Download as pdf or txt
Download as pdf or txt
You are on page 1of 58

CHAPTER 2

THE EVOLUTION OF
CONSPICUOUS AND DISTINCTIVE
COLORATION FOR
COMMUNICATION IN BIRDS
GREGORY S. BUTCHER and SIEVERT ROHWER

1. INTRODUCTION

Our goal is to elucidate the selective pressures that account for the
origin and maintenance of conspicuous and distinctive coloration
(hereafter referred to collectively as colorfulness) in birds. Our ideas
are relevant to the study of other animals as well. We propose an "ad-
aptationist program" (Gould and Lewontin, 1979) that features the use
of the comparative method (Hailman, 1976; Ridley, 1983; Clutton-Brock
and Harvey, 1984), color manipulations, and tests of a priori predic-
tions. Our major effort is to compare and contrast the major hypotheses
that have been proposed to account for colorfulness in birds. Progress
in understanding bird coloration will come more quickly when re-
searchers evaluate more than one hypothesis at a time. The availability
of multiple hypotheses serves two major purposes. First, it frees the
researcher from psychological dependence on the hypotheses being
tested (Loehle, 1987). Second, with careful planning, it permits strong
inference testing (Platt, 1964) among the most relevant hypotheses.

GREGORY S. BUTCHER and SIEVERT ROHWER • Department of Zoology and Burke


Museum, University of Washington, Seattle, Washington 98195. Present address for
G. S. B.: Cornell Laboratory of Ornithology, Ithaca, New York 14850.

51
D. M. Power (ed.), Current Ornithology
© Springer Science+Business Media New York 1989
52 GREGORY S. BUTCHER and SIEVERT ROHWER

First, we discuss the goals of the review and introduce the com-
parative method and color manipulations as potentially powerful tools
for determining the benefits of conspicuous coloration. Next we define
our major color terms (conspicuous, distinctive, and cryptic) and pro-
vide a short historical overview of coloration theories. In the next three
sections, we discuss five important intraspecific patterns of bird color-
ation that have been well documented and that are exactly the kinds
of patterns for which we hope to provide functional explanations. Four
of the patterns suggest general functional explanations; the fifth pattern
(color dimorphism/polymorphism unrelated to sex or age) seemingly
cannot be explained by a single functional hypothesis. The second half
of this review (1) presents updated versions of a number of color hy-
potheses; (2) discusses the type of evidence required to (a) distinguish
similar hypotheses, (b) provide strong support for a given hypothesis,
or (c) reject a given hypothesis; and (3) presents the evidence currently
available concerning the hypothesis. Because this area of study is rel-
atively undeveloped, there are few adequate tests of hypotheses; there-
fore, we do not dwell on the weaknesses of those that are available.
Although biologists have speculated about the causes of colorful-
ness for many years (see our historical overview below), this area of
study is relatively undeveloped. For instance, hypothesis testing has
been rare and lags dramatically behind hypothesis generation. Most
color hypotheses, as they have been presented to date, have been in-
troduced as plausible explanations, supported by only a few well-cho-
sen examples, not systematic tests of predictions. Rarely has more than
one hypothesis been considered at a time, and rarely have broad in-
terspecific tests been undertaken. We hope that by summarizing the
numerous hypotheses available for testing we can help to move the
field of animal coloration from the era of hypothesis generation to that
of hypothesis testing.
The major tool for ferreting out general causes of the origin of
colorfulness should be the comparative method (elutton-Brock and
Harvey, 1984, and references cited therein). From each hypothesis con-
cerning colorfulness, we can predict a correlation between colorfulness
and one or more ecological or behavioral characteristics of populations.
If the predictions are confirmed in a wide range of taxa, that charac-
teristic may be affirmed as an important cause of colorfulness. The
comparative method provides a useful check on adaptationist zeal by
revealing phylogenetic correlations of traits that indicate that a trait
may be nonadaptive. Unfortunately, few studies of bird coloration have
used the comparative method. One of the major reasons for the lack of
use of the comparative method is that avian phylogenies are currently
CONSPICUOUS COLORATION 53

in much dispute (Cracraft, 1981; Olson, 1982; Sibley and Ahlquist,


1983). Studies of individual species are useful for determining what
the current functions of color might be and for predicting cross-tax-
onomic patterns, but these studies alone cannot help us separate orig-
inal functions from secondary uses of coloration.
A major limitation of the comparative method is that it cannot
prove a functional relationship. Also, because it requires data from a
wide variety of species, the quality of data is not always as high as
might be desired. Thus, broad comparative studies must be supple-
mented by strong intraspecific tests of function. Color manipulations
are an important tool for these intraspecific tests. Each of the color
hypotheses provides a functional explanation for colorfulness from which
the effects of color manipulations can be predicted; if the manipulation
produces the predicted change, it provides strong evidence that the
coloration has the proposed function in that species. Color manipula-
tions might include making an individual look more similar to a syn-
topic congener, making an individual of one age or sex class look as
though it were a member of another class, increasing or decreasing the
conspicuousness of an individual, or obscuring a particular patch of
color.
One limitation of the approach we have taken is that we consider
each hypothesis separately or in pairs in the sections that follow and
make predictions assuming that each hypothesized benefit alone ac-
counted for conspicuousness. It is far more likely that more than one
of the proposed benefits combine to promote the conspicuousness of
animals, and it may be that consideration of each benefit separately
causes us to overlook patterns that reveal important combinations of
causes (Hilborn and Stearns, 1982). On the other hand, there are far
too many hypotheses to consider all possible interactions. Thus, the
first step is to delineate predictions clearly for single hypotheses. If we
are lucky, important causes will create patterns that reveal themselves
despite the interactions of other causes. When an investigator is study-
ing a particular system, many of the hypotheses we consider can be
rejected on a priori grounds, reducing the number of possible benefits
and making the consideration of interactions among possible benefits
more feasible.

1.1. Classifying Coloration: Conspicuous, Distinctive, or


Cryptic
We define signals as sensory cues (e.g., color, vocal, pheromonal)
that have been modified by natural selection to increase the effective-
54 GREGORY S. BUTCHER and SIEVERT ROHWER

ness with which they transmit information. This information must mod-
ify the behavior of other organisms such that the signaler benefits, on
average, from the signal. In addition, the receiver must, on average,
benefit from responding to cues that resemble the signal. This definition
has two major implications. First, signals need not be mutually bene-
ficial to senders and receivers. Deceptive signals that induce behavior
that is harmful to the responder are possible and are included in the
definition of a signal. Signals can have high costs to senders as well,
provided they have compensating benefits. Second, signals are different
from signs, which also transmit information to other organisms. By
definition signs have not been modified by natural selection for infor-
mation transfer, and the information they convey may be detrimental
to the individual that generates them. For example, the noise made by
a mouse rustling in the leaves may attract a predator, but mice have
certainly not been designed by natural selection to make such noise.
If an animal's coloration is to function as a signal, it must be either
conspicuous or distinctive (Burtt, 1986). We define distinctive as "un-
usual, easy to distinguish, easy to remember." Any spectrally bright
color or any bold pattern that is rare in the environment should be
distinctive. We define conspicuous as "easy to see at a distance." Large
patches of bright spectral colors, white, or black should all be con-
spicuous, if they contrast with the background against which the co-
loration is viewed. Because conspicuous coloration involves long-dis-
tance communication, background contrast and large patch size are
more important for conspicuousness than for distinctiveness. Complex
coloration (the juxtaposition of small, contrasting patches of color) is
an important component of distinctiveness, but not of conspicuousness.
Distinctive coloration may serve as a sign stimulus or as a social releaser
(Dilger, 1956; Colgan, 1983). When it is unclear whether conspicuous-
ness or distinctiveness is favored, we use the tern colorful to refer to
the two simultaneously.
Some color patches that appear at close range to be quite conspic-
uous are actually rather cryptic at a moderate distance (Hailman, 1977;
Endler, 1978). Thus, careful observation in the natural habitat is re-
quired before conspicuousness can be evaluated (Lythgoe, 1979; Burtt,
1986). Because cryptic coloration is the first line of defense against
predation for most animals (Endler, 1978; but see Hailman, 1977; Baker
and Parker, 1979), individuals may be selected for short-range distinc-
tiveness for communication with conspecifics and for long-range cryp-
sis for protection against predators. Certain patterns, such as the stripes
of zebras, achieve an excellent balance between distinctiveness and
crypsis (Endler, 1978).
CONSPICUOUS COLORATION 55

When conspicuousness of plumage is favored for intraspecific com-


munication, a number of compromises may be favored because of the
color conflict (Hingston, 1933; Hamilton and Peterman, 1971; eott, 1985)
between communication and crypsis. Most authorities argue that the
benefit of crypsis is lessened visibility to predators (Endler, 1978, and
references cited therein), but a major benefit of crypsis might be less-
ened visibility or lessened threat to competitors. Morton (1976) attri-
butes the crypsis of young Yellow Warblers (Dendroica petechia) to
the fact that they must often intrude on areas defended by adults to
obtain food. Similarly, young hummingbirds lack the iridescent gorget
of adults perhaps because they must intrude on territories defended by
adults to feed; in support of this idea, young hummingbirds were shown
to be more successful at intruding than were adults (Ewald and Rohwer,
1980). In Harris' Sparrow (Zonotrichia querula), reduced threat rather
than reduced visibility is suggested as the function of the more cryptic
coloration of young birds because they were attacked less frequently
than adults in open areas on a per-encounter basis (Rohwer and Ewald,
1981).
One of the ways available to birds to compromise between con-
spicuousness and crypsis is dynamic coloration: iridescence, erectile
crests, air sacs, and coverable patches of color, such as the epaulets of
the Red-winged Blackbird (Agelaius phoeniceus). Dynamic coloration
may also be favored to increase flexibility in signaling; for instance, a
male Red-winged Blackbird may signal that he is an owner by exposing
his epaulets and that he is an intruder by covering his epaulets (Hansen
and Rohwer, 1986). Other compromises between crypsis and conspic-
uousness include the restriction of conspicuousness to particular parts
of the body (often the face, chest, wings, or tail, but rarely the back),
seasonal color dimorphism, and age- and sex-limited expression of
conspicuousness. In many species in which the cost of conspicuousness
is particularly great, animals may use other modes of communication
(especially vocalizations for birds) for the same purposes for which
other species use color. It is probably because of the color conflict that
it is often difficult to characterize species on the cryptic-conspicuous
axis.
Another reason for the difficulty in characterizing species on the
cryptic-conspicuous axis is the extreme difficulty in quantifying con-
spicuousness in natural settings (Hailman, 1977, 1979; Lythgoe, 1979;
Burtt, 1986). Quantification of conspicuousness requires determining
(1) the discriminatory abilities of the various observers of an animal,
(2) biologically relevant distances between the observers and the an-
imla, (3) the color properties of the environments in which the animal
56 GREGORY S. BUTCHER and SIEVERT ROHWER

is seen, (4) the color properties of the backgrounds against which the
animal is seen, and (5) the color properties of the animal itself. The
complexity is overwhelming. Endler (1978) suggests that birds have
visual acuity and color vision similar to that of humans (but see Chen
and Goldsmith, 1986, and references cited therein, most published after
Endler's paper). Burtt (1986) provides a rationale for starting with the
assumption that all diurnal terrestrial vertebrates have similar abilities
in visual discrimination; however, he points out important violations
of this assumption. For example, birds are sensitive to ultraviolet light
(Parrish et al., 1984; Chen and Goldsmith, 1986), but humans are not.
Faced with this complexity, biologists will differ in their approach.
Some will study the visual abilities of birds in the laboratory (Hailman,
1967; Parrish et al., 1984; Chen and Goldsmith, 1986). Some will at-
tempt quantitative studies of conspicuousness in nature (Hailman, 1979;
Burtt, 1986; Endler, 1987). Some will be satisfied for now with quali-
tative and general discussions of relative conspicuousness (this chapter,
and most of the papers cited herein). Many will decide that conspic-
uousness is too fuzzy a concept for them and will stop reading right
here. Despite the inherent difficulties in quantifying conspicuousness,
we believe that conspicuousness is the major factor that is selected for
or against in animal communication. We believe that progress in sen-
sory physiology concerning the visual discriminatory abilities of a va-
riety of animals, added to careful field observations of the visual context
within which animals are seen, will allow us to avoid or correct mis-
takes in our qualitative assessments of relative conspicuousness.

1.2. A Brief History of Color Hypotheses


New hypotheses concerning the adaptive significance of the col-
orfulness have appeared ever since Darwin's publication of The Descent
of Man, and Selection in Relation to Sex in 1871. Darwin wrote that
colorfulness and "ornaments of all kinds ... apparently serve to excite,
attract, or fascinate the females" (p. 394 of 2nd edition, 1890). Although
Darwin recognized that colorful males are subject to intense intrasexual
selection, he nonetheless maintained that the coloration itself serves
primarily to attract females. Wallace (1889) rejected Darwin's conten-
tion that colorfulness was due to female choice, believing, as did many
of Darwin's critics, that female choice of male colors implied an un-
acceptable "aesthetic sense" in animals. Instead, Wallace proposed that
colorfulness evolved to increase the distinctiveness of species-specific
coloration in order to prevent hybridization and to promote the dis-
CONSPICUOUS COLORA nON 57

covery of flocks by lost conspecifics. Moynihan (1960) expanded Wal-


lace's ideas on the uses of colorfulness by flocking birds.
Hingston (1933) rejected both sexual selection and species recog-
nition arguments and proposed that colorfulness was used in all cases
for warning (to potential predators) or threat (to potential rivals), often
failing to distinguish the two. Hingston also developed the idea of the
color conflict between crypsis and conspicuousness (see Section 1.1).
Huxley (1938) presented the first balanced review of the theories of
bird coloration, including a classification of the possible functions of
coloration that remains essentially complete today. Cott (1946; see also
Cott and Benson, 1970) extended the category of warning coloration
for birds by presenting evidence that many birds are unpalatable and
that colorfulness may be positively correlated with unpalatability. Baker
and Parker (1979) further extended the theory of warning coloration in
birds by suggesting that some birds are so vigilant and so agile that
they are unprofitable prey for most predators at most times and that
individuals that are relatively "unprofitable prey" signal this fact through
colorfulness.

2. THE THREE RULES OF AVIAN COLOR DIMORPHISM

Many intraspecific patterns of bird coloration have been described;


a number of these patterns are referred to in this chapter. Three of these
patterns recur in our work over and over again, and it is primarily these
patterns that we wish to be able to explain. We refer to them as the
three rules of avian color dimorphism. They are as follows:
1. When birds are sexually color dimorphic, males are more col-
orful than females.
2. When the age classes of birds feature different colors, adults are
more colorful than young birds.
3. When birds are seasonally color dimorphic, breeding season
birds are more colorful than nonbreeding season birds.
All three patterns of dimorphism have been documented for a va-
riety of species (Darwin, 1871), for waterfowl (Kear, 1970), for the Eu-
ropean avifauna (Baker and Parker, 1979), and for selected groups of
North and Central American passerines (Skutch, 1940). Sexual and
seasonal color dimorphisms have been documented for African plo-
ceids (Moreau, 1960; Crook, 1964) and for North and Central American
passerines (Skutch, 1957; Hamilton, 1961; Hamilton and Barth, 1962;
Bailey, 1978). In addition, Burley (1981) documented patterns of sexual
58 GREGORY S. BUTCHER and SIEVERT ROHWER

dimorphism in a number of flocking species, and Goodwin (1960) doc-


umented patterns of sexual color dimorphism in pigeons. Andersson
(1983, and references cited therein) documented patterns of seasonal
color dimorphism in ducks. Rohwer and Butcher (1988) documented
age-related color dimorphisms in North American passerines during
the nonbreeding and breeding seasons, including information on sea-
sonal changes. The latter added to the work of Rohwer et a1. (1980),
which concentrated on the breeding season.
There are exceptions to each of these rules. For example, in the
three species of phalaropes (genus Pha1aropus) and in the Painted Snipe
(Rostratu1a bengha1ensis). females are more colorful than males (Dar-
win, 1871). Associated with this reversal of colorfulness is a reversal
in the role of the sexes in courtship and parental care. In some species
of woodpeckers, young birds of both sexes have more colorful head
patches than do adults (Darwin, 1871). It is probably relevant that wood-
peckers nest in holes and therefore are relatively immune from pre-
dation while in the nest. Finally, in Brachyramphus murrelets (Marbled
and Kittlitz's), winter adults are a high-contrast black and white, whereas
summer adults are a cryptic mottled brown. In these species, unlike
most alcids, pairs form at sea in winter and nest as isolated pairs on
the ground or in trees in the summer.
When these three rules are considered in combination, they indi-
cate that some form of sexual selection must be one of the important
selection pressures favoring colorful birds (Darwin, 1871; contra Baker
and Parker, 1979). Three forms of sexual selection are commonly in-
voked to explain colorful birds: epigamic species recognition (finding
a mate of the proper species), agonistic sexual selection (intrasexual
competition over mates), and epigamic sexual selection (mate choice).
Given that it is during the breeding season that colorful birds are
likely to expose their nest and sedentary offspring to visually hunting
predators, it is particularly striking that breeding-season birds are more
likely to be conspicuous than nonbreeding-season birds (Andersson,
1983). Thus, it is during the season in which conspicuous coloration
should impose the highest cost that birds are most likely to be colorful.
In most species, males are under stronger sexual selection than are
females (Darwin, 1871; Trivers, 1972). consistent with the finding that
males are more likely to be conspicuous than females. Finally, young
birds are commonly assumed to be handicapped in breeding relative
to adults (Lack, 1954, 1966, 1968; Selander, 1965, 1972); the increased
likelihood that they will be cryptic may therefore be a method of either
compensating for these handicaps (Rohwer et a1., 1980; Lyon and Mont-
gomerie, 1986) or of increasing survival at the expense of reproduction
CONSPICUOUS COLORATION 59

while young (Lack, 1954, 1966, 1968; Selander, 1965, 1972; Studd and
Robertson, 1985; Rohwer and Butcher, 1988), or both.
Baker and Parker (1979) document the validity of these three rules
for European birds, but they argue that the rules are consistent
with the idea that colorfulness evolved as a signal to predators. They
argue that males and adults are more likely to be vigilant against pred-
ators and to be able to escape attack than are females and young, hence
that males and adults should signal to predators (using conspicuous
coloration) that they represent unprofitable prey. Moreover, they argue
that all adult birds in temperate regions are more profitable in winter
than in summer, because (1) birds must spend more time looking for
food in winter than in summer, (2) there is less cover in winter than
in summer, (3) the predator has fewer alternatives in winter than in
summer, and (4) young birds become relatively more experienced in
late winter than they were in summer. Andersson (1983) counters that
the patterns of seasonal color dimorphism, in particular, do not favor
Baker and Parker's (1979) interpretation because seasonal color changes
are not timed to coincide with the presence or absence of inexperienced
young, but rather are timed to coincide with the presence or absence
of courtship. The best example is the eclipse plumage of male ducks:
in species with an eclipse plumage, the males become cryptic at exactly
the time when inexperienced offspring first appear. Males of species
with extensive early-winter courtship molt out of eclipse plumage ear-
lier than males in species with only spring courtship (Andersson, 1983).

