-
entierugnssystem der Brieftauben, Zeitschrift fur Tierpsy-
chologie, 1961, vol. 18, no. 2, pp. 221–244.
Schmidt-Koenig, K., The sun compass, Experientia, 1990,
vol. 46, no. 4, pp. 336–342.
Schulten, K., Swenberg, C.E., and Weller, A., A biomag-
netic sensory mechanism based on magnetic field modu-
lated coherent electron spin motion, Zeitschrift fur Physika-
lische Chemie (NF), 1978, vol. 111, no. 1, pp. 1–5.
Shumakov, M.E. and Zelenova, N.P., The ontogeny of
non-visual orientation of the Eurasian blackcap (Sylvia atri-
capilla), in Tezisy dokladov XII Pribaltiiskoi ornitol. konf.
(Abstracts of the XII Baltic Ornithol. Conf.), Vilnus, 1988,
pp. 255–257.
Skiles, D.D., The geomagnetic field, its nature, history and
importance for biology, in Biogennyi magnetit i magnit-
oretseptsiya. Novoe o biomagnetizme (Biogenic Magnetite
and Magnetoreception. New in Biomagnetism), Moscow:
Mir, 1989, vol. 1, pp. 63–144.
Sokolov, L.V., Vysotsky, V.G., and Bardin, A.V., The post-
breeding dispersion of the pied flycatcher (Ficedula hypo-
leuca) on the Curonian Spit, in Issledovaniya po faune i
ekologii ptits Palearktiki (Studies on the Fauna and Ecology
BIOLOGY BULLETIN Vol. 43 No. 8 2016
ORIENTATION AND NAVIGATION OF MIGRATING BIRDS 803
of Palearctic Birds), Trudy Zool. Inst. AN SSSR, Lenin-
grad, 1987, vol. 163, pp. 126–135.
Sokolov, L.V., Bolshakov, K.V., Vinogradova, N.V., Dol-
nik, T.V., Lyuleeva, D.S., et al., Checking the ability of
young finches to capture and find the future nesting terri-
tory, Zool. Zh., 1984, vol. 63, no. 11, pp. 1671–1681.
Sokolov, L.V., Filopatriya i dispersiya ptits (Phylopatry and
Dispersion of Birds), Trudy Zool. Inst. AN SSSR, Lenin-
grad, 1991, vol. 230.
Solov’yov, I., Mouritsen, H., and Schulten, K., Acuity of a
cryptochrome and vision-based magnetoreception system
in birds, Biophys. J., 2010, vol. 99, no. 1, pp. 40–49.
Thorup, K., Bisson, I.-A., Bowlin, M.S., Holland, R.A.,
Wingfield, J.C., et al., Evidence for a navigational map
stretching across the continental U.S. in a migratory song-
bird, Proc. Natl. Acad. Sci. U. S. A., 2007, vol. 104, no. 46,
pp. 18115–18119.
Tolman, E.C., Cognitive maps in rats and men, Psychol.
Rev., 1948, vol. 55, no. 4, pp. 189–208.
Treiber, C.D., Salzer, M.C., Riegler, J., Edelman, N.,
Sugar, C., et al., Clusters of iron-rich cells in the upper beak
of pigeons are macrophages not magnetosensitive neurons,
Nature, 2012, vol. 484, no. 7394, pp. 367–370.
Waldvogel, J.A., Olfactory navigation in homing pigeons:
are the current models atmospherically realistic?, Auk,
1987, vol. 104, no. 3, pp. 369–379.
Walker, M.M., Dennis, T.E., and Kirschvink, J.L., The
magnetic sense and its use in long-distance navigation by
animals, Curr. Opin. Neurobiol., 2002, vol. 12, no. 6,
pp. 735–744.
Wallraff, H.G., Does celestial navigation exist in animals?,
Cold Spring Harbor Symp. Quant. Biol., 1960, vol. 25,
pp. 451–461.
Wallraff, H.G., Das Navigationssystem der Vögel,
München–Wien: R. Oldenbourg, 1974.
Wallraff, H.G., Olfaction and homing in pigeons: nerve-sec-
tion experiments, critique hypotheses, J. Comp. Physiol. A,
1980, vol. 139, no. 3, pp. 209–224.
Wallraff, H.G., The olfactory component of pigeon naviga-
tion: steps and analysis, J. Comp. Physiol. A, 1981, vol. 143,
no. 4, pp. 411–422.
