From Pheromones to Behavior - Physiological Reviews

Transcription

From Pheromones to Behavior - Physiological Reviews
Physiol Rev 89: 921–956, 2009;
doi:10.1152/physrev.00037.2008.
From Pheromones to Behavior
ROBERTO TIRINDELLI, MICHELE DIBATTISTA, SIMONE PIFFERI, AND ANNA MENINI
Department of Neuroscience, University of Parma, Parma; and International School for Advanced Studies,
Scuola Internazionale Superiore di Studi Avanzati (SISSA) and Italian Institute of Technology, SISSA Unit,
Trieste, Italy
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
I. Introduction
II. Mammalian Pheromones
A. Pheromones in various animal orders
B. Major urinary proteins and major histocompatibility complex molecules
C. Exocrine gland-secreting peptides
III. Chemosensory Systems Detecting Pheromones
A. Vomeronasal epithelium
B. Olfactory epithelium
C. Other olfactory organs
IV. Putative Pheromone Receptors
A. Vomeronasal receptors
B. Odorant receptors
C. Trace amine-associated receptors
D. Olfactory-specific guanylyl cyclase type D receptor
V. Pheromone Transduction
A. Vomeronasal sensory neurons
B. Olfactory sensory neurons
C. Grueneberg ganglion
VI. Signaling From Sensory Epithelia to the Brain
A. Main and accessory olfactory bulbs
B. Signaling beyond the olfactory bulbs
C. Pheromone-Activated hormonal systems
VII. Pheromone-Mediated Behaviors
A. Male behaviors
B. Female behaviors
VIII. Conclusions and Future Challenges
921
922
922
925
926
927
927
929
929
930
930
933
934
934
935
935
939
940
940
940
943
944
944
945
946
947
Tirindelli R, Dibattista M, Pifferi S, Menini A. From Pheromones to Behavior. Physiol Rev 89: 921–956, 2009;
doi:10.1152/physrev.00037.2008.—In recent years, considerable progress has been achieved in the comprehension of
the profound effects of pheromones on reproductive physiology and behavior. Pheromones have been classified as
molecules released by individuals and responsible for the elicitation of specific behavioral expressions in members
of the same species. These signaling molecules, often chemically unrelated, are contained in body fluids like urine,
sweat, specialized exocrine glands, and mucous secretions of genitals. The standard view of pheromone sensing was
based on the assumption that most mammals have two separated olfactory systems with different functional roles:
the main olfactory system for recognizing conventional odorant molecules and the vomeronasal system specifically
dedicated to the detection of pheromones. However, recent studies have reexamined this traditional interpretation
showing that both the main olfactory and the vomeronasal systems are actively involved in pheromonal communication. The current knowledge on the behavioral, physiological, and molecular aspects of pheromone detection in
mammals is discussed in this review.
I. INTRODUCTION
Chemosensation is one of the sensory modalities that
animals have evolved to analyze the chemical properties
www.prv.org
of the external world, enabling the detection and the
discrimination of a very high number of molecules with
different structures. Substances carrying a chemical message among animals are termed “semiochemicals,” from
0031-9333/09 $18.00 Copyright © 2009 the American Physiological Society
921
922
TIRINDELLI ET AL.
determinants. For example, trace pheromones allow attracting a conspecific, while alarm pheromones alert
other conspecifics about an external challenge and can
evoke aggressive behaviors.
A first classification groups intraspecific signaling
molecules in releaser and primer pheromones (239). Releaser pheromones lead to individual recognition and
evoke short-latency behavioral responses in the conspecific like in aggressive attacks, or mating, or territory
marking. Conversely, primer pheromones induce delayed
responses that are commonly mediated through the activation of the neuroendocrine system. Thus, through the
emission of chemical cues, an individual forces the hormonal balance of a receiver, thus modulating its reproductive status.
In mammals, pheromones have been reported to be
involved in a variety of behavioral effects also in unrelated
species and, surprisingly, some pheromonal molecules are
shared, with different purposes, between species.