3. THE THREE RULES AND THE THREAT HYPOTHESIS

The threat hypothesis argues that conspicuously colored individ-


uals have evolved because they are better at defending resources. How
colorfulness might work as a threat within a monomorphic class of
animals is discussed in Section 6.7. The status-signaling hypothesis is
a subset of the threat hypothesis that argues that differences in con-
spicuousness have evolved to signal differences in the ability of indi-
viduals to win agonistic contests (Rohwer, 1975, 1982). The status-
signaling hypothesis might explain the three rules of color dimorphism
as follows: (1) males are more colorful than females in the nonbreeding
season when they are larger than females and thus better fighters, or
when defense of resources is more important for males than for females;
(2) older birds are more colorful than younger birds because their greater
experience gives them greater fighting ability; and (3) breeding birds
60 GREGORY S. BUTCHER and SIEVERT ROHWER

are more colorful than nonbreeding birds because competition for mates
and breeding space is common during the breeding season and because
there is rarely any way to avoid competition and still achieve high
mating success. Balph et a1. (1979) and Whitfield (1987) argue that the
status-signaling hypothesis should be restricted to cases in which dif-
ferences in coloration signal differences in fighting ability within age
or sex classes; however, we argue here that the greatest power and value
of the status-signaling hypothesis may be its ability to explain why
some species advertise information about age or sex while other species
hide it.
Males dominate females during the nonbreeding season in 14 of
the 18 sexually color-dimorphic species for which we obtained infor-
mation; the data are equivocal in two species (Table I). In captive flocks
of Chaffinches (Fringilla coelebs), females dyed to resemble males im-
proved their status against both males and females, even though most
males remained dominant to the dyed females (Marler, 1955). In two
species, females dominated males, even though they are more cryptic
than males. In eight species, the sexes are the same color. Only 3 of
the 26 species have a prebreeding molt; thus, it is uncertain for most
of these species that their plumage is an adaptation to the nonbreeding
season. The association of dominance and conspicuous coloration in
this sample may merely reflect the frequency of male dominance and
of male conspicuousness in the sample.
We should note here the interesting article by S. M. Smith (1980)
that suggests that female dominance may be the rule within breeding
pairs. This suggestion, even if proved true, is not particularly relevant
to the status-signaling hypothesis. The threat hypothesis suggests that
coloration should be useful in deterring fights among relative strangers,
not among individuals with a long-term close relationship (e.g., mates
or neighbors).
Experiments on two bird species in which adult males and females
dominate first-year males and females during the nonbreeding season
are consistent with the status-signaling hypothesis. In both species,
adults are more colorful than first-year birds. Experimental manipu-
lations in White-crowned Sparrows (Zonotrichia leucophrys) showed
that first-year birds dyed to mimic adults rise in status to equal that of
adults; thus, the signal normally correlated with age (and thus status)
is believed independently of other cues of age or fighting ability (Par-
sons and Baptista, 1980; Fugle et al., 1984). Similarly, first-year Harris'
Sparrows dyed to mimic adult males were able to dominate all other
first-year birds in a flock of all first-year birds (Rohwer, 1985). Similar
experimental results on a variety of species or the demonstration that
TABLE I
Male versus Female Dominance during the Nonbreeding Season o . b

Dominant Color
Species' sex? dimorphic? Prebreeding molt? Field/aviary'? Reference
Dendrocopos (Picoides) pubescens Male Yes No Field Kilham (1974)
Dendrocinc1a fuliginosa Female No Field Willis (1972)
Zootheria (Ixoreus) naevia( us) Male Yes No ? Martin (1970)
Parus atricapillus Male No No Field Odum (1942); Glase (1973)
Parus carolinensis Male No No Field Dixon (1963)
Parus major Male Variable No Field Brian (1949)
Zosterops lateralis Male Yes ? Both Kikkawa (1961)
Zonotrichia (MelospizaJ melodia Male No No Field Knapton and Krebs (1976)
Zonotrichia querula Male Variable Yes Field Rohwer et a1. (1981)
Junco hyemalis Male Variable No Field Ketterson (1979)
Icterus spurius Male Yes No Field Butcher and Wimberger
(unpublished manuscript)
Fringilla coelebs Male Yes No Both Hinde (1956)
Fringilla montifringilla Male Yes No Both Hinde (1956)
Carduelis chloris Male Yes No Aviary Hinde (1956)
Carduelis tristis Male Yes Yes Both Coutlee (1967)
Curduelis carduelis Male No No Both Hinde (1956)
Acanthis (Carduelis) flammea Male Yes No Aviary Dilger (1960)
Carpodacus purpureus Female Yes No Both Samson (1977)
Carpodacus cassinii Female Yes No Both Samson (1977)
Carpodacus mexicanus For = Yes No Av/field Thompson (1960)
Loxia curvirostra Male Yes No Both Tordoff (1954)
Pyrrhula pyrrhula Female? Yes No Both Hinde (1956)
Coccothraustes coccothraustes Male Yes No Both Hinde (1956)
Coccothraustes vespertinus Male Yes No(sl?) Field Balph et al. (1979)
Cyanocitta stelleri Male No No Field J. L. Brown (1963)
Garrulus glandarius Male No Yes Aviary Goodwin (1951)

"Spring molt: Dwight (1900) and Forbush and May (1929) for eastern North America; Oberholser (1974) for western North America; Witherby et 01, (1943)
for England.
bPrebreeding molt: sl? refers to the possibility of a very restricted prebreeding molt. All prebreeding molts in this group are considered partial molts.
'Scientific names and sequence from Clements (1981), based on Morony et oJ. (1975); in parentheses, from A.O.V. (1983).
62 GREGORY S. BUTCHER and SIEVERT ROHWER

a large number of unusual color patterns are correlated with unusual


dominance patterns will be required to establish the generality of the
status-signaling hypothesis.

4. VARIABILITY, MONOMORPHISM, AND THE THREAT


HYPOTHESIS

Consistent with the three rules of avian color dimorphism is the


fact that, in many species, males and females are equally colorful, young
and adult birds are equally colorful, and/or breeding-season and non-
breeding-season adults are equally colorful. In these species, either
there must be a powerful constraint against color dimorphism, or there
must be selective pressures other than sexual selection that favor col-
orful birds. In addition, many species show within-class color varia-
bility. We argue that the importance of coloration for threat is most
consistent with the known patterns of the presence and absence of the
three color dimorphisms.
There are four major kinds of plumage variability within a bird
species: (1) no variability (all individuals either cryptic or conspicuous);
(2) between-class variability, with two to four more or less distinct color
classes; (3) within-class variability within one, some, or all classes; and
(4) variability unrelated to sex or age. For most of the species discussed
in this section, the plumage variability is correlated (at least to some
extent) with age and sex; for color polymorphisms that are independent
of age and sex, see Section 5. Plumage variability in birds is poorly
documented, with essentially no established methodology for quanti-
fication; thus, few species can be confidently assigned to the above four
categories.
A previous discussion (Rohwer, 1982) suggested that variable plu-
mages might be more characteristic of the nonbreeding season and that
monomorphic plumages might be more characteristic of the breeding
season. Nonetheless, examples of all four kinds of variability can be
found in both seasons; thus, the discussion of factors influencing which
pattern will characterize a species should address both seasons.
A major test of the status-signaling hypothesis is its ability to ac-
count for the four categories of plumage variability listed above.

4.1. Monomorphism
If the threat hypothesis is relevant, monomorphic cryptic species
should fall into one of four classes: (1) species for which aggression is
CONSPICUOUS COLORATION

unimportant; (2) species in which crypsis for avoiding predation is


more important than colorfulness for success at agonistic encounters;
(3) species that must sneak resources that are defended by more dom-
inant species; or (4) species for which coloration cannot facilitate win-
ning or reduce the number of aggressive encounters (see Section 6.7).
Monomorphic colorful species should be those in which all indi-
viduals must win agonistic contests to survive or reproduce. If a pop-
ulation can be divided into relatively good fighters and relatively poor
fighters, a major question might be: In what situations should (or can)
poor fighters disguise their lack of fighting ability? Poor fighters should
try to disguise their poor fighting ability when there are few avenues
to survival or reproductive success apart from winning fights. Poor
fighters may be able to disguise their poor fighting ability when contests
are often decided merely on the basis of priority (see Section 6.7) or
in other circumstances in which contestants are reluctant to fight, per-
haps because fighting is risky (Rohwer, 1982) (see Section 6.7). Because
both plumage and fighting ability are often correlated with age and sex,
plumage monomorphism may result when poor fighters (young birds,
birds of the subordinate sex, or other poor fighters) mimic the plumage
of good fighters in order to hide information about relative fighting
ability.
Poor fighters may not mimic good fighters because either (1) they
suffer less aggression by being cryptic or by identifying themselves as
subordinates, or (2) they can successfully obtain resources without fighting
dominants. Crypsis is a benefit to a poor fighter when he can sneak
resources defended by dominants (Ewald and Rohwer, 1980); subor-
dinates might suffer less aggression from a dominant than other dom-
inants would when dominants benefit from living in a group and prefer
subordinates as company because they provide less competition (Roh-
wer and Ewald, 1981).

4.2. Between-Class Variability


Between-class plumage variability might occur in two major situ-
ations under the threat hypothesis: (1) when one class of individuals
defends resources and other classes do not (Morton, 1976), and (2) when
individuals of the species live in groups (at least part of the time) and
there is a within-group dominance hierarchy (status signaling). These
situations may apply primarily to the nonbreeding season, but indi-
viduals of many species do not defend exclusive all-purpose territories
during the breeding season and may form groups at that time, especially
64 GREGORY S. BUTCHER and SIEVERT ROHWER

for foraging (Lyon and Montgomerie, 1986; Montgomerie and Lyon,


1986).

4.3. Within-Class Variability


A major unachieved goal is to explain why some variable species
have two color classes, some three, some four, while others are contin-
uously variable (both within and between classes). Two conditions may
favor continuous variability over discrete variability under the status-
signaling hypothesis: (1) a large amount of intraclass competition (Roh-
wer et 01., 1980; Rohwer, 1982), or (2) an uncertain environment, so
that individuals with intermediate fighting ability do not know whether
it would be better to signal dominance or subordinance at the time of
the molt (Rohwer, 1982).
A major unresolved issue in understanding continuously variable
species is whether or not plumage coloration is correlated with dom-
inance within age and sex classes. Some studies have found no cor-
relation within classes (Balph et 01., 1979; Watt, 1986a,b; Jackson et
01., 1988, contra Rohwer et 01., 1981). A study on Great Tits (Porus
major) (Jarvi and Bakken, 1984) showed that bib coloration was strongly
correlated with dominance within age and sex classes. In first-year male
Purple Martins (Progne subis), Rohwer and Niles (1979) found that birds
that were more like adult males dominated birds that were more like
females.
If dominance within age and sex classes is not correlated with
plumage differences, then the individual recognition hypothesis (Col-
lias, 1943; Shields, 1977; Whitfield, 1986, 1987) may be invoked to
explain why some species are continuously variable. The individual-
recognition hypothesis argues that some species are more variable than
others because individuals of those species use plumage variability to
remember the outcome of individual encounters. Thus, long-lasting
intransitive dominance relationships should be more common in con-
tinuously variable species than in less variable species. In other words,
dominance relationships in continuously variable species should be
more stable between two individuals (because of better memory of
individuals) and should be less linear in the flock as a whole (because
individuals can use individual rather than class cues to decide whether
they can defeat an individual). None of these predictions has been tested
to date. A final note: the individual-recognition hypothesis is unable
to explain why a color characteristic might be correlated with age and
sex class; thus, the status-signaling hypothesis may be useful even in
CONSPICUOUS COLORATION 65

cases where within-class status is not correlated with plumage differ-


ences. The null hypothesis for the occurrence of continuous variability
is that it is an unselected by-product due to little or no selection for
monomorphism (Balph et 01., 1979).

5. COLOR DIMORPHISMS/POLYMORPHISMS
INDEPENDENT OF SEX, AGE, AND SEASON

A number of species of birds and other animals exhibit color di-


morphisms or polymorphisms independent of sex, age, and/or season.
Many of these involve melanism. In birds, these dimorphisms/poly-
morphisms are found in the families Procellariidae, Ardeidae, Accip-
itridae, Falconidae, Stercorariidae, Strigidae (Huxley, 1955), Cuculidae
(Payne, 1967), Laniidae (Hall et 01., 1966), and Emberizidae (Johnson
and Brush, 1972; Wunderle, 1981, 1983). To date, no generalizations
about the adaptive significance of these color patterns can be made. In
fact, in the case of the Snow Goose (Chen caerulescens), one research
group argues that there are no adaptive differences among blue, white,
or intermediate morphs (Rockwell et 01.,1985; Cooke et al., 1985; Find-
lay et 01., 1985).
If a color dimorphism/polymorphism is not selectively neutral, it
may be transitory because of the selective advantage of one color morph
compared with the other(s), or it may be stable as a result of frequency-
dependent selection or geographical differences in selective pressures.
Frequency dependence requires that each color morph be selectively
advantageous when rare relative to the equilibrium frequency at which
fitnesses are equal. When color polymorphisms are related to geograph-
ical differences, a morph-ratio cline may result. Frequency-dependent
selection for a color characteristic used in conspecific signaling has yet
to be demonstrated satisfactorily in any species, although opposing
selective pressures have been demonstrated in some and speculated
about in others.
For example, two fish polymorphisms are apparently maintained
by female preference for conspicuous morphs and by predation pressure
in favor of cryptic morphs (Semler, 1971; Moodie, 1972; Endler, 1980,
1983, 1987). Female Three-spined Sticklebacks (Gasterosteus aculea-
tus) prefer red males, perhaps because red males are better at defending
their eggs from conspecific predation (Semler, 1971). However, red
males suffer higher predation from other fish species (Moodie, 1972).
66 GREGORY S. BUTCHER and SIEVERT ROHWER

Female guppies (Poecilia reticu1ata) prefer conspicuous males with


many red and iridescent spots (Endler, 1980; Kodric-Brown, 1985) and
with large tails (Bischoff et a1., 1985). Two reasons have been proposed
for this preference. First, red spots contain carotenoids that must be
obtained in the diet; thus, red spots are an honest signal of foraging
ability (Endler, 1980; Kodric-Brown, 1985). Second, conspicuous males
are vulnerable to predation; thus, a living conspicuous male has proved
his ability to avoid predation (Endler, 1980; Bischoff et a1., 1985; Ko-
dric-Brown, 1985). On average, male guppies have fewer spots and
smaller tails in populations with more dangerous predatory fish (Endler,
1980; Endler in Bischoff et a1., 1985), indicating that predation selects
for crypsis in guppies.
Melanism often has correlated physiological effects (Ford, 1975;
Murton and Westwood, 1977; Borowski, 1981), which may lead to geo-
graphically related advantages of the morphs. There appear to be at
least three examples of this in birds. In the Rock Dove (Columba 1ivia),
melanic males breed earlier and for a longer period and are favored in
areas that support longer breeding seasons (Murton et a1., 1973). In
bananaquits (Coereba fJaveo1a), on the island of Grenada in the West
Indies, black morphs live in parts of the island with more rainfall and
spend more time in the shade than do the yellow-and-black morphs
(Wunderle, 1981). In Parasitic Jaegers (= Arctic Skuas) (Stercorarius
parasiticus), dark males have larger territories and are preferred as
mates by at least some females; light males breed at a younger age than
do dark males (O'Donald, 1980, and references cited therein). Given
these examples, one is tempted to predict important physiological dif-
ferences among dark ("blue") and light Snow Geese, but none has been
discovered in intensive studies (Cooke et a1., 1985; Findlay et a1., 1985;
Rockwell et a1., 1985).
In general, light ventral coloration is favored in aerial predators
because of its crypsis when viewed against the sky by terrestrial prey
(Mock, 1981; G6tmark, 1987; but see Kiltie, 1988). However, in many
species of raptors, there is a ventrally expressed dimorphism or poly-
morphism in coloration. The avoidance-image hypothesis (Paulson, 1973;
Rohwer, 1983; Rohwer and Paulson, 1987) suggests that potential prey
individuals of certain species may more consistently avoid predators
of the same color morph as one that has attacked them previously and
thus be more vulnerable to attack from a predator with a novel color
morpho Payne (1967) makes a parallel argument to explain the color
polymorphism of certain brood parasitic cuckoos. For hawks, four in-
direct tests support this hypothesis (Paulson, 1973; Rohwer and Paul-
son, 1987): (1) the color differences are primarily expressed ven-
CONSPICUOUS COLORATION 57