Wallraff, H.G., The magnetic map of homing pigeons: an
evergreen phantom, J. Theor. Biol., 1999, vol. 197, no. 2,
pp. 265–269.
Wallraff, H.G. and Andreae, M.O., Spatial gradients in
ratios of atmospheric trace gases: a study stimulated by
experiments on bird navigation, Tellus, 2000, no. 4,
pp. 1138–1157.
Wallraff, H.G., Avian Navigation: Pigeon Homing As a Par-
adigm, Berlin: Springer, 2005.
Wallraff, H.G., Do olfactory stimuli provide positional
information for home-oriented avian navigation?, Anim.
Behav., 2014, vol. 90, pp. e1–e6.
Wallraff, H.G., Ratios among atmospheric trace gases
together with winds imply exploitable information for bird
navigation: a model elucidating experimental results, Bio-
geosciences, 2013, vol. 10, no. 11, pp. 6929–6943.
Weindler, P., Beck, W., Liepa, V., and Wiltschko, W.,
Development of migratory orientation in pied flycatchers in
different magnetic inclinations, Anim. Behav., 1995, vol. 49,
no. 1, pp. 227–234.
Weindler, P., Baumetz, M., and Wiltschko, W., The direc-
tion of celestial rotation influences the development of stel-
lar orientation in young garden warblers (Sylvia borin), J.
Exp. Biol., 1997, vol. 200, no. 15, pp. 2107–2113.
Weindler, P., Böhme, F., Liepa, V., and Wiltschko, W., The
role of the daytime cues in the development of magnetic
orientation in a night-migrating bird, Behav. Ecol. Socio-
biol., 1998, vol. 42, no. 4, pp. 289–294.
Wiltschko, W., Über den Einfluß statischer Magnetfelder
auf die Zugorientierung der Rotkehlchen (Erithacus rubec-
ula), Zeitschrift für Tierpsychologie, 1968, vol. 25, no. 5,
pp. 537–558.
Wiltschko, W. and Balda, R.P., Sun compass orientation in
seed-caching scrub jays (Aphelocoma coerulescens), J. Comp.
Physiol. A, 1989, vol. 164, no. 6, pp. 717–721.
Wiltschko, W. and Wiltschko, R., Magnetic compass of
European robins, Science, 1972, vol. 176, no. 4030, pp. 62–
64.
Wiltschko, R. and Wiltschko, W., Avian navigation: a com-
bination of innate and learned mechanisms, Adv. Study
Behav., 2015, vol. 47, pp. 229–310.
Wiltschko, W., Wiltschko, R., and Keeton, W.T., Effects of
a “permanent” clock-shift on the orientation of young
homing pigeons, Behav. Ecol. Sociobiol, 1976, vol. 1, no. 3,
pp. 229–243.
Wiltschko, W., Wiltschko, R., and Keeton, W.T., The effect
of a “permanent” clockshift on the orientation of experi-
enced homing pigeons. I. Experiments in Ithaca, New York,
USA, Behav. Ecol. Sociobiol., 1984, vol. 15, no. 4, pp. 263–
272.
Wiltschko, W., Daum, P., Fergenbauer-Kimmel, A., and
Wiltschko, R., The development of the star compass in Gar-
den Warblers, Sylvia borin, Ethology, 1987, vol. 74, no. 4,
pp. 285–292.
Wiltschko, W., Balda, R.P., Jahnel, M., and Wiltschko, R.,
Sun compass orientation in seed-caching corvids: its role in
spatial memory, Animal Cognition, 1999, vol. 2, no. 4,
pp. 215–221.
Wiltschko, R., Munro, U., Ford, H., and Wiltschko, W.,
Contradictory results on the role of polarized light in com-
pass calibration in migratory songbirds, J. Ornithol., 2008,
vol. 149, no. 4, pp. 607–614.
Wiltschko, R., Munro, U., Ford, H., and Wiltschko, W.,
Response to the comments by R. Muheim, S. Åkesson, and
J.B. Phillips to our paper “Contradictory results on the role
of polarized light in compass calibration in migratory song-
birds,” J. Ornithol., 2008a, vol. 149, no. 4, pp. 663–664.
Yeagley, H.L. and Whitemore, F.C., A preliminary study of
a physical basis of bird navigation, J. Appl. Phys., 1947,
vol. 18, no. 12, pp. 1035–1063.
Zapka, M., Heyers, D., Hein, C.M., Engels, S., Schnei-
der, N.-L., et al., Visual, but not trigeminal, mediation of
magnetic compass information in a migratory bird, Nature,
2009, vol. 461, no. 7268, pp. 1274–1277.