trally,the side most apparent to potential prey; (2) the color differences
are more likely to be present in species whose diet includes prey species
that are more visually acute and are more likely to modify their behavior
based on experience; (3) migratory species of Buteo are more likely to
be color polymorphic than are resident species; and (4) among poly-
morphic species of Buteo, migrants usually occur in more than two
morphs and residents usually occur in two. A direct test of the avoid-
ance-image hypothesis has yet to be performed: the hypothesis predicts
that naive prey should be more readily captured by predators with light
venters, while prey that have been attacked by light predators should
be more readily captured by predators with dark venters (Rohwer, 1983;
Rohwer and Paulson, 1987).
Mock (1981) reviews a number of hypotheses that have been ad-
vanced to explain color differences within heron species. He presents
experimental data showing that more fish were trapped around models
of white herons than around models of dark herons when the models
were placed in clear, shallow water on sunny days where herons had
previously been seen foraging. These data support the argument that
white herons are more cryptic to prey viewing them against bright skies.
However, Caldwell (1986) suggests that dark herons may be more cryp-
tic to prey in other habitats. In addition, Caldwell (1986) showed that
more hawks were attracted to white herons than to dark herons in
Panama. Kushlan (1977) showed that white models of wading birds
placed in potential foraging areas attracted more individual wading
birds than did dark models. Thus, the balance between dark and light
coloration in wading birds may be determined by the relative benefits
of conspicuousness or crypsis to conspecifics, guild members, preda-
tors, and prey in a variety of habitats.
There are strong geographic differences in color-morph ratios in
some species (see Mock, 1981 and references cited therein). In three
species, Great Blue Heron (Ardea herodias), Grey Heron (A. cinerea),
and Reddish Egret (Egretta, Dichromanassa, or Hydranassa rufescens),
white morphs appear to predominate in areas in which the opportunity
for hunting in clear, open water with bright skies predominates. Sim-
ilarly, in Reef Herons (Egretta sacra), Holyoak (1973) found the light
morph to predominate in Polynesian islands with white coral beaches
and clear water. However, to what extent different color morphs dem-
onstrate microhabitat selection within a geographical area (Mock, 1981)
remains controversial; such selectivity would be expected if white her-
ons have a hunting advantage and a relative immunity to predation in
open water.
Unlike other herons, Little Blue Herons (Egretta. Florida. or Hy-
68 GREGORY S. BUTCHER and SIEVERT ROHWER

dranassa caerulea) show an age-related color dimorphism. Caldwell


(1981) has shown that the white immatures have a foraging advantage
over the dark adults in mixed-species flocks because they are attacked
less frequently than adults by the socially dominant Snowy Egret (Egretta
thula), a white species.
Although a number of studies have addressed the relative con-
spicuousness of dark and white morphs in wading birds, the physio-
logical hypotheses have yet to be tested in either wading birds or rap-
tors.
The case of the White-throated Sparrow (Zonotrichia albicollis)
may represent a bridge between those color dimorphisms that are and
those that are not sex related. The black-and-white-striped crowns of
adults are linked with the presence of one or more 2m chromosomes;
tan-and-brown-crowned adults have only 2 chromosomes (Thorney-
croft, 1975). In an aviary study, Watt et al. (1984) found that, on average,
black-and-white-crowned males had the highest social rank, followed
by tan-and-brown males and tan-and-brown females. Black-and-white-
crowned females had the lowest social rank. If these results are sup-
ported by further study, it suggests that black-and-white crown color-
ation is favored in adult males and that tan-and-brown crown coloration
is favored in adult females. These results suggest that a genetic linkage
between the color morphs and sex would rapidly spread to fixation if
introduced into the population.
Another major question that occurs in color differences that are
not sex related is that of assortative and disassortative mating. Non-
random mating of either type may indicate selection for optimal out-
breeding (Bateson, 1983), depending on the conditions. Assortative mating
has been shown in Snow Geese (Cooke and Davies, 1983; Findlay et
al., 1985) and in Parasitic Jaegers (= Arctic Skuas) (O'Donald et aI.,
1974; O'Donald, 1959). If assortative mating is an adaptive behavior,
individuals that mate with similar individuals should have greater re-
productive success than those that mate with dissimilar individuals;
however, in Snow Geese (Cooke and Davies, 1983; Findlay et aI., 1985),
there was no evidence of differential success of like and unlike pairs.
Cooke and Davies (1983) and Findlay et al., (1985) argue that assortative
mating is merely a consequence of imprinting and that imprinting func-
tions primarily to ensure conspecific matings. Disassortative mating
has been shown in White-throated Sparrows (Lowther, 1961), Rock
Doves (Murton et aI., 1973), and the fish genus Xiphophorus (Borowski,
1981). Disassortative mating may be favored when producing either
diverse or heterozygous young, or both, is favored. In disassortative
mating, imprinting for species recognition is not a confounding vari-
CONSPICUOUS COLORA nON 69

able; thus, it seems more likely to prove a behavioral adaptation in and


of itself.

6. CONTRASTING THE HYPOTHESES

The preceding sections suggest some of the broad patterns of bird


coloration that we seek to explain. It is clear that broad patterns of bird
coloration will continue to be interpreted in different ways, until more
decisive tests of individual hypotheses are available. Thus, we devote
the remainder of this chapter to proposing the kinds of evidence that
we believe should be collected to determine which hypotheses are most
likely to explain particular intraspecific and interspecific patterns. We
present the hypotheses in terms of dichotomous choices that can be
made in determining which hypotheses are most likely to apply. First,
we ask whether coloration serves a signaling function or a physiological
function. If coloration is a signal, we ask whether it is directed primarily
to conspecifics, to potential predators, or to potential prey. If coloration
is directed to conspecifics, we ask whether we can separate breeding-
season function from nonbreeding-season function. If coloration is a
signal of species identity, is it a signal to competitors (agonistic species
recognition) or to potential mates (epigamic species recognition)? If an
adaptation to the breeding season, is the coloration directed at potential
competitors (agonistic sexual selection) or at potential mates (epigamic
sexual selection)? If a signal to potential conspecific mates, is color-
fulness an object of mate choice or of attraction? If females choose males
on the basis of their coloration, what benefit might they derive? If
females are passively attracted to males based on their coloration, are
they attracted to conspicuous males, or are they attracted to males with
an arbitrary patch of color? If a signal to potential conspecific compet-
itors, is it a signal of quality (fighting-ability hypothesis) or presence
(priority hypothesis)?

6.1. Physiology or Communication


Does the coloration have a physiological or a communicative func-
tion? This is the first of many dichotomies that may often produce the
answer "both." Given that a communicative function favors a conspic-
uous or distinctive coloration, which color is employed for that function
may well be determined by the physiological properties of the colors.
Most of the studies on the physiological properties of colors have
concentrated on black or white (Hamilton, 1973; Hailman, 1977; Wals-
70 GREGORY S. BUTCHER and SIEVERT ROHWER

berg et al., 1978; Burtt, 1979, 1981, 1986). Most of the sexual, seasonal,
and age-related color differences that we hope to explain are between
colorful (black, white, or bright hues) and cryptic (brown, gray, or olive).
Unfortunately, the above studies do not contrast the physiological dif-
ferences between black or white individuals and cryptic individuals.
A major exception is Walsberg's (1982) paper on sexual color dimorph-
ism in the Phainopepla (Phainopepla nitens). He demonstrated that the
difference in radiative heat load due to the difference between the black
male plumage and the gray female plumage was minor compared with
the differences in heat load that could be achieved in other ways (e.g.,
changes in orientation); in addition, Walsberg found that, under similar
conditions, the black males suffered a higher heat load than did the
gray females. In contrast to the expectations of Walsberg et a1. (1978),
black males had a higher heat load even in high wind conditions.
Therefore, he attributes the sexual color dimorphism to signaling re-
quirements. Given that the Phainopepla lives in the hot, arid desert of
the southwestern United States, Walsberg's result may prove to be gen-
eral for thermal colors, and in general, the all-black coloration of desert
birds may be suboptimal in the heat of the day. All-black coloration in
desert birds may be favored because it increases conspicuousness in
signaling situations or because the melanin in black feathers effectively
reduces feather abrasion from blowing sand (Burtt, 1986).
A number of physiological functions might be served by particular
colors or patterns, including improving vision, promoting optimal heat
or water balance, providing mechanical protection against abrasion,
and providing a shield against ultraviolet radiation. If analysis suggests
that a color pattern may be favored because of its physiological effects,
see the papers cited in this section (Burtt, 1981, 1986, are the most
convenient entries into this literature). For the remainder of the review,
noncryptic coloration is assumed to function as a signal.

6.2. Signaling to Conspecifics, Predators, or Prey


If coloration is used for communication, it is aimed at conspecifics,
at potential predators, or at potential prey? Most students of bird co-
loration have argued that colorfulness is a signal to conspecifics and
that the major cost of conspicuousness is increased vulnerability to
predation (Endler, 1978, and references cited therein). However, the
warning hypotheses argue that colorfulness evolved to signal to pred-
ators that colorful individuals are undesirable prey items (Cott, 1946;
Cott and Benson, 1970; Baker and Parker, 1979). Thus, the two classes
CONSPICUOUS COLORA nON 71

TABLE II
Parental Care in Monomorphic Species versus Dimorphic Species

Parental care"

Both sexes Female alone

Monomorphism 18
Dimorphism 58

"The association between sexual color monomorphism and biparental care in waterfowl (Kear, 1970).
"Numbers refer to individual species. However, almost all the monomorphic species are more closely
related to each other than they are to the dimorphic species. Thus, the association between mono-
morphism and biparental care in waterfowl might be best considered just one data point.

of hypotheses appear to make diametrically opposed predictions con-


cerning the association of conspicuousness and predation rate.

6.2,1. Are Conspicuous or Cryptic Birds More Vulnerable to


Predation?
We can think of three possible ways to test the question: are con-
spicuous or cryptic birds more vulnerable to predation?
1. Compare the vulnerability to predation of cryptic and conspic-
uous classes of birds: Baker and Parker (1979; see also Baker and Houn-
some, 1983; Baker, 1985) made an excellent case that many of the
interspecific differences in coloration can be explained by differences
in vulnerability to predation, with the more vulnerable species being
more cryptic, Specifically, they found the following patterns: (1) species
that live on the ground or on the water are less colorful and more
sexually color dimorphic than other species; (2) adults of smaller spe-
cies are less colorful and more dimorphic than adults of larger species;
(3) species with exposed nests are less colorful than species with con-
cealed nests; and (4) classes of birds that provide more parental care
are less colorful than classes of birds that provide less parental care.
Two other large interspecific surveys of sexual color dimorphism and
monomorphism, in ducks and in North American passerines, also sug-
gest that vulnerability to predation may be responsible for many of the
patterns of occurrence of conspicuous coloration. Kear (1970) found
that in most monomorphic species of waterfowl (which, in general, are
not gaudily colored), both sexes shared parental duties, whereas in most
dimorphic species (having, in general, highly colorful males), females
raised the young alone (Table II). Most male passerines in North Amer-
ica feed their young. However, males of monomorphic species are much
72 GREGORY S. BUTCHER and SIEVERT ROHWER

TABLE III
Incubation of Eggs by Males of Monomorphic and Dimorphic Species
Male
incubation°

Yes No

Monomorphism 24
Dimorphism 27

"North American passerines, association at the species level between sexual color dimorphism and
male incubation of eggs (Verner and Willson, 1969).
bNumbers refer to individual species. However, a comparative analysis using known phylogenies of
the included species has not yet been done, thus the biological persuasiveness of this correlation
has not yet been established.

more likely to incubate eggs than are males of dimorphic species (Table
III) (Verner and Willson, 1969). These results indicate that presence at
the nest selects strongly for crypsis. However, the intenstiy of sexual
selection on males is inversely proportional to the extent of male pa-
rental care (Trivers, 1972; Murton and Westwood, 1977), and differ-
ences in the relative intensity of sexual selection on the two sexes may
explain these two patterns of sexual color dimorphism and mono-
morphism.
2. Compare annual predation rates on banded (ringed) birds: Baker
and Hounsome (1983) compared 14 species of passerine birds (includ-
ing both sexes of four sexually color dimorphic species) and found that
conspicuous birds were less likely to be preyed on than cryptic birds.
Unfortunately, their sample size was small (and poorly defined).
3. Compare the proportion of cryptic and conspicuous individuals
in the diet of predators that feed on dimorphic prey with the proportion
available in the foraging areas of the predator (e.g., by studying remains
at a hacking site of cache of a predatory bird): In the springtime in
Iceland, conspicuously white male Rock Ptarmigan (Lagopus mutus)
suffer higher predation from Gyrfalcons (Falco rusticolus) than do the
cryptic brown females (Gardarsson, 1971, cited in Lyon and Montgo-
merie, 1985; O. Nielson, personal communication).
Despite the apparent promise of these three tests, "sexual selection
theory makes no particular prediction about the relation between pre-
dation rate and conspicuousness among species" (Lyon and Montgo-
merie, 1985), especially if interspecific risk of predation varies for rea-
sons other than conspicuousness, as is likely (Lyon and Montgomerie,
1985, see also Krebs, 1979). Each of these three proposed tests suffers
CONSPICUOUS COLORATION 73

from the fact that a number of factors could produce a bias of individuals
in the diet of a predator. As with any correlative study, cause and effect
are unclear. Results of any of the three tests could be consistent both
with the idea that conspicuousness advertises the unprofitability of
potential prey and with the idea that sexual selection favors more con-
spicuous individuals when predation costs are low. Clearly, what needs
to be tested is whether predators decide not to pursue certain classes
of individuals because of their conspicuous coloration. Thus, a different
approach is needed.

6.2.2. Do Predators Prefer Cryptic Prey?


The most direct way to study the warning hypotheses of Cott (1946)
and Baker and Parker (1979) is to examine the actual choices between
conspicuous and cryptic prey made by predators. This could be done
by following domestic animals (e.g., cats), radio-tagged predators, or
perhaps trained raptors. An experimenter could offer the predators a
series of choices between two potential prey items: one cryptic and one
conspicuous. Since individual predators may form search images for
particular prey types (Croze, 1970; Curio, 1976), such a study should
include a number of individual predators, with a significant number
of prey items per individual predator. The warning hypotheses predict
that predators should consistently refuse to chase conspicuous indi-
viduals.
A second important test is to compare the success rate of predators
trained to chase conspicuous birds and those trained to chase cryptic
birds. This would test Baker and Parker's (1979) idea that conspicuous
individuals ought to be more difficult to capture than cryptic individ-
uals.

6.2.3. To Whom Is a Signal Directed?


Another major way to distinguish between signaling to predators
or to conspecifics is to study the context of a signal (i.e., who is present
when the signal is given and to whom the signal is directed). This can
best be done with dynamic coloration or coloration that is restricted to
a particular part of the body. For example, male Red-winged Blackbirds
(Agelaius phoeniceus) expose their epaulets to conspecific males and
females during the breeding season when they are on their own territory
(Orians and Christman, 1968). When dangerous predators are present
nearby, when they forage away from their breeding territory, when they
intrude on occupied territories, and during the nonbreeding season,
74 GREGORY S. BUTCHER and SIEVERT ROHWER

male Red-wings cover their epaulets. Thus, both agonistic and epigamic
sexual selection seem relevant to the evolution of red epaulets, but
signaling to predators cannot apply.
Many birds have distinctive patches of color on their face, throat,
or breast. We believe that the most reasonable explanation for this
restriction of color is that the color is favored for signaling to conspe-
cifics and that it is restricted to lower the probability that it will be
seen by predators.
Another example is the phenomenon of dynamic or flash coloration
on the wings and tails of birds. Moynihan (1960) argues that these
patches are directed at flockmates and function to coordinate the takeoff
and flight of flocking birds. Baker and Parker (1979) argue that flash
coloration is directed at predators as either (1) protean defense, or (2)
perception advertisement. Hailman (1960) and others (Selander and
Hunter, 1960; Monroe, 1964) reported that flash coloration of the wings
was directed at potential prey and was used to flush prey items. Finally,
Burtt (1986) argues that flash coloration is directed at conspecifics in
the courtship displays of warblers. Baker and Parker (1979) found that
in European birds the presence of flash coloration on the upper surface
is correlated with small size and the amount of time spent on the ground
or in the water. Both are indications of vulnerability to predation and
are consistent with their unprofitable prey hypothesis. However, Baker
and Parker also found a correlation between flash coloration and non-
breeding gregariousness, as predicted by the flash-synchronization hy-
pothesis (Moynihan, 1960). The best resolution of these conflicting
hypotheses may be to observe the use of these patches in nature care-
fully, concentrating on what classes of animals are present when these
colors are flashed and how they behave. Experimentally obliterating
the patches and observing subsequent behavior and response should
also prove quite feasible.
The distinction between signaling to predators or to conspecifics
also needs to be made in studies of the function of conspicuous rump
patches in ungulates (Harvey and Greenwood, 1978, and references
cited therein). The major way to decide this question is to determine
which class of individuals is most likely to view the conspicuous rump
patch and to determine how the two classes of individuals respond
when they do see the rump patch. Similar questions can be asked of
the stotting behavior of ungulates and of the alarm calls of birds and
other animals (Harvey and Greenwood, 1978; Caro, 1986).
Surprisingly to us, Baker and Parker (1979) rejected the reasoning
that if a signal is used in an epigamic context epigamic sexual selection
CONSPICUOUS COLORATION 75

is most likely to account for its evolution. Baker and Parker (1979)
argued as follows:
When some feature of male plumage is used during solitary courtship dis-
play to a female and is concealed or less conspicuous at other times ... ,
the possibility that the feature evolved through sexual selection might seem
greater than when a feature is continuously visible. Even here, however,
the possibility that the feature evolved initially in response to predation-
risk pressures requires consideration .... The critical factor seems likely to
be whether a displaying male represents a profitable or unprofitable prey
to a predator.

We believe that Baker and Parker's argument here is untenable. If a


color pattern is displayed only to females, the rational deduction is
that male-female communication produces the important selection
pressure favoring that color pattern. Similarly, if a color pattern is
displayed only to predators, predator-prey communication is the most
likely selection pressure favoring the color pattern. Many color patches
are displayed in more than one situation, making conclusions less
straightforward.
If these sorts of analyses indicate that communication between
potential prey and predators might be important for the evolution of
certain types of bird coloration, then we refer you to the cited references
and to Hailman (1977). For the remainder of this chapter, we assume
that signaling to conspecifics (or to ecologically similar or closely re-
lated species) is the function of colorfulness.

6.3. Breeding or Nonbreeding Adaptation


If a signal to conspecifics, is coloration an adaptation to the breed-
ing or to the nonbreeding season? The major question we wish to ad-
dress here is whether we can assume that the coloration that charac-
terizes an individual in one season is independent of the selection
pressures on the individual's coloration during another season. For
those bird species with two molts a year, the answer is unequivocally
yes. Unfortunately, many species of birds do not feature two molts a
year.
Molt is an energetically costly event. Studies of the energetic cost
of postnuptial molts show that the basal metabolic rate is elevated
20-40% for the duration of the molt, typically 3-6 weeks (King, 1980;
Walsberg, 1983). It may be that the cost of two molts a year is sufficiently
high to override any benefit that some birds might achieve by producing
different plumages for different seasons. On the other hand, it may be
76 GREGORY S. BUTCHER and SIEVERT ROHWER

that the species that require different plumages are precisely those that
molt twice a year. (Some species molt twice a year without changing
color, indicating that a twice-a-year molt can be favored for other rea-
sons, presumably high feather wear.)
There is a second method (in addition to two molts a year) to
produce plumage independence: feather tipping and feather abrasion
(Darwin, 1871). Almost all birds molt after breeding, and almost all
seasonally dimorphic species are more cryptic during the nonbreeding
season than during the breeding season. A number of birds have cryptic
tips on their new feathers that wear away with time (e.g., the European
Starling and the Snow Bunting). However, most of the species that we
know lack a spring molt do not look dramatically different in spring
than in fall, because the cryptic tips are narrow. Thus, there may be a
constraint on the ability of feather tips to produce a dramatically dif-
ferent plumage in two seasons. Further research into feather abrasion
and spring molts is critical to our understanding of seasonal indepen-
dence. Burtt's (1986) technique of experimental abrasion should facil-
itate studies of feather tips.
A lot of evidence suggests that colorfulness is more important dur-
ing the breeding season than during the nonbreeding season. First,
species that are seasonally color dimorphic are more conspicuous dur-
ing the breeding season than during the nonbreeding season-the third
rule of avian color dimorphism. Second, a reversal of sex role during
the breeding season results in reversed sexual color dimorphism (Le.,
in phalaropes and the Painted Snipe). Third, colorfulness is positively
correlated with polygyny in the European avifauna (Baker and Parker,
1979), and extravagant plumage (colorfulness plus elaborate and elon-
gated feathers) is positively correlated with promiscuity/polygyny in
the Birds of Paradise (LeCroy, 1981) (Table IV).
If plumage is not seasonally independent and if sexual selection
is a major selective pressure favoring colorfulness, older birds might
be expected to be colorful year-round and younger birds cryptic year-
round. This expectation follows from the assumption that the proba-
bility of reproductive success increases with age and experience (Lack,
1954,1966,1968; Studd and Robertson, 1985) and, therefore, that older
birds should maximize their reproductive success despite some cost to
survival, whereas younger birds should maximize their survival at some
cost to reproductive success. The pattern of age dimorphism is con-
sistent with the unprofitable prey hypothesis as long as younger birds
are more vulnerable to predators than older birds. Thus, if sexual se-
lection is a major selective pressure favoring conspicuousness and if
plumage is not seasonally independent, older birds may be more con-
CONSPICUOUS COLORATION 77

TABLE IV
Polygyny and Dimorphism and Polygyny and Elaborate Plumage in Birds of
Paradise

Sexually Males with


Mating system/ color elaborate
species dimorphic? plumes?

Monogamous (presumed or known)


Macgregoria pulchra (known) No No
Lycocorax pyrrhopterus (presumed) No No
Manucodia
ater (known) No No
jobiensis (presumed) No No
chalybatus (known) No No
comrii (presumed) No No
Phonygammus keraudrenii No Two plumes
(known)
Paradigalla
carunculata (presumed) No No
brevicauda (presumed) No No

Promiscuous (male dispersion")


Semioptera wallacei (TU) Yes Yes
Orepanornis albertisi (OT) Yes Yes
Astrapia stephaniae (TL) Yes Yes
Parotia
sefilata (EL) Yes Yes
carolae (EL) Yes Yes
lawesii (EL) Yes Yes
wahnesi (EL) Yes Yes
Pteridophora alberti (EL) Yes Yes
Cicinnurus regius (EL) Yes Yes
Oiphyllodes magnificus (ONT) Yes Yes
Paradisaea
apoda (TL) Yes Yes
raggiana (TL) Yes Yes
minor (TL) Yes Yes
decora (EL) Yes Yes
rubra (TL) Yes Yes
guilielmi (TL) Yes Yes

Resource-defense polygyny
Epimachus mcyeri (OT) Yes Yes
Lophorina superba (OT) Yes Yes

Unknown
Loria loriae Yes No
Loboparadisea scricca No No
Cnemophilus macgregorii Yes One plume

(continued)
78 GREGORY S. BUTCHER and SIEVERT ROHWER

Table IV (Continued)
Sexually Males with
Mating system/ color elaborate
species dimorphic? plumes?

Ptilorus
paradiseus Yes No
victoriae Yes No
magnificus Yes No
Seleucidis melanoleuca Yes Yes
Drepanornis bruijinii Yes Yes
Epimachus fastuosus Yes Yes
Astrapia
nigra Yes Yes
splendidissima Yes No
mayeri Yes Yes
rothschildi Yes Yes
Diphyllodes respublica Yes Yes
Paradisea rudolphi Yes Yes

"Taxonomy from Morony et al. (1975).


bOata from Gilliard (1969), LeCroy (1981). Beehler (1983), and Beehler and Pruett-Jones (1983).
'The monogamous and unadorned species are all believed to be more closely related to each other
than to the other species. Thus. the data presented above are best considered to provide merely one
data point in support of a correlation between extravagant plumage and a promiscuous mating system.
dMale dispersion: TL. true lek; EL. expoloded lek; OT. dispersed territorial; ONT. dispersed nonter-
rilorial.

spicuous than would be adaptive in the nonbreeding season and youn-


ger birds may be more cryptic than would be adaptive in the breeding
season. If these predictions are true, we should be able to demonstrate
costs (in terms of predation or competition) of adult male conspicu-
ousness in the nonbreeding season and costs (in terms of reproductive
success) of young male crypsis in the breeding season. Correlative evi-
dence of the latter cost is beginning to appear, in that more conspicuous
first-year males have higher mating success than less conspicuous first-
year males in Indigo Buntings (Payne, 1982) and Black-headed Gros-
beaks (Hill, 1988) and perhaps in Pied Flycatchers (R0skaft and Jiirvi,
1983) and Darwin's Finches (Price, 1984). However, there are no ex-
perimental tests that prove that coloration is the cause of the higher
mating success (Baker, 1985).
Additional evidence that more cryptic young males are handi-
capped in breeding relative to more colorful young males is the fact
that, in three species, first-year males continue to molt in spring for up
to a month after arriving on the breeding grounds while adult males
are initiating breeding, suggesting that acquiring adult-male-like color-
CONSPICUOUS COLORATION 79

ation is more important to them than is early breeding (American Red-


start: Rohwer et aI., 1983; Indigo Bunting: Rohwer, 1986; "Baltimore"
Oriole: S. Rohwer and Manning, unpublished data).
If a species exhibits two molts per year, suggesting that plumage
coloration in winter and summer is independent, then winter may be
the preferred season for studying the adaptive significance of conspic-
uous coloration, because there are fewer competing hypotheses in win-
ter. The hypotheses that seem most relevant to winter include threat,
especially status signaling (Sections 4, 5, and 6.7). signaling within
flocks (Section 6.2.3). agonistic species recognition (Section 6.4)' or
signaling to predators or prey (Section 6.2). Most of the discussion that
follows assumes that conspicuous coloration is adaptive in the breeding
season.

6.4. Agonistic or Epigamic Species Recognition


Is coloration a signal of species identity? If so, is it a signal to
potential mates (epigamic species recognition) or to potential compet-
itors (agonistic species recognition)? The species-recognition hy-
potheses are different from most of the other hypotheses we consider
in that they primarily predict interspecific differences in coloration.
The null hypothesis to all adaptive explanations of color differences
among species is that color differences among species arise due to
random genetic drift. There are at least three selective alternatives to
the species-recognition hypothesis. First, color differences among spe-
cies might accumulate in allopatry as a result of selection pressures for
optimal conspicuousness in different environments (Hailman, 1977).
Second, color differences within a species might spread rapidly because
of an evolutionary "arms race" (Dawkins and Krebs, 1979) between
good and poor fighters (Rohwer, 1982; West-Eberhard, 1983; Weldon
and Burghardt, 1984). Under this scenario, good fighters benefit by
looking different from poor fighters and poor fighters benefit by looking
the same as good fighters. The outcome of such an arms race is prob-
lematic (see Section 6.7); nonetheless, it provides a powerful mecha-
nism for generating dramatic color differences between species (Roh-
wer, 1982; West-Eberhard, 1983; Weldon and Burghardt, 1984). Such
a mechanism might be responsible for "leapfrog" patterns of color dif-
ferences (Remsen, 1984). Third, color differences between species might
develop rapidly because of the "runaway" process of nonrandom mat-
ing (Fisher, 1930; see also Section 6.6).
There are two species-recognition hypotheses. The epigamic ver-
sion argues that species color differences evolved as premating isolating
80 GREGORY S. BUTCHER and SIEVERT ROHWER

mechanisms to prevent interspecific hybridization (Dobzhansky, 1937;


Mayr, 1942, 1965). The agonistic version argues that species color dif-
ferences evolved to prevent erroneous interspecific aggression (the "poster
color" hypothesis proposed by Lorenz, 1962, 1966) in either the breed-
ing or nonbreeding season.
Naturalists have difficulty in assessing arguments that a particular
characteristic is selectively favored because it facilitates species rec-
ognition. On the one hand, we know that many birds are capable of
fine distinctions among individuals (stranger-neighbor discrimination
in bird song; parent-offspring recognition through simple calls; stable
dominance relationships within flocks), and therefore we assume that
species recognition ought to be automatic, using whatever differences
already exist. On the other hand, we know that interspecific hybrid-
ization does occur (although we rarely know the circumstances), and
we know that birds often respond to minimal stimuli (models, feather
tufts, red spots) as though they were conspecifics. In addition, the re-
sults of numerous studies are consistent with the interpretation that a
particular song, display, or color pattern is used for species discrimi-
nation (McKaye et al., 1984; Pierotti, 1987). However, no studies have
shown that these characteristics originated because of their function as
a species identity badge (but see Burley, 1986). Finally, we have the
twin problems that specific characteristics are often redundant (E. E.
Williams and Rand, 1977), or they may be completely absent (mor-
phological sibling species). Clearly, what is needed is to examine many
species that face different species-recognition problems to look for pat-
terns in the occurrence and redundancy of characteristics that may
serve as isolating mechanisms.
In its most general form, the epigamic species-recognition hypoth-
esis may apply to all bird species. However, another form of the ar-
gument applies only to some species. Mate recognition may be innate
or learned through sexual imprinting on a caregiving individual during
a sensitive period early in life (Hailman, 1959; Immelman, 1972; Colgan,
1983). Fine distinctions among species of very similar plumage can be
learned through sexual imprinting (Schutz, 1965). Immelman (1975)
argues that males that do not care for the young (and therefore are
unavailable for imprinting) need to be distinctive to permit innate mate
recognition.
Immelman's argument that species that rely on innate mate rec-
ognition are those that are distinctively colored cannot apply to most
passerine birds. Many conspicuously colored male and female passer-
ines feed both nestlings and fledglings and are available for sexual
imprinting. However, the argument may apply to ducks (Sibley, 1957;
CONSPICUOUS COLORA nON 81

Schutz, 1965) and to many promiscuous or polygynous species in which


males are more conspicuous than females and do not care for their
offspring.
To argue that a particular color pattern originated because it facil-
itated species recognition is to argue that the color pattern originated
either (1) after secondary contact with a potentially hybridizing species
(usually a closely related congener) in a situation where hybrids are
selected against (epigamic species recognition: Remington 1968), or (2)
after contact with an unrelated aggressive species of very similar size
and form in a situation where interspecific aggression is maladaptive
(agonistic species recognition). When two populations come into con-
tact, there will be some variability within the populations. Under the
epigamic hypothesis, females that prefer males that are most distinct
from the other species will be favored, as will the males that are most
distinct. Under the agonistic hypothesis, individuals that do not fight
the other species will be favored, and individuals that are more distinct
from the other species will suffer less aggression.
The recognition badge will be selected for as long as two potentially
hybridizing and aggressing species remain in syntopy and as long as
hybridization and interspecific aggression are selected against. The spe-
cies recognition badge is an ideal system for the evolution of runaway
sexual selection (Fisher, 1930). Under the epigamic version, any in-
crease in male divergence from the other species might allow females
to choose conspecific mates faster and with fewer errors. Thus, females
might be selected to prefer extreme males, and males might be selected
to be extreme. Under the agonistic version, any increase in color di-
vergence might lower the probability of erroneous aggression. The result
could be males that are much more divergent from their syntopic con-
geners than is necessary to avoid interspecific hybridization or aggres-
sion.
If the above reasoning is valid, color character divergence should
be common where two congeners overlap, especially in areas of sec-
ondary contact. It is in areas of overlap that hybridization or aggression
will occur and provide the selective pressure for developing the specific
differences in plumage that prevent erroneous mating or fights (ac-
cording to the species-recognition hypotheses). Thus, color differences
between species should originate in areas of overlap. Whether such
differences spread throughout the range of the species will depend on
rates of gene flow and selection pressures outside the area of sympatry.
In contrast to the prediction, color character divergence is sur-
prisingly rare (Rohwer et 01., 1980). When it has been found, the di-
vergent characters often have been quite minor: eye-patch in nuthatches
82 GREGORY S. BUTCHER and SIEVERT ROHWER

(Grant, 1975), eye-ring and iris color in gulls (N. G. Smith, 1966, but
see Pierotti, 1987), throat color in partridges (Watson, 1962), epaulet
color in blackbirds (Hardy and Dickerman, 1965; Hardy, 1967; Orians
and Christman, 1968), and iris color in grackles (Peterson, 1980). Pied
Flycatchers (Ficedula hypoleuca) provide a more interesting case: where
they overlap the range of the dominant Collared Flycatcher (Ficedula
collaris), males are browner and more femalelike than in other parts of
their range (R0skaft et aI., 1986). The general failure of the prediction
of character displacement suggests that most color differences between
species were developed in allopatry as a result of genetic drift or dif-
ferential selection pressures and were not developed in sympatry as a
result of selection for improved premating reproductive isolation. This
latter selection might operate by favoring individuals that are better
able to discriminate color differences that evolved in allopatry (' 'reinforce-
ment"), but even this process is in great dispute (Butlin, 1987).
A second prediction that follows from both species-recognition
hypotheses is that syntopic congeners or family members should be
more different from each other in coloration than are allopatric con-
geners or family members. According to this prediction, sibling species
should not be syntopic or, if they are, other behavioral isolating mech-
anisms should be present. In support of the prediction, Hamilton (1973)
found that the complexity of coloration in a group of Indian Ocean reef
fish increased as the number of potentially sympatric species belonging
to the same family increased. This correlation did not work at the
generic level even though in theory, the closer the relationship, the
stronger the trend should be.
Much of the support for the species-recognition hypotheses derives
from the observations that (1) islands rarely support more than one
congener, and (2) birds on islands are often less conspicuous than their
mainland congeners (Mayr, 1942, 1972; Sibley, 1957; Grant, 1965; Mo-
reau, 1966; Lack, 1970). However, island birds may also be more con-
spicuous than their mainland counterparts (e.g., females of the genera
Petroica and Pachycephala in the southwest Pacific (Mayr, 1942) and
female Red-winged Blackbirds (Agelaius phoeniceus) on Cuba and Puerto
Rico (Ricklefs and Cox, 1972). In addition, other hypotheses may be
consistent with the patterns of plumage differences between island and
mainland birds (Selander, 1965; Zahavi, 1981; Baker and Hounsome,
1983).
The epigamic version of the species-recognition hypothesis pre-
dicts that individuals that more closely resemble syntopic congeners
than normal adults should be handicapped in getting a mate. However,
two lines of evidence suggest that birds that deviate from species-spe-
cific plumage still are able to breed. In a number of bird species, males
CONSPICUOUS COLORATION 83

breed in a sub adult plumage that is much less distinctive than the adult
male pluamge (see Rohwer et al., 1980, for North American passerines).
In addition, males of three species underwent dramatic color manip-
ulations but were still able to breed: (1) male Red-winged Blackbirds
whose epaulets were blackened so that they closely resembled Rusty
Blackbirds (Euphagus carolinus) or Brewer's Blackbirds (Euphagus
cyanocephalus) were capable of attracting conspecific mates and suc-
cessfully breeding if they were able to defend their territories (Smith,
1972); (2) male Common Yellowthroats (Geothlypis trichas) whose black
face mask was obscured were able to attract females on the same sched-
ule as were controls (Lewis, 1972); and (3) male "Bullock's" Orioles
(Icterus galbula bullockii) whose black coloration was bleached to orange
were able to attract females on the same schedule as were controls
(Butcher, 1984).
These experiments demonstrate that a dramatic change in colora-
tion may cause little problem for mate attraction. The Red-winged
Blackbird study is the most relevant, since the males were made to look
like syntopic congeners; however, it is less than definitive for two rea-
sons: (1) Peek (1972) found that his males with dyed epaulets were
unable to attract females until the dye wore off, and (2) Smith (1972)
did not compare harem sizes of experimentals and controls, even though
Red-winged Balckbirds are polygynous.
The agonistic version of the species-recognition hypothesis pre-
dicts that colorful animals should be aggressive and that intraspecific
aggression should be more common than interspecific aggression. Lor-
enz (1962, 1966) found support for both points in Caribbean coral reef
fish. Zumpe (1965) studied the behavior of eight species of conspicuous
butterfly fish (Chaetodontidae) in an aquarium and found that intra-
specific aggression was much more frequent and intense than inter-
specific aggression. In addition, she found that interspecific aggression
was often directed toward a heterospecific that was most similar in
coloration to the aggressor. Ehrlich et a1. (1977) rejected the conspecific
aggression hypothesis for coral reef fish because (1) they found no
correlation between conspicuous coloration and territoriality in coral
reef fish, and (2) when the fish they studied (Chaetodontidae) were
aggressive, they fought both heterospecifics and conspecifics. No sim-
ilar studies to these just mentioned have been performed on birds.