... В области изучения ориентации и навигации птиц в последние 10-15 лет происходит существенный прогресс, который был отражён в недавних обзорах, к которым мы и отсылаем заинтересованного читателя (Chernetsov 2016;Mouritsen 2018). В конце 1960 -начале 1970-х гг. ...
Article
Full-text available
This review deals with the topics of bird migration research that have seen significant progress in the recent years. In these fields, earlier views have been significantly re-considered. They include the manifestation of migration, i.e. length of migratory flights, their speed, duration of stopovers, flight altitude, which taken together form the spatio-temporal realization of migration. New data on this array of research questions make it possible to ask new questions about physiological adaptations that make such flights possible, and factors that govern these processes. Significant progress is also evident in the study of long-distance orientation and navigation of migrating birds. However, it mainly concerns the mechanisms of magnetoreception and use of magnetic cues for orientation. Besides, students of bird migration could not ignore the ‘-omics’ revolution, which, like the studies of the manifestation of migration, was mainly caused by technological advances, mainly by the advent of next generation sequencing techniques. The new data obtained by these methods raised the hope to sort out the genetic basis of endogenous control of bird migration. However, the gap between the expectations and the actual results remains. To close this gap, new research into the molecular and cellular mechanisms that underlie migratory behavior is necessary. The results available now mainly refer to the processes that are relatively well understood, mainly the mechanisms of energy turnover and stress response.
... During such movements, they need to select and maintain a certain compass direction. This ability is termed orientation, and the set of behavioral adaptations that provide them with this ability is called the compass system [Chernetsov, 2016;Mouritsen, 2018]. Moreover, during local journeys within the familiar home range, if the animal is faced with the task of moving in a straight line for any prolonged period of time, it must have a compass system based on global cues, not just on local landmarks (Finkelstein et al., 2016). ...
Article
The existence of a magnetic compass system was fi rst demonstrated in birds. Since then, extensive of data have been accumulated on the operation of the magnetic compass in birds and its relationship with visual reception. The currently dominant concept is that the receptor supporting operation of the magnetic compass in birds is located in the retina. The most popular hypothesis for how magnetic fi eld receptors work is the radical pair hypothesis, in which cryptochrome, more specifi cally, the cryptochrome 4a isoform, is the candidate for the role of the primary magnetoreceptor molecule. Recent research has yielded data on the interaction of cryptochrome with various proteins involved in the phototransduction cascade, along with promising data from electrophysiological studies combining visual (light) and magnetic stimulation. In addition, a number of morphological studies of the avian retina are also helping to narrow the range of possible cells for the role of magnetoreceptor, with double cones being currently the most likely candidate. In this review, we discuss the latest research in this area.
... Chernetsov [54] argued convincingly of the need for data from Africa and Southern Europe to fully understand movements in Palearctic-African migrants. It has been suggested [54,55] that the speed of movement during spring migration in these species is higher, at least within Europe, than during the autumn return migration. After crossing the Mediterranean, migratory movements within Europe to breeding areas seem to take no more than 30 days [54]. ...
Article
Full-text available
Simple Summary We investigate the first arrival dates in spring of 31 migrant bird species for the period 1989–2022 in the Tatarstan Republic of Russia, towards the eastern limit of Europe. Most (28) of these species significantly advanced their migration during the study period. A number of species traits were investigated to see if these influenced the advances, but no clear relationships were identified, other than that the earliest-arriving species had advanced most. Higher spring temperatures in Tatarstan were associated with earlier arrival, but the possible influences on these birds of the temperatures at wintering grounds and on migration routes remain to be investigated. Abstract The first arrival dates of 31 species of migrant birds in the Tatarstan Republic of Russia were monitored for the 34-year period from 1989–2022. Trends in first arrival date were evaluated using regression against the year value. Patterns in arrival data with respect to species traits (habitat, migration distance, body weight, etc.) were evaluated using redundancy analysis. Relationships between first arrival dates and Tatarstan temperatures were also evaluated using regression methods of first-arrival date on monthly mean temperatures. Almost all (28 of 31) species revealed a significantly earlier migration arrival date; however, associations between arrival patterns and species traits were equivocal. Warmer temperatures were significantly associated with earlier arrival in 26 of the 31 species, but the relationship was insufficient to explain the average 11-day advance in species. For these species and in this location only the timing and location of arrival are well recorded; the exact wintering areas and migration routes, and the timing of these phases are less well understood. When these become better known, an investigation of the influence of environmental conditions (including temperature) on departure timing and passage timing and speed is recommended.