6.5. Agonistic or Epigamic Sexual Selection


If an adaptation to the breeding season, is the coloration directed
at potential competitors (agonistic sexual selection) or at potential
mates (epigamic sexual selection)? Students of bird song have tried to
84 GREGORY S. BUTCHER and SIEVERT ROHWER

distinguish between agonistic and epigamic functions; the answer in


many cases is that the two functions cannot be distinguished. If bird
song, which can be readily turned off and on in specific situations,
cannot be assigned an epigamic or agonistic function, it seems unlikely
that bird coloration can be assigned one or the other function. There
are a number of reasons to believe that the two functions will often be
inextricably linked. If advertising is the function of coloration or song,
mate advertising and territorial advertising are perfectly compatible
behaviors. If coloration is linked with high genetic quality, both male
and female conspecifics will benefit from using the information.
Nonetheless, there may be a number of cases in which an epigamic
or agonistic function can be assumed over the other. If a color patch is
usually displayed only to one of the sexes, communication to that sex
is the function of the coloration. If agonistic contests occur before mat-
ing occurs, the relative importance can be estimated using a modifi-
caation of techniques advocated by Arnold and Wade (1984).
Male Red-winged Blackbirds display their epaulets both to males
and to females (Orians and Christman, 1968). In two studies in which
the epaulets of territorial males were dyed black (Peek, 1972; Smith,
1972), experimental males had great difficulty defending their territo-
ries as compared with controls. When experimental birds tried to de-
fend their territories, intruding males remained for long periods and
seemed to ignore the experimental male, even when he apparently went
through all the usual maneuvers of territorial defense. In one study
(Smith, 1972), experimental males that maintained their territories had
no obvious problem attracting females (although harem sizes of exper-
imentals and controls were not compared); in the other study, exper-
imental males attracted females only after the black paint covering their
epaulets was chipped away. There are two possible explanations for
the inability of experimental males to defend their territories: (1) they
may have been mistaken for Euphagus blackbirds (Brewer's or Rusty)
that are all-black, or (2) they may have lacked a required sign stimulus
of threat. Hansen and Rohwer (1986) compared these two hypotheses
and found that Red-winged Blackbird males were able to discriminate
between mounted Brewer's Blackbirds and epaulet-blackened Red-
winged Blackbirds. Noting that intruding Red-winged Blackbirds cover
their epaulets, Hansen and Rowher concluded that experimental birds
were mistaken for intruders, i.e., they lacked the sign stimulus of threat
normally displayed by owners.
Burley et a1. (1982; see also Harvey, 1986, and references cited
therein) found that both sexes of zebra finches preferred particular
colors of plastic leg bands. Ratcliffe and Boag (1987) showed that the
CONSPICUOUS COLORATION 85

colors of leg bands did not affect male-male interactions and that female
preferences for male leg bands were not expressed when male-male
interactions were an important determinant of mating success.
Another study has contrasted the importance of epigamic versus
agonistic sexual selection, in this case not for colorfulness, but rather
for a conspicuous morphological feature. Andersson (1982b) showed
that female Long-tailed Widowbirds (Euplectes progne) mate prefer-
entially with males that have experimentally lengthened tails as com-
pared with males with norlIlal or experimentally shortened tails. An-
dersson was able to demonstrate that the improved mating success of
the long-tailed males was not attributable to increased success in
male-male competition.
A mostly observational study suggests that the colorful plumage
of adult male manakins functions in male-male competition and not
in female choice (Foster, 1987). Foster found that males in subadult
plumages successfully copulated with females when adult males were
absent, either naturally (rare) or when experimentally removed. She
also found that adult males were more aggressive to other adult males
than they were to subadult males. Thus, the subadult plumage lowered
male aggression, but not at the expense of attractiveness to females.
Under normal circumstances, male dominance relationships ensured
that subadult males did not copulate with females.
Houde (1987) found in guppies that male-male competition did
not occur over mates; thus, the brightly colored spots of male guppies
are favored only because of female preference (Houde, 1987, and ref-
erences cited therein).
We know of no other studies that were able to separate out epigamic
from agonistic sexual selection. Most studies have considered either
one or the other. We consider agonistic sexual selection first, then
epigamic.

6.5.1. Agonistic Sexual Selection


The strongest evidence to date that coloration has a threat function
comes from the nonbreeding season. When female Chaffinches (Frin-
gilla coelebs) were dyed to resemble males, the experimental birds rose
in status against both males and females, even though most males re-
lIlained dominant to the dyed females (Marler, 1955). Similarly, when
immature White-crowned Sparrows (Zonotrichia 1eucophrys) (Parsons
and Baptista, 1980; Fugle et a1., 1984) and Harris' Sparrows (Zonotri-
chia queru1a) (Rohwer, 1985) were dyed to resemble adults, they rose
in status relative to other immature birds of their species. Thus, the
86 GREGORY S. BUTCHER and SIEVERT ROHWER

signal normally correlated with sex (in Chaffinches) or age (in White-
crowned and Harris' Sparrows), hence with dominance status, was
respected independently of any other cues of status or fighting ability.
There is a lot of evidence that coloration acts as a releaser of aggres-
sion during the breeding season (Lack, 1939; Peiponen, 1960, in Eibl-
Eibesfeldt, 1970; D. Smith in the television movie, "Why Birds Sing");
however, coloration cannot be favored because it acts as a releaser of
aggression unless for some peculiar reason the colorful individual ben-
efits by being attacked (Rowland, 1982). The studies of color as a releaser
of aggression usually involve presenting a relevant patch of color to a
defending individual. If these results can be taken as evidence that
coloration has a threat function during the breeding season, it is because
we predict that if the same patch of coloration were presented to an
intruding individual, the intruder would be deterred. Recently, R0skaft
and Rohwer (1987) tested this prediction using male Red-winged Black-
bird dummies: they found that male Red-wings landed closer to dum-
mies with normal red epaulets than to dummies with enlarged red
epaulets and that male Red-wings landed closer to dummies with black-
ened epaulets than to an empty pole. When the four treatments were
presented at once, male Red-wings were more likely to land near a
dummy with blackened epaulets or an empty pole than they were to
land near either mount with epaulets. Thus, the red epaulet acts as a
deterrent to male Red-wings, whereas the black body coloration appears
to attract male Red-wings (R0skaft and Rohwer, 1987).
Evidence that coloration is important for territorial defense comes
from a number of species in which males and females defend separate
territories during part of the year and in which both sexes are colorful,
i.e., Great Tit (Lack, 1966; Fretwell, 1972), European Robin (Erithocus
rubeculo) (Lack, 1965), Red-headed Woodpecker (Melonerpes erythro-
cepholus) (Kilham, 1978), and three species of hummingbirds (Wolf,
1969, 1975; Bleiweiss, 1985). The woodpecker and the hummingbirds
are especially noteworthy because most species in their families are
sexually color dimorphic.
Lorenz (1962, 1966) found a correlation between colorfulness and
territoriality in a Caribbean community of coral reef fish. By contrast,
Ehrlich et 01. (1977) found that most species of colorful butterfly fish
(Chaetodontidae) on the Great Barrier Reef of Australia were neither
territorial nor aggressive. In addition, a literature search by Ehrlich et
01. (1977) demonstrated many species of coral reef fish (especially in
Pomacentridae) that were territorial but cryptic in coloration. Both Pe-
terman (1971) and Brown et 01. (1973) found associations between so-
ciality and coloration, but they were not the ones predicted by agonistic-
CONSPICUOUS COLORATION 87

signaling hypotheses. Peterman found that all schooling, rock-dwelling,


and coral-dwelling fishes in the Bahamas are colorful, whereas seven
of nine substrate-dwelling fishes are not. The schooling fishes are es-
sentially nonaggressive, despite being colorful. In Kenya, all eight social
species studied were colorful and essentially nonaggressive (Brown et
01., 1973).

6.5.2. Epigamic Sexual Selection


Most of the studies of female preferences for particular color pat-
terns have occurred in polymorphic species (many of these studies were
considered in Section 5), but in addition there are at least three ex-
perimental studies and one correlative study. Two of the experimental
studies were discussed above (Andersson, 1982b; Burley, 1986). The
other experimental study, involving an aviary population of weaver-
birds, demonstrated that the lack of a particular color pattern lowered
the mating success of males (Collias et 01., 1979): young females more
than older females refused to lay eggs with males that had their normally
yellow underwings dyed black. The young females often formed tem-
porary pair bonds with the dyed males, but deserted them before laying
eggs.
Hamilton and Zuk (1982) argued that brightly colored feathers should
show the effects of ectoparasites or general poor health due to disease
more readily than do duller colors. Thus, they argued that females
should choose brightly colored males that appear healthy and free of
ectoparasites. In support of their hypothesis, they found a weak but
significant correlation between bright male coloration and parasite load
in North American passerines. Read (1987) extended their study by
adding European passerines and by testing for a number of confounding
factors and reconfirmed the correlation.

6.6. Mate Choice or Mate Attraction


If a signal to potential conspecific mates, is colorfulness an object
of mate choice or an object for mate attraction? The epigamic portion
of nonrandom mating in birds can be divided into mate attraction, mate
choice, mate seeking, or forced copulation. Mate seeking and forced
copulation are both consistent with cryptic coloration, but either mate
attraction or mate choice might favor conspicuous coloration in the
attractive or chosen mate. Both mate attraction and mate choice might
operate simultaneously in a single species, so the presence of each
needs to be assessed independently.
88 GREGORY S. BUTCHER and SIEVERT ROHWER

Mate choice (sometimes referred to as "active mate choice") im-


plies, empirically, that in searching for a quality breeding situation,
females discriminate between conspecific males. As a consequence of
acquiring information and of rejecting some males, females pay at least
some small time or energetic cost for their behavior. To justify paying
this cost, females must derive some tangible benefit, which may be
environmental (Le., nonheritable characteristics such as a good nesting
territory) or genetic (Le., "good genes" that would be preferred by all
females or complementary genes that some females would choose and
others reject). A number of hypotheses, many of which are controver-
sial, have been proposed to explain the benefits that females receive
from choosing mates. These are considered in Section 6.6.1. A recent
laboratory experiment on the red flour beetle (Tribolium castaneum)
tested for a correlation between attractiveness and heritable differences
in the viability of offspring but found none (Boake, 1986). A number
of studies with the same aim will be required to evaluate whether this
process occurs in nature.
Mate attraction usually involves situations in which males arrive
on a breeding territory or an advertising site (such as a lek) and advertise
for mates. Males differ in mating success, but only because they are
differentially successful in attracting females, not because of female
choice. Thus, females do not reject males or otherwise incur costs of
mate choice, and differential male mating success can occur even if
there is no benefit to female choosiness. Other investigators have used
such terms as "mate choice," "mate preference," or "passive mate choice"
for this phenomenon; however, because differential mate attraction has
no cost or benefit to the female, these terms are paradoxical and con-
fusing. Female attraction might occur to a nonadaptive (= selectively
neutral) trait of males (Lande, 1981; Kirkpatrick, 1982, 1987) or to a
trait that renders males more conspicuous (Parker, 1982, 1983). These
possibilities are considered in Section 6.6.2.
In deciding whether either mate choice or mate attraction might
favor colorfulness in birds, it is necessary to perform a color manipu-
lation by obscuring any color pattern believed to be the object of epi-
gamic selection. If obscuring the color pattern lowers the mating success
of experimental individuals, a close look at the results should enable
a determination as to whether the color pattern was an object of choice
or attraction. Choice implies that at least some females visit and reject
some males. The number of males visited by females should be ob-
servable in a number of field and laboratory situations (Janetos, 1980;
Thornhill, 1980; Brown, 1981). If the color pattern is used as an object
of mate choice, then visitation rates of females to experimental males
CONSPICUOUS COLORATION 89

should remain high, but their mating success should be low. If con-
spicuousness were the important quality of the color patch, then a
manipulation that dramatically lowers conspicuousness should have a
greater effect on mating success of experimentals than an equally dra-
matic manipulation that has less effect on conspicuousness. If females
prefer a color patch for arbitrary reasons, then the experimental males
with that patch obscured should show low mating success, but the
system should include little "shopping" behavior among females.
These predictions are essentially untested. Butcher (1984) found
extensive "shopping" behavior in female "Bullock's" Orioles (Icterus
galbula bullockii), but they did not discriminate against males whose
black plumage had been bleached to a dull orange. The results suggest
that there is mate choice in female "Bullock's" Orioles, but that the
choice is not for black male coloration or for the black-orange color
contrast. Unfortunately, the manipulation did not significantly lower
the conspicuousness of the male orioles, nor were sample sizes suffi-
cient to detect a subtle change in the experimental males' acceptability
to females.
Houde (1987) suggested that mate choice was more important than
mate attraction in female preferences for conspicuous male guppies
because she could demonstrate no difference in the distance of attrac-
tion of females to males that differed in conspicuousness and because
females did differ in their sexual responsiveness and in their willing-
ness to copulate with males that differed in conspicuousness.

6.6.1. Mate Choice


If females choose mates based on their coloration, what evolu-
tionary benefit might they derive? Females who nest on a male's ter-
ritory should choose males with good territories, and females whose
mates feed them or their young should choose males that are good
foragers; these conclusions are not controversial. However, in these
species, females should be able to assess territory quality and foraging
ability directly, and they should not rely on male coloration as indi-
cators.
Females might also choose males on the basis of genetic comple-
mentarity. There should be an optimal level of inbreedingloutbreeding
(Bateson, 1983); in addition, there might be particular loci at which
homozygosity or heterozygosity is favored. A color patch might be an
indication of genotype in either of these cases, allowing assortative or
disassortative mating (see Section 5).
Far more controversial is the idea that a female might pick a mate
90 GREGORY S. BUTCHER and SIEVERT ROHWER

for his "good genes." The idea is that the male's good genes should be
correlated with higher survival and reproductive success of his off-
spring. A number of investigators reject the idea of mate choice for
improved offspring survival because they believe that there is little
additive genetic variance for traits that are closely tied to fitness (Ar-
nold, 1983, and references cited therein). However, even if there is a
small amoung of variance, mate choice will be favored (Maynard Smith,
1985); thus, a low rate of occurrence of slightly deleterious mutations
seems sufficient to account for mate choice for improved offspring sur-
vival (Kondrashov, 1984; Manning, 1984). Some data are beginning to
accumulate about additive genetic variance for fitness in natural pop-
ulations; to date, the data are inconclusive Oones, 1987). Partridge (1980)
showed in Drosophila melanogaster that the offspring of females given
a choice of mates were more successful than the offspring of females
who were forced to copulate with a single male. Similar results are
found in plants with and without a choice of pollen (Stephenson and
Bertin, 1983).
Even if there is additive genetic variance for survival or competitive
ability, why should conspicuousness be positively correlated with such
genes? Conspicuous coloration is energetically cheap and easy to mimic;
thus, it should not provide good information concerning genetic quality.
In addition, conspicuous coloration exposes its bearer to predators and
rivals, which should lower the survival of conspicuous individuals
relative to cryptic ones. There are at least three ways in which coloration
might facilitate mate choice within species: (1) by serving as a handicap,
(2) by advertising a display, or (3) by advertising the health of the
plumage. These are considered in the next three paragraphs.
To argue that conspicuous coloration indicates survival or com-
petitive ability is to argue that the individual has been able to survive
or be present on territory despite the handicap of being conspicuous
(Zahavi, 1981; Borgia, 1979). Thus, a conspicuous bird that is alive and
available as a potential mate is more likely to be of high genetic quality
than a similar cryptic individual because it has been subjected to a
higher probability of detection by predators (the handicap principle:
Zahavi, 1975, 1977; Kodric-Brown, 1985) or competitors (Borgia, 1979).
Models of this hypothesis require that (1) only individuals of high
quality acquire conspicuous coloration (phenotypic plasticity), (2) fe-
males choose the best among the colorful individuals, or (3) the benefit
of being preferred as a mate (the Fisher effect) is included (Andersson,
1982a; Dominey, 1983; Kodric-Brown and Brown, 1984; Nur and Has-
son, 1984; Maynard Smith, 1985; Andersson, 1986).
A male might demonstrate his health and agility in an energetically
CONSPICUOUS COLORATION 91

costly or difficult-to-perform display; conspicuous coloration may be


selected to accentuate this display. In this case, the display provides
the information about health and agility and the coloration advertises
the display.
Finally, brightly colored feathers may show the effects of ectopar-
asites or general poor health due to disease more readily than do duller
colors (Hamilton and Zuk, 1982). Females would prefer brightly colored
males because they can be more sure that they are free of ectoparasites
or other disease.
To date, the theories in this area are way ahead of the data. Direct
tests of these three ideas may be difficult to devise, but acceptance of
them will require confirmation of unique predictions in field and lab-
oratory studies.

6.6.2. Mate Attraction


If females are passively attracted to males on the basis of their
coloration, are they attracted to conspicuous males, or are they at-
tracted to males with an arbitrary patch of color? One of the most
nonintuitive results (for biologists trained in the adaptationist school)
of recent theoretical models of sexual selection has been the finding by
Lande (1981) and Kirkpatrick (1982) that females might mate prefer-
entially with males with certain characteristics even if those charac-
teristics decreased male survival and even if the females derived no
benefit at all from mating with those males.
There are two keys to making the model work: (1) females suffer
no cost from their preference for these males, and (2) the increase in
mating success of these preferred males must be greater than the loss
in survival or fecundity caused by the characteristic that also causes
the female preference. The result is nonintuitive for two reasons: (1)
the models appear to be about female preference, but female preference
has no cost or benefit, and (2) it does not seem logical that a trait that
lowers male survival should be favored by sexual selection.
Because the models involve no costs or benefits for female pref-
erences, selection acts on the male trait alone. Thus, we prefer to con-
sider these as models of the mate attracting behavior of males. This
change in focus makes the results seem more reasonable.
In reference to the second point, it is exactly this idea of a trade-
off between reproduction and survival that has become a central tenet
of life-history theory (Williams, 1966; Stearns, 1976). In life-history
theory, the amount of effort expended on current reproduction is lim-
ited primarily by the effect of such efforts on survival, hence on future
92 GREGORY S. BUTCHER and SIEVERT ROHWER

reproduction. Mate attraction is vital for successful reproduction, thus


any trait that improves an individual's success in attracting a mate will
be favored as long as the resulting increase in reproductive success
outweighs any resulting loss in survival probability or fecundity (Kirk,
patrick, 1982). This is so because fitness is a multiplicative function of
survival success times mating success times fecundity.
Parker (1982, 1983) has also modeled the mate-attracting behavior
of males under the assumption that there is no cost or benefit to females;
however, his models add one critical element missing in the Lande and
Kirkpatrick models. Parker argues that males are competing to increase
their radius of attraction to increase the chance that they will be heard
or seen by receptive females. Thus, the key element of Parker's models
is that mate attraction selects for conspicuous male traits, not just ar-
bitrary male traits.
Parker's model makes two predictions not made by the Lande and
Kirkpatrick models: (1) that exaggerated coloration should always be
in the direction of increasing conspicuousness, and (2) that a color
manipulation that affects conspicuousness should affect mating success
more severely than an equally drastic color manipulation that does not
affect conspicuousness.

6.7. Agonistic Presence or Quality


If an agonistic signal to conspecifics during the breeding season,
does coloration signal presence (the priority hypothesis) or quality (the
fighting-ability hypothesis)? Priority effect is one of a number of names
for the observation that individuals in a number of species often win
contests because they have priority over their opponents with respect
to a particular resource, such as a territory or a mate (Maynard Smith,
1982, and references cited therein). The priority effect consists of two
distinct but interrelated phenomena: (1) challenger defeat (i.e., defend-
ers defeat challengers in escalated contests), and (2) challenger inhi-
bition (i.e., challengers retreat from defenders without fully escalating).
Challenger defeat may occur for one of two categories of reasons
(Maynard Smith, 1982, and references cited therein): (1) defenders are
on average better fighters than challengers, either because defenders
are individuals that are able to pursue the resource earlier or because
a high frequency of fights early in the period of resource possession
results in only the best fighters being defenders, or (2) resources are
more valuable to defenders than to challengers because (a) defenders
know how to optimize energy gain on the territory (Davies and Houston,
1981), (b) defenders have stable relationships with their neighbors (Krebs,
CONSPICUOUS COLORATION 93

1982), or (c) defenders know less than challengers about the options
available to challengers (Rohwer, 1982).
Challenger inhibition may occur (1) because defenders would be
expected to win an escalation for either of the two above reasons, (2)
because fights are so risky that a challenger would be better off to forego
a challenge even if he could win (Rohwer, 1982; Maynard Smith, 1982),
(3) because there are good alternatives to fighting defenders, or (4)
because of an "arbitrary" asymmetry "defender wins" that is uncor-
related with fighting ability or resource value (Maynard Smith, 1982).
Good alternatives to fighting defenders occur when there is excess suit-
able habitat (resulting in the ideal free distribution of Fretwell, 1972),
or when defenders frequently abandon resources (for reasons other than
losing fights; e.g., Davies, 1978). The priority effect may be reinforced
when information about relative fighting ability or differential valuation
of resources is not available because of typical intensity displays or
plumage monomorphism. When fighting is risky (point 2) but the al-
ternatives are poor (the opposite of point 3), a "war of attrition" may
be initiated by the challenger rather than escalation or abandonment
of the resource to the defender.
The threat hypothesis assumes that coloration improves the success
of colorful individuals in agonistic interactions. Coloration is advan-
tageous for defenders when it improves challenger defeat or challenger
inhibition. Threat coloration must work in one of three ways: by low-
ering the rate of agonistic encounters of colorful individuals, by causing
an encounter to be shorter or less escalated, or by improving the like-
lihood that colorful individuals will win an escalated fight. Unlike
antlers or other morphological features, coloration cannot directly im-
prove its bearer's fighting ability; therefore, coloration is unlikely to
improve challenger defeat; rather, it must aid in challenger inhibition.
Coloration may act as a distance signal of threat (favoring con-
spicuousness) or as a sign stimulus of threat (favoring distinctiveness).
If conspicuousness is most important, threat coloration should operate
by lowering the number of encounters with competitors. If distinctive-
ness is most important, threat coloration should operate by reducing
repeat encounters, by shortening the length of encounters or by low-
ering the probability that a competitor will escalate against a colorful
individual.
Why should coloration improve challenger inhibition? The present
authors split on this issue: one of us (Rohwer, 1982) supports the fight-
ing-ability hypothesis (F AH), whereas Butcher favors a priority hy-
othesis (PH) (as outlined later in this section).
By the F AH, colorfulness is favored in good fighters when distinc-
94 GREGORY S. BUTCHER and SIEVERT ROHWER

tive coloration renders fight winners more memorable, lessening the


likelihood that challengers will repeat a fight with a distinctive winner.
The FAH should apply best to species in which fighting is risky; in
these species, challengers should collect information about a large num-
ber of defenders (to determine the worst fighter) and fight only rarely
(to avoid the risks). As the colorful badge of fighting ability spreads,
challengers might avoid fights with all badge bearers after one defeat
with a badge bearer. As a result, all defenders in a species may become
colorful because poor fighters mimic good fighters to avoid being sin-
gled out by challengers. As a result, defenders should be conspicuous
and largely monomorphic. However, exceptionally good fighters are
selected to diverge from the appearance of the mimics to restore the
benefits of an honest badge of fighting ability. Therefore, different pop-
ulations (species) should be very different in plumage because new
patches (of arbitrary color and location) are constantly being favored
in good fighters of a population and mimicked by poor fighters of that
population. This process could be responsible for the "leapfrog" pattern
of color differences among subspecies in the Andes Mountains of South
America (Remsen, 1984) and for the very different appearance of many
pairs of hybridizing species.
By contrast, the PH argues that colorfulness is favored in any in-
dividual that benefits from the priority effect, regardless of relative
fighting ability. Thus, colorfulness, within the context of defense of
resources, is a badge of priority that works whenever the priority effect
works. Colorfulness serves as either a distance signal or as a sign stim-
ulus of threat.
The two hypotheses make a number of exclusive predictions. The
FAH predicts that a color badge should be most likely to evolve when
challenger inhibition and challenger defeat are not complete (because
some defenders are better able to defeat challengers than others). Ac-
cording to the FAH, defenders should be cryptic if challenger inhibition
is complete because priority should be the asymmetry used to decide
the contest, not coloration. By contrast, the PH asserts that color is the
badge of priority and should be most beneficial when challenger in-
hibition is most complete and there are few escalations. The FAH pre-
dicts that if an artificial color badge is added to a defender that is a
good fighter, he should face fewer challengers (especially fewer repeat
visits from challengers); if a badge is added to a defender that is a poor
fighter, he should face more challengers (more repeat visits). The PH
predicts that in both cases, the number of challengers should stay the
same or decline, since they are both defenders and both should benefit
CONSPICUOUS COLORATION 95

from advertising their priority. If intrusion ratef! were to rise in both


groups, a species recognition function is suggested.

7. CONCLUSIONS

The study of bird coloration is exciting because the theories of bird


coloration derive from such a number of different fields in biology. This
review would have been much different without recent advances in
our understanding of thermal physiology, aggression, sexual selection,
and communication. Research on bird coloration has the potential to
contribute to each of these fields.
On the other hand, the large number of potential explanations for
color patterns has probably inhibited progress because of the inherent
difficulty in addressing a large number of hypotheses simultaneously.
Most studies of bird coloration to date have advocated a single hy-
pothesis, without seriously testing that hypothesis against alternatives.
However, the review dispelled any notions we may have had that a
single hypothesis might account for a wide variety of color patterns by
accounting for the costs and benefits of conspicuous coloration under
various circumstances. We hope we have suggested a number of ap-
proaches to testing a large number of the available hypotheses. We argue
that conspicuousness or distinctiveness are the essential qualities for
signals, despite the difficulty of quantifying these qualities in nature.
We present three major empirical rules of avian color dimorphism:
When there are differences, males are more colorful than females, adults
are more colorful than young birds, and breeding-season birds are more
colorful than non breeding-season birds. These patterns indicate that
some form of sexual selection must be an important selective factor
favoring colorfulness. Thus, we spend much of the review covering a
variety of sexual selection hypotheses in detail. The three rules are also
consistent with the threat hypothesis; that coloration helps individuals
defend resources during the breeding and nonbreeding seasons. The
threat hypothesis may also help explain monomorphic species and
species with plumage variability within sex and age classes.
A large number of species feature color dimorphisms or polymor-
phisms that are independent of sex, age, and season. These species
provide many of the best animals for studying the variety of selective
pressures of plumage coloration; these studies confirm that a wide va-
riety of selective pressures are important.
A lot of evidence suggests that black and white provide physio-
96 GREGORY S. BUTCHER and SIEVERT ROHWER

logical benefits to their bearers. However, few studies have compared


black and white birds with similar cryptically colored individuals. Thus,
the interrealtionship between physiological colors and signaling colors
is still poorly understood.
A recent monograph (Baker and Parker, 1979) was dedicated to the
idea that most conspicuous coloration functions to signal predators that
the colorful individuals are unprofitable prey. Most of their arguments
support the idea that conspicuous birds are less vulnerable to predation
than are cryptic birds, but this correlation is consistent with the sexual
selection hypotheses. To date there have been no tests of the two central
ideas of the unprofitable-prey hypothesis: (1) that predators choose not
to pursue conspicuous prey, and (2) that conspicuous prey are more
difficult to catch than cryptic prey. Paying attention to the context of
colorful displays should help determine whether colorfulness is aimed
at conspecifics, predators, or prey.
Very few species of birds molt twice a year. Thus, plumage worn
during one season may be primarily an adaptation to another season.
If sexual selection is the major force favoring colorfulness and predation
is the major force favoring crypsis, older birds of species that are colorful
year-round may be too colorful in the nonbreeding season for maximal
survival, and younger birds that are relatively cryptic year-round may
be too cryptic in the breeding season for maximal breeding success.
Conspicuous coloration has traditionally been considered as an
isolating mechanism, preventing interspecific hybridization. However,
little evidence and little theory exist to support this idea. Rather, species
isolation appears to be an unselected by-product of color differences
that arise during allopatry.
Both agonistic and epigamic sexual selection seem to favor col-
orfulness in some species. A few field and laboratory studies have been
able to separate out the two, but in many species, both will prove
important. Both forms of sexual selection for conspicuous coloration
provide strong theoretical challenges in explaining precisely why col-
orfulness is favored.
There are three major epigamic sexual selection arguments, al-
though much of the recent discussion has focused on just two. Color-
fulness might help females choose males with "good genes" for a variety
of reasons, or females might choose males with an arbitrary color badge
despite detrimental effects on offspring survival. The third alternative
is that males might compete among themselves to be more conspicuous
to females, with females mating more frequently with the more con-
spicuous males. Theoretical models exist for all three possibilities, but
very few data are available from the field or laboratory to determine
CONSPICUOUS COLORATION 97

whether any of these mechanisms are likely to be important in nature.


We expect that a number of these studies are probably already under-
way, and developments in this area should be dramatic during the next
few years.
We have initiated a debate about how colorfulness might aid re-
source defenders within relatively monomorphic classes of individuals.
Rohwer favors the fighting-ability hypothesis, which states that good
fighters benefit by having a unique badge of fighting ability, but poor
fighters benefit by mimicking good fighters. Butcher favors the priority
hypothesis, which states that all individuals that can win contests just
because of their priority over the contested resource should advertise
their priority through conspicuous coloration.
Our review of the status of research on bird coloration confirms
that a number of fascinating questions remain. Are conspicuous birds
signaling to predators? Are some classes of birds stuck in a suboptimal
plumage for half the year because of the cost of molt? Does conspicuous
male coloration help females choose "good genes" for survival? Are
there delicately balanced opposing selective pressures that can account
for the maintenance of plumage polymorphisms or dimorphisms that
are unrelated to age or sex? We hope that the next few years will produce
answers to these and other questions that will dramatically increase
our understanding of the importance of conspicuous plumage colora-
tion-and extravagant morphological structures-in the lives of birds.

ACKNOWLEDGMENTS. Gordon Orians has provided helpful comments


on a number of drafts of this manuscript. George Williams, Michael
Beecher, and Allen Rutberg provided detailed and helpful comments
on an early version; William Langbauer critiqued a more recent version.
Malte Anderson, Mercedes Foster, Torbjorn Jarvi and Morten Bakken,
Anders Moller, Geoffrey Parker, Trevor Price, and William Searcy pro-
vided unpublished manuscripts. We thank all these individuals for
their help.

REFERENCES
A. O. U., 1983. Check-Jist of North Americon Birds. 6th ed .. American Ornithologists'
Union. Washington. D.C.
Andersson. M .. 1982a. Sexual selection. natural selection. and quality advertisement.
BioI. J. Linn. Soc. 17:375-393.
Andersson. M .. 1982b. Female choice selects for extreme tail length in a widowbird.
Noture (Lond.) 299:818-820.
98 GREGORY S. BUTCHER and SIEVERT ROHWER

Andersson, M., 1983, On the functions of conspicuous seasonal plumages in birds, Anim.
Behav.31:1262-1264.
Andersson, M., 1986, Evolution of condition-dependent sex ornaments and mating pref-
erences: Sexual selection based on viability differences, Evolution 40:804-816.
Arnold, S. J., 1983, Sexual selection: The interface of theory and empiricism, in: Mate
Choice (P. Bateson, d.), pp. 67-107, Cambridge University Press, Cambridge.
Arnold, S. J., and Wade M. J., 1984, On the measurement of natural and sexual selection:
Applications, Evolution 38:720-734.
Bailey, S. F., 1978, Latitudinal gradients in colors and patterns of passerine birds, Condor
80:372-381.
Baker, R R, 1985, Bird coloration: In defence of unprofitable prey, Anim. Behav. 33:
1387-1388.
Baker, R R., and Hounsome, M. V., 1983, Bird coloration: Unprofitable prey model
supported by ringing data, Anim. Behav. 31:614-615.
Baker, R E., and Parker, G. A., 1979, The evolution of bird coloration, Philos. Trans. R.
Soc. Lond. 8287:63-130.
Balph, M. H., Balph, D. F., and Romesburg, H. C., 1979, Social status signaling in winter
flocking birds: An examination of a current hypothesis, Auk 96:78-93.
Bateson, P., 1983, Optimal outbreeding, in: Mate Choice (P. Bateson, ed.), Cambridge
University Press, Cambridge, pp. 257-277.
Beehler, B., 1983, Frugivory and polygamy in birds of paradise, Auk 100:1-12.
Beehler, B., and Pruett-Jones, S. G., 1983, Display dispersion and diet of birds of paradise:
A comparison of nine species, Behav. Ecol. Sociobiol. 13:229-238.
Bischoff, R J., Gould, J. L., and Rubenstein, D. 1., 1985, Tail size and female choice in
the guppy (Poecilia reticulata), Behav. Ecol. Sociobiol. 17:253-256.
Bleiweiss, R, 1985, Iridescent polychromatism in a female hummingbird: Is it related to
feeding strategies?, Auk 102:701-713.
Boake, C. R B., 1986, A method for testing adaptive hypotheses of mate choice, Am.
Nat. 127:645-666.
Borgia, G., 1979, Sexual selection and the evolution of mating systems, in: Sexual Se-
lection and Reproductiv8 Competition in Insects (M. S. Blum and N. A. Blum, eds.),
Academic, New York, pp. 19-80.
Borowski, R, 1982, Tailspots of Xiphophorus and the evolution of conspicuous poly-
morphism, Evolution 35:345-358.
Brian, A. D., 1949, Dominance in the Great Tit (Parus major), Scott. Nat. 61:144-155.
Brown, J. H., Cantrell, M. A., and Evans, S. M., 1973, Observations on the behaviour and
coloration of some coral reef fish (Family: Pomacentridae), Mar. Behav. Physiol. 2:
63-71.
Brown, J. L., 1963, Aggressiveness, dominance and social orgainization in the Stellar Jay,
Condor 65:460-484.
Brown, L., 1981, Patterns offemale choice in mottled sculpins (Cottidae, Teleostei), Anim.
Behav. 29:375-382.
Burley, N., 1981, The evolution of sexual indistinguishability, in: Natural Selection and
Social Behavior (R D. Alexander and D. W. Tinker, eds.), Chiron, New York, pp.
121-137.
Burley, N., 1986, Comparisons of the band-colour preferences of two species of estrildid
finches, Anim. Behav. 34:1732-1741.
Burley, N., Krantzberg, G., and Radman, P., 1982, Influence of colour-banding on the
conspecific preferences of zebra finches, Anim. Behav. 30:444-455.
Burtt, E. H., Jr., 1979, Tips on wings and other things, in: The Behavioral Significance
CONSPICUOUS COLORATION 99