... Faktem jest, że bodźce świetlne przyciągają ptaki do źródła światła, a następnie powodują ich dezorientację na skutek zaburzenia mechanizmów orientacji w przestrzeni, a w konsekwencji trudność z uwolnieniem się ze świetlnej pułapki. Szczególnie silnie efekt ten oddziałuje na ptaki podczas jesiennej migracji (Zhao 2014, Chernetsov 2016. Przywabione sztucznym światłem i zdezorientowane ptaki często ulegają kolizjom z różnymi elementami infrastruktury (budynki, wieże telekomunikacyjne, przewody elektryczne itp.) lub padają z wyczerpania, krążąc wokół punktów świetlnych. ...
Book
Full-text available
Skala śmiertelności ptaków w wyniku kolizji ze szkłem jest trudna do oszacowania i nadal brakuje w Polsce metodycznie zaplanowanych i wykonanych badań, które te wartości mogłyby przybliżyć. Wynika to między innymi z ograniczonej wykrywalności ofiar (aktywność drapieżników i padlinożerców) i często skutkuje tym, że wartości są niedoszacowane. Pewne wyobrażenie dają wyniki badań opartych na nauce obywatelskiej i rejestrach prowadzonych przez organizacje pozarządowe zajmujące się ochroną ptaków w Polsce i na świecie. Na podstawie zbieranych w ten sposób danych możemy również dowiedzieć się, jakim obrażeniom ulegają ptaki w wyniku kolizji i jakie będą w związku z tym ich szanse na przeżycie. Ogromną wiedzą i doświadczeniem w tym zakresie dysponują ośrodki rehabilitacji dzikich zwierząt, do których najczęściej trafiają ofiary kolizji. Zanim jednak ptak po kolizji znajdzie się u odpowiednich specjalistów, często podejmowane są próby samodzielnego udzielenia mu pomocy bądź – gdy brakuje dostępu do wyspecjalizowanego ośrodka – pacjent przekazywany jest do najbliższego lekarza weterynarii. Metodyka postępowania z dzikim pacjentem, który dodatkowo uległ urazowi, różni się znacząco od postępowania ze zwierzętami udomowionymi, stąd kluczowe wydaje się, aby podstawowe wytyczne odnośnie pierwszej pomocy, diagnostyki i leczenia ptaków pokolizyjnych dostępne były dla wszystkich lekarzy weterynarii. Szeroko zakrojone działania wielu instytucji w zakresie edukacji ekologicznej przyczyniły się do znacznego wzrostu świadomości i wrażliwości przyrodniczej społeczeństwa. Nadal jednak nie wszyscy wiedzą, jak zachować się w sytuacji, gdy mamy do czynienia z dzikim zwierzęciem, które jest chore lub odniosło obrażenia. Najczęściej działania są intuicyjne i niestety nie mają wiele wspólnego z faktycznym udzieleniem pomocy, a nie zawsze jest możliwość dostarczenia pacjenta do wyspecjalizowanego ośrodka rehabilitacji czy doświadczonego lekarza weterynarii. Doświadczenie osób zaangażowanych w prace ośrodków rehabilitacji pokazuje, że ptaki, które ulegają kolizjom, często odnoszą poważne obrażenia niejednokrotnie doprowadzające do śmierci lub konieczności przeprowadzenia eutanazji. Na terenie województwa pomorskiego, ale i całego kraju, nadal niewiele jest funkcjonujących, wyspecjalizowanych ośrodków rehabilitacji. Istniejące, pomimo ogromnego zaangażowania, często borykają się z problemami finansowymi czy lokalowymi i szczególnie w okresie wzmożonej liczby przyjęć mogą mieć problem z udzieleniem pomocy wszystkim potrzebującym pacjentom.
... В ходе таких перемещений у них возникает необходимость выбирать и поддерживать определенное компасное направление относительно сторон света. Такую способность принято называть способностью к ориентации, а совокупность поведенческих адаптаций, обеспечивающих эту способность -компасной системой (Chernetsov, 2016;Mouritsen, 2018). Кроме того, и при локальных перемещениях в пределах знакомого животному по предыдущему опыту индивидуального участка, если перед животным стоит задача сколько-нибудь продолжительное время двигаться по прямой, оно должно обладать компасной системой, основанной на глобальных источниках информации, а не только на локальных ориентирах (Finkelstein et al., 2016). ...