of Color (E. H. Burtt, Jr., ed.), Garland STPM Press, New York and London, pp.
75-110.
Burtt, E. H., Jr., 1981, The adaptiveness of animal colors, Bioscience 31:723-729.
Burtt, E. H., Jr., 1986, An analysis of physical, physiological, and optical aspects of avian
coloration with emphasis on wood-warblers, Ornithol. Monog. [AOUl 38:1-126.
Butcher, G. S., 1984, Sexual color dimorphism in orioles (the Genus Icterus): Tests of
communication hypotheses, unpublished Ph.D. dissertation, University of Wash-
ington, Seattle.
Butlin, R., 1987, Speciation by reinforcement, Trends Ecol. Evol. 2:8-13.
Caldwell, G. S., 1981, Attraction to mixed-species heron flocks, Behav. Ecol. Sociobiol.
8:99-103.
Caldwell, G. S., 1986, Predation as a selective force on foraging herons: Effects of plumage
color and flocking, Auk 103:494-505.
Caro, T. M., 1986, The functions of stotting in Thomson's gazelles: Some tests of the
predictions, Anim. Behav. 34:663-684.
Chen, D.-M., and Goldsmith, T. H., 1986, Four spectral classes of cone in the retinas of
birds, J. Compo Phys. A159:473-479.
Clements, J., 1981, Birds of the World: A Checklist, 3rd ed., Facts on File, New York.
Clutton-Brock, T. H., and Harvey, P. H., 1984, Comparative approaches to investigating
adaptation, in: Behavioural Ecology: An Evolutionary Approach 0. R. Krebs and N.
B. Davies, eds.), 2nd ed., Blackwell Scientific Publications, Oxford, pp. 7-29.
Colgan, P., 1983, Comparative Social Recognition, Wiley, New York.
Collias, E. c., Collias, N. E., Jacobs, c. H., McAlary, F., and Fujimoto, J. T., 1979, Ex-
perimental evidence for facilitation of pair formation by bright color in weaverbirds,
Condor 81:91-93.
Collias, N. E., 1943, Statistical analysis of factors which make for success in initial
encounters between hens, Am. Nat. 77:519-538.
Cooke, F., and Davies, J. c., 1983, Assortative mating, mate choice, and reproductive
fitness in Snow Geese, in: Mate Choice (P. Bateson, ed.), Cambridge University Press,
Cambridge, pp. 279-295.
Cooke, F., Findlay, C. S., Rockwell, R. F., and Smith, J. A., 1985, Life history studies in
the Lesser Snow Goose (Ansel' caerulescens caerulescens). III. The selective value
of plumage polymorphism: new fecundity, Evolution 39:165-177.
Cott, H. B., 1946, The edibility of birds, Proc. Zool. Soc. Lond. 116:371-524.
Cott, H. B., 1985, Palatability of birds and eggs, in: A Dictionary of Birds [B. Campbell
and E. Lack, eds.], Poyser, Calton, p. 427.
Cott, H. B., and Benson, C. W., 1970, The palatability of birds, mainly based upon
observations of a tasting panel in Zambia, Ostrich Suppl. 8:357-384.
Coutlee, E. 1., 1967, Agonistic behavior in the American Goldfinch, Wilson Bull. 79:
89-100.
Cracraft, J., 1981, Toward a phylogenetic classification of the recent birds of the world
(Class Aves), Auk 98:681-714.
Crook, J. H., 1964, The evolution of social organization and visual communication in the
weaverbirds (Ploceinae), Behaviour Suppl. 10:1-178.
Croze, H., 1970, Searching image in Carrion Crows, Z. Tierpsychol. Beiheft 5:1-86.
Curio, E., 1976, The Ethology of Predation, Springer-Verlag, Berlin.
Darwin, c., 1871, The Descent of Man, and Selection in Relation to Sex, John Murray,
London (1890, 2nd ed.)
Davies, N. B., 1978, Territorial defence in the speckled wood butterfly (Pararge aegeria),
the resident always wins, Anim. Behav. 26:138-147.
100 GREGORY S. BUTCHER and SIEVERT ROHWER

Davies, N. B., and Houston, A. I., 1981, Owners and satellites: The economics of territory
defense in the Pied Wagtail, Motacilla alba, J. Anim. Ecol. 50:157-180.
Dawkins, R., and Krebs, J. R., 1979, Arms races between and within species, Proc. R.
Soc. Lond. B Biol. Sci. 205:489-504.
Dilger, W. C., 1956, Hostile behavior and reproductive isolating mechanisms in the avian
genera Catharus and Hylocichla, Auk 73:313-353.
Dilger, W. C., 1960, Agonistic and social behavior of captive red polls, Wilson Bull. 72:
114-132.
Dixon, K. 1., 1963, Some aspects of social organization in the Carolina Chickadee, Pro-
ceedings XIII International Ornithological Congress, June 17-24, 1962, C. G. Sibley,
ed., American Ornithologists' Union, Baton Rouge, Louisiana, pp. 240-258.
Dobzhansky, T., 1937, Genetics and the Origin of Species, Columbia University Press,
New York.
Dominey, W. J., 1983, Sexual selection, additive genetic variance and the "phenotypic
handicap," ]. Theor. Bioi. 101:495-502.
Dwight, J., 1900, The sequence of plumages and moults of the passerine birds of New
York, Ann. NY Acad. Sci. 13:73-360 (1975: repr.).
Ehrlich, P. R., Talbot, F. H., Russell, B. C., and Anderson, G. R. V., 1977, The behaviour
of chaetodontid fishes with special reference to Lorenz' "poster coloration" hypoth-
esis, J. Zool. Lond. 183:213-228.
Eibl-Eibesfeldt, 1.,1970, Ethology: The Biology of Behavior, Holt, Rinehart, and Winston,
New York.
Endler, J. A., 1978, A predator's view of animal color patterns, in: Evolutionary Biology,
Vol. 11 (M. K. Hecht, W. C. Steere, and B. Wallace, eds.), Plenum, New York, pp.
319-364.
Endler, J. A., 1980, Natural selection on color patterns in Poecilia reticulata, Evolution
34:76-91.
Endler, J. A., 1983, Natural and sexual selection on color patterns in poeciliid fishes,
Environ. Biol. Fishes 9:173-190.
Endler, J. A., 1987, Predation, light intensity, and courtship behaviour in Poecilia reti-
culata (pisces: Poeciliidae), Anim. Behav. 35:1376-1385.
Ewald, P. W., and Rohwer, S. A., 1980, Age, coloration, and dominance in nonbreeding
hummingbirds: A test of the asymmetry hypothesis, Behav. Ecol. Sociobiol. 7:273-279.
Findlay, C. S., Rockwell, R. F., Smith, J. A., and Cooke, F., 1985, Life history studies of
the Lesser Snow Goose (Anser caerulescens caerulescens). VI. Plumage polymor-
phism, assortative mating, and fitness, Evolution 39:904-914.
Fisher, R. A., 1930, The Genetical Theory of Natural Selection, Clarendon Press, Oxford
(1958, 2nd ed., Dover, New York).
Forbush, E. H., and May, J. B., 1929, Birds of Massachusetts and Other New England
States. III. Land Birds from Sparrows to Thrushes, Norwood Press, Norwood
Massachusetts.
Ford, E. B., 1975, Ecological Genetics, 4th ed., Chapman and Hall, London; Wiley, New
York.
Foster, M. S., 1987, Delayed maturation, neoteny, and social system differences in two
manakins of the genus Chiroxiphia, Evolution 41:547-558.
Fretwell, S. D., 1972, Populations in a Seasonal Environment, Princeton University Press,
Princeton, New Jersey.
Fugle, G. N., Rothstein, S. I., Os enberg, C. W., and McGinley, M. A., 1984, Signals of
CONSPICUOUS COLORATION 101

status in wintering White-crowned Sparrows, Zonotrichia leucophrys gambelii, Anim.


Behav. 32:86~93.
Gardarsson, A., 1971, Food ecology and spacing behavior of Rock Ptarmigan Lagopus
mutus in Iceland, unpublished Ph.D. dissertation, University of California, Berkeley.
Gilliard, E. T., 1969, Birds of Paradise and Bower Birds, Weidenfeld and Nicolson,
London; Natural History Press, Garden City, New Jersey.
Glase, J., 1973, Ecology of social organization in the Black-capped Chickadee, Living Bird
12:235~267.

Goodwin, D., 1951, Some aspects of the behaviour of the jay Garrulus glandarius, Ibis
93:414~442, 602~625.
Goodwin, D., 1960, Sexual dimorphism in pigeons, Bull. Br. Ornithol. Club 80:45~52.
G6tmark, F., 1987, White underparts in gulls function as hunting camouflage, Anim.
Behav. 35:1786~1792.
Gould, S. J., and Lewontin, R. c., 1979, The spandrels of San Marco and the Panglossian
paradigm: A critique of the adaptationist programme, Proc. R. Soc. Lond. B 205:
581~598.
Grant, P. R., 1965, Plumage and the evolution of birds on islands, Syst. Zool. 14:47~52.
Grant, P. R., 1975, The classic case of character displacement, in: Evolutionary Biology,
Vol. 8 (T. Dobzhansky, M. K. Hecht, and W. C. Steere, eds.), Plenum, New York, pp.
237~337.
Hailman, J. P., 1959, Why is the male Wood Duck strikingly colorful?, Am. Nat. 93:
383~384.
Hailman, J. P., 1960, A field study of the Mockingbird's wing-flashing behavior and its
association with foraging, Wilson Bull. 72:346~357.
Hailman, J. P., 1967, Spectral discrimination: An important correction, J. Opt. Soc. Am.
57:281~282.
Hailman, J. P., 1976, Uses of the comparative study of behavior, in: Evolution, Brain,
and Behavior: Persistent Problems (R. B. Masterton, W. Hodos, and H. Jerison, eds.),
Erlbaum, Hillside, New Jersey, pp. 13~22.
Hailman, J. P., 1977, Optical Signals: Animal Communication and Light, Indiana Uni-
versity Press, Bloomington.
Hailman, J. P., 1979, Environmental light and conspicuous colors, in: The Behavioral
Significance of Color (E. H. Burtt, Jr., ed.), Garland STPM Press, New York, pp.
289~354.

Hall, B. P., Moreau, R. E., and Galbraith, I. C. J., 1966, Polymorphism and parallelism in
the African bush-shrikes of the genus Malaconotus (including Chlorophoneus), Ibis
108:161~182.
Hamilton, T. H., 1961, On the functions and causes of sexual dimorphism in breeding
plumage of North American species of warblers and orioles, Am. Nat. 95:121~123.
Hamilton, T. H., and Barth, R. H., Jr., 1962, The biological significance of seasonal change
in male plumage appearance in some New World migratory species, Am. Nat. 96:
129~144.

Hamilton, W. D., and Zuk, M., 1982, Heritable true fitness and bright birds: A role for
parasites?, Science 218:384~387.
Hamilton, W. J. III, 1973, Life's Color Code, McGraw-Hill, New York.
Hamilton, W. J. III, and Peterman, R. M., 1971, Counters hading in the colourful reef fish
Chaetodon lunula: Concealment, communication, or both?, Anim. Behav. 19:357~364.
Hansen, A. J., and Rohwer, S., 1986, Coverable badges and resource defence in birds,
Anim. Behav. 34:69~76.
102 GREGORY S. BUTCHER and SIEVERT ROHWER

Hardy, J. W., 1967, Evolutionary and ecological relationships between three species of
blackbirds (Icteridae) in central Mexico, Evolution 21:196-197.
Hardy, J. W., and Dickerman, R. W., 1965, Relationships between two forms of Red-
winged Blackbirds in Mexico, Living Bird 4:107-130.
Harvey, P. H., 1986, Birds, bands, and better broods?, Trends Ecol. Evol. 1:8-9.
Harvey, P. H., and Greenwood, P. H., 1978, Anti-predator defence strategies: Some ev-
olutionary problems, in: Behavioural Ecology: An Evolutionary Approach (J. R. Krebs
and N. B. Davies, eds.), Blackwell Scientific Publications, Oxford, pp. 129-15l.
Hilborn, R., and Stearns, S. C., 1982, On inference in ecology and evolutionary biology:
The problem of multiple causes, Acta Biotheor. 31:145-164.
Hill, G. E., 1988, The function of delayed plumage maturation in male Black-headed
Grosbeaks, Auk 105:1-10.
Hinde, R. A., 1956, A comparative study of the courtship of certain finches (Fringillidae),
Ibis 98:1-23.
Hingston, R. W. G., 1933, The Meaning of Animal Colour and Adornment, E. Arnold,
London.
Holyoak, D. T., 1973, Significance of colour dimorphism in Polynesian populations of
Egretta sacra, Ibis 115:419-420.
Houde, A. E., 1987, Mate choice based on naturally occurring color-pattern variation in
a guppy population, Evolution 41:1-10.
Huxley, J. S., 1938, Threat and warning coloration in birds, with a general discussion of
the biological functions of color, in: Proceedings of the Eighth International Orni-
thological Congress, F. C. R. Jourdain, ed., The University Press, Oxford, pp. 430-455.
Huxley, J. S., 1955, Morphism in birds, Acta Int. Ornithol. Congr. XI, July 1934, A.
Portman and E. Sutter, eds., Birkhauser Verlag, Basel and Stuttgart, Switzerland, pp.
309-328.
Immelman, K., 1972, The influence of early experience upon the development of social
behaviour in estrildine finches in: Proceedings of the XV International Ornithological
Congress August 30-September 5, 1970, (K. H. Voous, ed.), E. J. Bull, Leiden, Neth-
erlands, pp. 316-338.
Immelman, K., 1975, The evolutionary significance of early experience, in: Function and
Evolution in Behaviour (G. Baerends, C. Beer, and A. Manning, eds.), Clarendon,
Oxford, pp. 243-253.
Jackson, W. M., Rohwer, S., and Winnegrad, R. L., 1988, Status signaling is absent within
age-and-sex classes of Harris' Sparrows, Auk 105:424-427.
Janetos, A. C., 1980, Strategies of female choice: A theoretical analysis. Behav. Ecol.
Sociobiol. 7:107-112.
Jarvi, T., and Bakken, M., 1984, The function of the variation in the breast stripe of the
great tit, Anim. Behav. 32:590-596.
Johnson, N. K., and Brush, A. H., 1972, Analysis of polymorphism in the Sooty-capped
Bush Tanager, Syst. Zool. 21:245-262.
Jones, J. S., 1985, The heritability of fitness: Bad news for good genes?, Trends Ecol. Evol.
2:35-38.
Kear, J., 1970, The adaptive radiation of parental care in waterfowl, in: Social Behaviour
in Birds and Mammals: Essays on the Social Ethology of Animals and Man (J. H.
Crook, ed.), Academic, London, pp. 357-39l.
Ketterson, E. D., 1979, Aggressive behavior in wintering dark-eyed juncos: Determinants
of dominance and their possible relation to geographic variation in the sex ratio,
Wilson Bull. 91:371-383.
CONSPICUOUS COLORATION 103

Kikkawa, j., 1961, Social behavior of the white-eye Zosterops latera lis in winter Hocks,
Ibis 103:428-442.
Kilham, 1., 1974, Early breeding season behavior of Downy Woodpeckers, Wilson Bull.
86:407-418.
Kilham, 1., 1978, Sexual similarity of Red-headed Woodpeckers and possible explana-
tions based on fall territorial behavior, Wilson Bull. 90:285-287.
Kiltie, R. A., 1988, Countershading: Universally deceptive or deceptively universal'!,
Trends Ecol. Evol. 3:21-24.
King. j. R., 1980, Energetics of avain moult, in: Acta XVII Congr. Internatl. Ornithol., R.
Niihring, ed., Verlag der Deutschen Ornithologen Gesellschaft, Berlin, pp. 312-317.
Kirkpatrick, M., 1982, Sexual selection and the evolution of female choice, Evolution 36:
1-12.
Kirkpatrick, M., 1987, Sexual selection by female choice in polygynous animals, Annu.
Rev. Ecol. Syst. 18:43-70.
Knapton, R. W., and Krebs, j. R., 1976, Dominance hierarchies in winter Song Sparrows,
Condor 78:567-569.
Kodric-Brown, A., 1985, Female preference and sexual selection for male coloration in
the guppy (Poecilia reticulata), Behav. Ecol. Sociobiol. 17:199-206.
Kodric-Brown, A., and Brown j. H., 1984, Truth in advertising: The kinds of traits favored
by sexual selection, Am. Nat. 124:309-323.
Kondrashov, A. S., 1984, Deleterious mutations as an evolutionary factor.!. The advantage
of recombination, Genet. Res. 44:199-217.
Krebs, j. R., 1979, Bird colours, Nature (Lond.) 282:14-16.
Krebs, j. R., 1982, Territorial defence in the Great Tit (Parus major): Do residents always
win?, Behav. Ecol. Sociobiol. 11:185-194.
Kushlan, j. A., 1977, The significance of plumage colour in the formation of feeding
aggregations of ciconiiforms, Ibis 119:361-364.
Lack, D., 1939, The behaviour of the robin. I, II. Proc. Zool. Soc. Lond. A 109:169-178.
Lack, D., 1954, The Natural Regulation of Animal Numbers, Clarendon. Oxford.
Lack, D., 1965, The Life of the Robin, Collins, London.
Lack, D., 1966, Population Studies of Birds, Clarendon, Oxford.
Lack, D., 1968, Ecological Adaptations for Breeding in Birds, Methuen, London.
Lack, D., 1970, The endemic ducks of remote islands, Wildfowl 5:5-10.
Lande, R., 1981, Models of speciation by sexual selection on polygenic traits, Proc. Nat!.
Acad. Sci. USA 78:3721-3725.
LeCroy, M., 1981, The genus Paradisaea display and evolution, Am. Mus. Nov. 2714:
1-52.
Lewis, D. M., 1972, Importance of face-mask in sexual recognition and territorial behavior
in the yellowthroat, Jack-Pine Warbler 50:98-109.
Loehle, c., 1987, Hypothesis testing in ecology: Psychological aspects and the importance
of theory maturation, Q. Rev. BioI. 62:397-409.
Lorenz, K., 1962, The function of colour in coral reef fishes, Proc. R. Inst. Gr. Br. 39:
282-296.
Lorenz, K., 1966, On Aggression, Harcourt, Brace, and World, New York.
Lowther, J. K., 1961, Polymorphism in the White-throated Sparrow, Zonotrichia albicollis
(Gmelin), Can. J. Zool. 39:281-292.
Lyon, B. E., and Motgomerie, R. D., 1985, Conspicuous plumage of birds: Sexual selec.tion
or unprofitable prey? Anim. Behav. 33:1038-1040.
104 GREGORY S. BUTCHER and SIEVERT ROHWER

Lyon, B. E., and Motgomerie, RD., 1986, Delayed plumage maturation in passerine birds:
Reliable signaling by subordinate males?, Evolution 40:604-615.
Lythgoe, J. N., 1979, The Ecology of Vision, Oxford University Press, New York.
Manning, J. T., 1984, Males and the advantage of sex, J. Theor. BioI. 108:215-220.
Marler, P., 1955, Studies of fighting in chaffinches. 2. The effect on dominance relations
of disguising females as males, Br. J. Anim. Behav. 3:137-146.
Martin, S. G., 1970, The agonistic behavior of Varied Thrush (Ixoreus noevius) in winter
assemblages, Condor 72:452-459.
Maynard Smith, J., 1982, Evolution and the Theory of Games, Cambridge University
Press, Cambridge.
Maynard Smith, J., 1985, Sexual selection, handicaps, and true fitness, J. Theor. BioI.
115:1-8.
Mayr, E., 1942, Systematics and the Origin of Species, Columbia University Press, New
York.
Mayr, E., 1965, Animal Species and Evolution, Belknap Press-Harvard University Press,
Cambridge, Massachusetts.
Mayr, E., 1972, Sexual selection and natural selection, in: Sexual Selection and the
Descent of Man 1871-1971 (8. Campbell, ed.), Aldine, Chicago, pp. 87-104.
McKaye, K. R, Kocher, T., Reinthal, P., Harrison, R, and Kornfield, 1., 1984, Genetic
evidence for allopatric and sympatric differentiation among color morphs of a Lake
Malawi cichlid fish, Evolution 38:215-219.
Mock, D. W., 1981, White-dark polymorphism in herons, in: Proceedings of the First
Welder Wildlife Symposium, pp. 145-161.
Monroe, B. 1., Jr., 1964, Wing-flashing in the red-backed scrub-robin, Erythropygia zam-
besiana, Auk 81:91-92.
Montgomerie, R D., and B. E. Lyon, 1986, Does longevity influence the evolution of
delayed plumage maturation in passerine birds?, Am. Nat. 128:930-936.
Moodie, G. E. E., 1972, Predation, natural selection, and adaptation in an unusual Three-
spined Stickleback, Heredity 28:155-167.
Moreau, R E., 1960, Conspectus and classification of the Ploceine weaverbirds, Ibis 102:
298-321.
Moreau, R E., 1966, The Bird Faunas of Africa and Its Islands, Academic, London.
Morony, J. J., Bock, W. J., and Farrand, J., 1975, Reference List of Birds of the World,
Department of Ornithology, American Museum of Natural History, New York.
Morton, E. S., 1976, The adaptive significance of dull coloration in yellow warblers,
Condor 78:423.
Moynihan, M., 1960, Some adaptations which help to promote gregariousness, in: Pro-
ceedings of the XII International Ornithological Congress, June 5-12, 1958, G. Berg-
man, K. O. Donner, and 1. von Haartman, eds., Tilgmann Kirjapaino, Helsinki, Fin-
land, pp. 523-541.
Murton, R K. and Westwood, N. J., 1977, Avian Breeding Cycles, Clarendon Press,
Oxford.
Murton, R K., Westwood, N. J., and Thearle, R J. P., 1973, Polymorphism and the
evolution of a continuous breeding season in the pigeon, Columbia livia, J. Reprod.
Fertil. Suppl. 19:563-577.
Nur, N., and Hasson, 0., 1984, Phenotypic plasticity and the handicap principle, J. Theor.
BioI. 110:275-297.
Oberholser, H. C., 1974, Bird Life of Texas, University of Texas Press, Austin, Texas.
O'Donald, P., 1959, Possibility of assortative mating in the Arctic Skua, Nature (Lond.)
183:1210.
CONSPICUOUS COLORATION 105

O'Donald, P., 1980, Genetic Models of Sexual Selection, Cambridge University Press,
Cambridge, Massachusetts.
O'Donald, P., Davis, J. W. F., and Broad, R. A., 1974, Variation in assortative mating in
two colonies of Arctic Skuas, Nature (Land.) 252:700-701.
Odum, E. P., 1942, Annual cycle of the Black-capped Chickadee. 3, Auk 59:499-531.
Olson, S. 1., 1982, A critique of Cracraft's classification of birds, Auk 99:733-739.
Orians, G. H., and Christman, G. M., 1968, A comparative study of the behavior of Red-
winged, Tricolored, and Yellow-headed Blackbirds. Univ. Calif. Pub1. Zoo1. 84:1-81.
Parker, G. A., 1982, Phenotype-limited evolutionarily stable strategies, in: Current Prob-
lems in Sociobiology (King's College Sociobiology Group, eds.), Cambridge Univer-
sity Press, Cambridge, pp. 173-201.
Parker, G. A., 1983, Mate quality and mating decisions, in: Mate Choice (P. Bateson, ed.),
Cambridge University Press, Cambridge, pp. 141-166.
Parrish, J. W., Ptacek, J. A., and Will, K. 1., 1984, The detection of near-ultraviolet light
by nonmigratory and migratory birds, Auk 101:53-58.
Parsons, J., and Baptista, L. F. 1980, Crown color and dominance in the White-crowned
Sparrow, Auk 97:807-815.
Partridge, 1., 1980, Mate choice increases a component of offspring fitness in fruit flies,
Nature (Lond.) 283:290-291.
Paulson, D. R., 1973, Predator polymorphism and apostatic selection, Evolution 27:269-277.
Payne, R. B., 1967, Interspecific communication signals in parasitic birds, Am. Nat. 101:
363-375.
Payne, R. B., 1982, Ecological consequences of song matching: Breeding success and
intraspecific song mimicry in Indigo Buntings, Ecology 63:401-411.
Peek, F. W., 1972, An experimental study of the territorial function of vocal and visual
display in the male Red-winged Blackbird (Agelaius phoeniceus), Anim. Behav. 20:
112-118.
Peiponen, V. A., 1960, Verhaltensstudien am blaukehlchen, Ornis Fenn. 37:69-83.
Peterman, R. M., 1971, A possible function of coloration in coral reef fishes, Copeia 1971:
330-331.
Peterson, R. T., 1980, A Field Guide to the Birds, Houghton-Mifflin, Boston.
Pierotti, R., 1987, Isolating mechanisms in seabirds, Evolution 41:559-570.
Platt, J. R., 1964, Strong inference, Science 146:347-353.
Price, T. D., 1984, Sexual selection on body size, territory and plumage variables in a
population of Darwin's finches, Evolution 38:327-341.
Ratcliffe, 1. M., and Boag, P. T., 1987, Effects of colour bands on male competition and
sexual attractiveness in zebra finches (Poephila guttata), Can. ,. Zool. 65:333-336.
Read, A. F., 1987, Comparative evidence supports the Hamilton and Zuk hypothesis on
parasites and sexual selection, Nature (Lond.) 328:68-70.
Remington, C. 1., 1968, Suture-zones of hybrid interaction between recently joined biotas,
Evol. BioI. 2:321-428.
Remsen, J. V., Jr., 1984, High incidence of "leapfrog" pattern of geographic variation in
Andean birds: Implication for the speciation process, Science 224:171-173.
Ricklefs, R. E., and Cox, G. W., 1972, Taxon cycles in the West Indian avifauna, Am.
Nat. 106:195-219.
Ridley, M., 1983, The Explanation of Organic Diversity, Oxford University Press, Oxford.
Rockwell, R. F., Findlay, C. S., Cooke, F., and Smith, J. A., 1985, Life history studies of
the Lesser Snow Goose (Anser caerulescens caerulescens). IV. The selective value
of plumage polymorphism: Net viability, the timing of maturation, and breeding
propensity, Evolution 39:178-189.
106 GREGORY S. BUTCHER and SIEVERT ROHWER

Rohwer, S., 1975, The social significance of avian winter plumage variability, Evolution
29:593-610.
Rohwer, S., 1982, The evolution of reliable and unreliable badges of fighting ability, Am.
Zool. 22:531-546.
Rohwer, S., 1983, Formalizing the avoidance-image hypothesis: Critique of an earlier
prediction, Auk 100:971-974.
Rohwer, S., 1985, Dyed birds achieve higher social status than controls in Harris' spar-
rows, Anim. Behav. 34:1325-1331.
Rohwer, S., 1986~A previously unknown plumage of first-year Indigo Buntings and
theories of delayed plumage maturation, Auk 103:281-292.
Rohwer, S., and Butcher, G. S., 1988, Winter versus summer explanations of delayed
plumage maturation in temperate passerine birds, Am. Nat. 131:556-572.
Rohwer, S., and Ewald, P. W., 1981, The cost of dominance and advantage of subordi-
nation in a badge signaling system, Evolution 35:441-454.
Rohwer, S., and Niles, D. M., 1979, The subadult plumage of male Purple Martins:
Variability, female mimicry, and recent evolution, Z. Tierpsychol. 51:282-300.
Rohwer, S., and Paulson, D. R, 1987, The avoidance-image hypothesis and color poly-
morphism in Buteo hawks, Ornis Scand. 18:285-290.
Rohwer, S., Fretwell, S. D., and Niles, D. M., 1980, Delayed maturation in passerine
plumages and the deceptive aquisition of resources, Am. Nat. 115:400-437.
Rohwer, S., Ewald, P. W., and Rohwer, F. C., 1981, Variation in size, appearance, and
dominance within and among the sex and age classes of Harris' Sparrows, J. Field
Ornithol. 52:291-303.
Rohwer, S., Klein, W. P. Jr., and Heard, S., 1983, Delayed plumage maturation and the
presumed prealternate molt in American Redstarts, Wilson Bull. 95:199-208.
Roskaft, E., and Jarvi, T., 1983, Male plumage colour and mate choice of female Pied
Flycatchers Ficedula hypoleuca, Ibis 125:396-400.
Roskaft, E., and Rohwer, S., 1987, An experimental study of the function of the epaulettes
and the black body colour of male red-winged blackbirds, Anim. Behav. 35:1070-1077.
Roskaft, E., Jarvi, T., Nyholm, N. E. I., Virolainen, M., Winkel, W., and Zang, H., 1986,
Geographic variation in secondary sexual plumage colour characteristics of the male
Pied Flycatcher, Ornis Scand. 17:293-298.
Rowland, W. J., 1982, The effects of male nuptial coloration on stickleback aggression:
A reexamination, Behaviour 80:118-126.
Samson, F. B., 1977, Social dominance in winter flocks of Cassin's Finch, Wilson Bull.
89:57-66.
Schutz, F., 1965, Sexuelle pragung bei Anatiden, Z. Tierpsychol. 22:50-103.
Selander, R K., 1965, On mating systems and sexual selection, Am. Nat. 99:129-141.
Selander, R. K., 1972, Sexual selection and dimorphism in birds, in: Sexual Selection
and the Descent of Man 1871-1971 (B. Campbell, ed.), Aldine, Chicago, pp. 180-230.
Selander, R. K., and Hunter, D. K., 1960, On the function of wing-flashing in mockingbirds,
Wilson Bull. 7:341-345.
Semler, D. E., 1971, Some aspects of adaptation in a polymorphism for breeding colours
in the Threespine Stickleback (Gasterosteus aculeatus), J. Zool. 165:291-302.
Shield, W. M., 1977, The social significance of avian winter plumage variability: A
comment, Evolution 31:905-906.
Sibley, C. G., 1957, The evolutionary and taxonomic significance of sexual dimorphism
and hybridization in birds, Condor 59:166-191.
Sibley, C. G., and Ahlquist, J. E., 1983, Phylogeny and classificaion of birds based on the
data of DNA-DNA hybridization, Current Ornithol. 1:245-292.
CONSPICUOUS COLORATION 107

Skutch, A. F., 1940, Some aspects of Central American bird-life, Sci. Mthly. 51:409-418,
500-511.
Skutch, A. F., 1957, The resident wood warblers of Central America, in: The Warblers
of America (1. Griscom and A. Sprunt, eds.), Devin-Adair, New York. pp. 275-285.
Smith, D. G., 1972, The role of the epaulets in the Red-winged Blackbird (Agelaius
phoeniceus) social system, Behaviour 41:251-268.
Smith, N. G., 1966, Evolution of some arctic gulls (Larus): An experimental study of
isolating mechanisms, Ornith. Monogr. (AOU) 4:1-99.
Smith, S. M., 1980, Henpecked males: The general pattern in monogamy?, J. Field Or-
nithol. 51:55-64.
Stearns, S. c., 1976, Life-history tactics: A review of the ideas, Q. Rev. BioI. 51:3-47.
Stephenson, A. G., and Bertin, R. I., 1983, Male competition, female choice, and sexual
selection in plants, in: Pollination Biology (1. Real, ed.), Academic. New York, pp.
109-149.
Studd, M. V., and Robertson, R. J., 1985, Life span, competition, and delayed plumage
maturation in male passerines: The breeding threshold hypothesis, Am. Nat. 126:
101-115.
Thompson, W. 1., 1960, Agonistic behavior in the House Finch. II. Factors in aggres-
siveness and sociality, Condor 62:378-402.
Thorneycroft, H. G., 1975, A cytogenetic study of the White-throated Sparrow, Zonotri-
chia albicollis, Evolution 29:611-621.
Thornhill, R, 1980, Mate choice in Hylobittacus apical is (Insecta: Mecoptera) and its
relation to some models of female choice, Evolution 34:519-538.
Tordoff, H. B., 1954, Social organization and behavior in a flock of captive, nonbreeding
Red Crossbills, Condor 56:346-358.
Trivers, R. 1., 1972, Parental investment and sexual selection, in: Sexual Selection and
the Descent of Man 1871-1971 (B. Campbell, ed.)., Aldine, Chicago, pp. 136-179.
Verner, J., and Willson, M. F., 1969, Mating systems, sexual dimorphism, and the role
of North American passerine birds in the nesting cycle, Omithol. Monogr. (AOU) 9:
1-76.
Wallace, A. R., 1889, Darwinism: An Exposition of the Theory of Natural Selection with
Some of Its Applications, Macmillan, London.
Walsberg, G. E., 1982, Coat color, solar heat gain, and conspicuousness in the Phaino-
pepla, Auk 99:495-502.
Walsberg, G. E., 1983, Ecological energetics: What are the questions?, in: Perspectives in
Ornithology (A. H. Brush and G. A. Clark, eds.), Cambridge University Press, Cam-
bridge, pp. 135-158.
Walsberg, G. E., Campbell, G. S., and King, J. R., 1978, Animal coat color and radiative
heat gain: Are-evaluation, J. Compo Phys. 126:211-222.
Watson, G. E., 1962, Sympatry in palearctic Alectoris partridges, Evolution 16:11-19.
Watt, D. J., 1986a, A comparative study of status signalling in sparrows (genus Zonotri-
chia), Anim. Behav. 34:1-15.
Watt, D. J., 1986b, Relationship of plumage variability, size and sex to social dominance
in Harris' sparrows, Anim. Behav. 34:16-27.
Watt, D. J., Ralph, C. J., Atkinson, C. T., 1984, The role of plumage polymorphism in
dominance relationships of the White-throated Sparrow, Auk 101:110-120.
Weldon, P. J., and Burghardt, G. M., 1984, Deception divergence and sexual selection,
Z. Tierpsychol. 65:89-102.
West-Eberhard, M. J., 1983, Sexual selection, social competition, and speciation, Q. Rev.
BioI. 58:155-183.
108 GREGORY S. BUTCHER and SIEVERT ROHWER

Whitfield, D. P., 1986, Plumage variability and territoriality in breeding turnstone Ar-
enaria interpres: Status signalling or individual recognition?, Anim. Behav. 34:
1471-1482.
Whitfield, D. P., 1987, Plumage variability, status signalling and individual recognition
in avian flocks, Trends Ecol. Evol. 2:13-18.
Williams, E. E., and Rand, A. S., 1977, Species recognition, dewlap function, and faunal
size, Am. Zool. 17:265-274.
Williams, G. C., 1966, Adaptation and Natural Selection. A Critique of Some Current
Evolutionary Thought, Princeton University Press, Princeton, New Jersey.
Willis, E. 0., 1972, The behavior of spotted antbirds, Ornithol. Monogr. (AOU) 10:1-162.
Witherby, H. F., Jourdain, F. C. R., Ticehurst, N. F., and Tucker, B. W., 1943, The Hand-
book of British Birds, Vol. I, Crows to Firecrest, Witherby, London.
Wolf, 1. 1., 1969, Female territoriality in a tropical hummingbird, Auk 86:490-504.
Wolf, 1. 1., 1975, Female territoriality in the Purple-throated Carib, Auk 92:511-522.
Wunderle, J. M., Jr., 1981, An analysis of a morph ratio cline in the Bananaquit (Coereba
flaveola) on Grenada, West Indies, Evolution 35:333-344.
Wunderle, J. M., Jr., 1983, A shift in the morph ratio cline in the Bananaquit of Grenada,
West Indies, Condor 85:365-367.
Zahavi, A., 975, Mate selection-A selection for a handicap, ,. Theor. BioI. 53:205-214.
Zahavi, A., 1977, The cost of honesty, J. Theor. BioI. 67:603-605.
Zahavi, A., 1981, Natural selection, sexual selection, and the selection of signals, in:
Evolution Today (G. G. E. Scudder and J. L. Reveal, eds.), Proceedings of the Second
International Congress on Systematic and Evolutionary Biology, July 17-24, 1980,
Hunt Institute for Botanical Documentation, Carnegie-Mellon University, Pittsburg,
Pennsylvania, pp. 133-138.
Zumpe, D., 1965, Laboratory observations on the aggressive behaviour of some butterfly
fishes, Z. Tierpsychol. 22:226-236.

You might also like