Article
The existence of the magnetic compass system was first shown in birds. Since then, a large amount of data has been accumulated on the performance of the avian magnetic compass and its relationship with visual reception. The current dominant concept is that the receptor for the magnetic compass in birds is located in the retina. The most popular hypothesis for the mechanism of operation of magnetic field receptors is the radical pair model, and a candidate for the role of the primary magnetoreceptor molecule is cryptochrome, and more specifically, its isoform, cryptochrome 4a. In recent years, data have been published on the interaction of cryptochrome with some proteins involved in the phototransduction cascade, as well as promising data from electrophysiological studies combining light and magnetic stimuli. In addition, a number of morphological studies of the avian retina also allow us to narrow down the range of promising cells for the role of a magnetoreceptor, and the double cone is currently the most likely candidate. In this review, we discuss the latest research data in this area.
... Data on animal migrations are not just a description of the pathways of birds and mammals, but unique information about their interaction with the [26,27]. Thus, the hardware provided by the URAGAN experiment makes it possible not only to monitor the moving animals of interest, but also to try to find out the reasons for changing the routes of their migrations and wintering sites. ...
... The young of many bird species use vector navigation/clock-andcompass strategy, i.e. an innate compass heading and information on the distance towards the population-specific winter range (Able, 2001;Berthold, 1991;Mouritsen and Larsen, 1998), or even assume some inherited mechanisms of controlling their position on the migration route (Chernetsov, 2016;Thorup et al., 2010). However, it was demonstrated that young birds, unlike adult ones, are not always capable of assessing the displacement from the migration route and compensating for it, and continue to fly in the initial direction of migration (Perdeck, 1958;Thorup et al., 2007). ...
Article
Full-text available
The orientation of naive animals during their first migration is extensively studied in birds and sea turtles, whereas the data for other groups such as amphibians are still scarce. To date, it is unknown whether young-of-the-year anurans perform a random or directional search for the hibernation sites, and what cues (global or local) do they use. We conducted a series of field experiments to study the orientation behavior of juvenile common frogs during their first wintering migration. We captured 1614 froglets from two subpopulations with different directions of migration and assessed their orientation in large circular outdoor arenas (20 m in diameter) on the opposite sides of the river. Before the migration, froglets used local cues and moved back towards the forest (summer habitat). At the start of migration, the froglets do not move randomly: they navigate towards the river using local cues; later, however, before approaching the hibernation site, they memorize the compass direction of migration and follow it using global cues. Orientation along a memorized compass heading begins to dominate in the hierarchy of orientation mechanisms, and this predominance is maintained even after reaching the hibernation site. Unlike in birds, no innate direction of migration was found.
... There is also evidence that blue light is needed, while monochromatic red light can result in disorientation. However, independent from light colour, strong light can cause birds to orient into a wrong direction, especially in dark nights (Ballasus and Hill 2009;Chernetsov 2016). ...
Technical Report
Full-text available
Review on the current knowledge of the impact of artificial light on marine and coastal habitats with special focus on the Waddensea and intertidal areas.
Chapter
Full-text available
Negatywne oddziaływanie sztucznego oświetlenia (Artificial Light at Night – ALAN) jest jedną z głównych, antropogenicznych przyczyn odpowie�dzialnych za bezpośrednią śmiertelność ptaków migrujących nocą. Do udokumentowanego, negatywnego wpływu ALAN na ptaki zaliczane są: efekt przyciągania i efekt dezorientacji. Efekt bariery z kolei jest przedmiotem badań, jednak brakuje jak dotąd jednoznacznego stwierdzenia na temat charakteru tego zjawiska. Większość badań wskazuje, że za efekt przyciągania, przyczyniający się do kolizji z różnymi konstrukcjami na lądzie i na morzu, odpowiedzialne jest oświetlenie o barwie niebieskiej i zielonej. Jednak czerwone i białe światło również wskazywane było w badaniach jako wywierające wpływ na przyciąganie ptaków. Z uwagi na te rozbieżności, konieczne są dalsze badania, w efekcie których będzie możliwe określenie, jaka barwa światła będzie najbezpieczniejsza dla ptaków. W chwili obecnej zasadne wydaje się rekomendowanie wykorzystania źródeł światła o barwach ciepłych (żółtej, bursztynowej) o temperaturze barwowej nieprzekraczającej 3000 K.
Article
Full-text available
A model of avian goal-oriented navigation is described that is based on two empirical findings building a bridge from ornithology to atmospheric chemistry. (1) To orient their courses homeward from distant unfamiliar areas, homing pigeons require long-term exposure to undisturbed winds at the home site and olfactory access to the environmental air at home and abroad. (2) Above Germany, ratios among some atmospheric trace gases vary along differently oriented spatial gradients as well as depending on wind direction. The model emulates finding (1) by utilising the analysed air samples on which finding (2) is based. Starting with an available set of 46 omnipresent compounds, virtual pigeons determine the profile of relative weights among them at each of 96 sites regularly distributed around a central home site within a radius of 200 km and compare this profile with corresponding profiles determined at home under varying wind conditions. Referring to particular similarities and dissimilarities depending on home-wind direction, they try to estimate, at each site, the compass direction they should fly in order to approach home. To make the model work, an iterative algorithm imitates evolution by modifying sensitivity to the individual compounds stepwise at random. In the course of thousands of trial-and-error steps it gradually improves homeward orientation by selecting smaller sets of most useful and optimally weighted substances from whose proportional configurations at home and abroad it finally derives navigational performances similar to those accomplished by real pigeons. It is concluded that the dynamic chemical atmosphere most likely contains sufficient spatial information for home-finding over hundreds of kilometres of unfamiliar terrain. The underlying chemo-atmospheric processes remain to be clarified.
Article
Full-text available
Contemporary data on the orientation and navigation of animals from different taxa with the geomagnetic field are considered in the review. The mechanisms of magnetoreception in animals are described. In this regard, several taxa of fishes are able to perceive the magnetic field via electroreceptors. Some animals can sense changes in the magnetic field polarity via sensory cells containing iron compounds. In addition, animals from different taxa are able to perceive the inclination of the magnetic field via a change in singlet and triplet yields of radical-pair reactions under the influence of the magnetic field. In the latter case, the molecules of cryptochrome can play a crucial role. It is considered that the cryptochrome excited by shortwave light can produce long-living radical pairs affected by the geomagnetic field.
Article
The sun compass was discovered by Kramer (1950) in caged birds showing migratory restlessness. Subsequent experiments with caged birds employing directional training and clock shifts showed that the sun azimuth is used, and the sun altitude ignored. In the laboratory, McDonald found the accuracy to be ±3° to ±5°. Caged birds trained at medium northern latitudes were able to allow for the sun's apparent movement north of the arctic circle, but not in equatorial and trans-equatorial latitudes. In homing experiments, and employing clock shifts, Schmidt-Koenig demonstrated that the sun compass is used by homing pigeon during initial orientation. This is the principal evidence for the existence of a map-and-compass navigational system. Pigeons living in equatorial latitudes utilize the sun compass even under the extreme solar conditions of equinox, achieving angular resolution of c3° in homing experiments. -from Authors
Article
New experiments have been made on homing pigeons deprived of olfactory perception by occlusion of both nostrils with cotton tampons, or by resection of the olfactory nerve on one side and occlusion of the nostril on the other. They confirm the strong disturbance of orientation of the birds at the moment of release, which had already been observed in subjects with both olfactory nerves resected. Observations conducted on subjects with resected nerves have not shown other disturbances in general behaviour. The possibility is considered that, at the home site, pigeons associate odorous stimuli carried by winds from the surroundings with the direction from which they come. Recognizing a known odour at the release point, they should thereby be able to orient themselves toward home by flying in a direction opposite to the one with which the odour had been associated.
Chapter
In view of the suggestion from earlier experiments that detection of magnetic and gravitational stimuli may be linked, we investigated the possibility that gravitational changes across the synodic lunar month might influence the initial bearings of homing pigeons under conditions when naturally occurring fluctuations in the geomagnetic field have been shown to influence the birds’ orientation behavior. Six separate series of releases were conducted during four different years at three different sites. In all six, a nearly linear correlation between the pigeons’ mean vanishing bearings (MVB) and the day of lunar month (DLM) was found. The monthly oscillation of MVB could apparently exist stably in either of two modes, each 180° out of phase with the other; in some series the oscillation was from new moon to new moon and in others from full moon to full moon. Statistical analyses of several types support the notions that the best frequency of the oscillations of MVB is indeed almost exactly that of the synodic lunar month and that the turning points of the oscillations are the syzygies (new moon and full moon). However, attempts to tie the variations in MVB directly to the gravitational changes occurring during the lunar month have so far been unsuccessful; hence it is not yet known what environmental stimuli are acting as the immediate causes of the birds' changing orientation behavior.