II. MAMMALIAN PHEROMONES
2. Marsupialia
A. Pheromones in Various Animal Orders
Pheromones have a very large chemical variability
(Table 1), and therefore they are preferably grouped according to their functions rather than to their structural
Physiol Rev • VOL
1. Proboscidea
Despite their size, which would obviously preclude
extensive behavioral studies, Asian and African elephants
have been largely investigated for their pheromonal responses by the late Bets Rasmussen. During musth, a
periodic condition in which mature males are characterized by highly aggressive behavior, elephants produce
large quantities of a specific pheromone, frontalin (1,5dimethyl-6,8-dioxabicyclo[3.2.1]octane), that is released
in temporal gland secretions, urine, and breath (307).
Adult males are mostly indifferent to frontalin, whereas
subadult males are highly reactive, often exhibiting repulsion or avoidance. Female chemosensory responses to
frontalin vary with the hormonal status. Females in the
follicular phase are the most responsive and often demonstrate mating-related behaviors subsequent to frontalin
stimulation (304). Furthermore, female Asian elephants
excrete a urinary pheromone, (Z)-7-dodecen-L-yl acetate,
to signal to males their readiness to mate (305). Interestingly, females of more than 100 species of insects use the
same compound as part of their pheromone blends to
attract insect males. The finding that the same molecule
may be shared by unrelated species for a different pheromonal purpose is only apparently surprising, owing to the
fact that structural constraints are often dictated by common metabolic pathways.
In marsupials, the gray short-tailed opossum (Monodelphis domestica) communicates by scent marking. The
male opossum possesses a prominent suprasternal scent
gland, whose extracts strongly attract female opossums
(131), stimulating the estrus in anestrous females via the
vomeronasal system.
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
the Greek “semeion” (sign). An important class of semiochemicals is constituted by pheromones. The term pheromone, based on the Greek “pherein” (to carry or transfer) and “hormon” (to stimulate or excite), was introduced in 1959 by the entomologists Peter Karlson and
Martin Lüscher (151) to identify specific biologically active substances “which are secreted to the outside by an
individual and received by a second individual of the same
species, in which they release a specific reaction, for
example, a definite behavior or a developmental process.”
In the same year, the first insect pheromone bombykol,
released by the female silkworm to attract mates, was
characterized by Butenandt et al. (43). The use of pheromones among insects has been extensively investigated,
and interested readers may consult several reviews on the
subject (4, 111, 130, 379).
Another class of semiochemicals is constituted by allelomones: substances produced by members of one species
that influence the behavior of members of another species
(for a review, see Ref. 333). Although in this review we will
mainly discuss the effects elicited by pheromones in mammals, we will also briefly mention some examples of allelomonal communication (5, 107, 108, 169).
In most mammals, the nasal cavity contains at least two
major sensory systems that may be involved in the detection
of pheromones: the vomeronasal and the main olfactory
systems. Evidence indicates that some behaviors are mediated exclusively by one or by a complementary action of
both systems. For example, Bruce (37) reported that, when
female mice that have been recently inseminated are exposed to urine of a male different from the one they mated
with, the pregnancy is interrupted and the female returns to
estrus. This effect depends on a functional vomeronasal
system (15). Differently, the nipple search in rabbit pups is
one of the typical examples of behavior strictly dependent
on an intact main olfactory system (119). In other instances,
behavioral responses are mediated by both the vomeronasal
and the main olfactory systems. This is the case of female
hamsters, whose ultrasonic calling during estrus is abolished by the inactivation of either the vomeronasal or the
main olfactory system (145). In recent years, the discovery of
distinct pheromone receptor families coupled to a complex
network of signal transduction pathways, together with advances in molecular genetics and functional electrophysiological and imaging techniques, have greatly advanced our understanding of the physiological role of pheromones.
923
FROM PHEROMONES TO BEHAVIOR
TABLE
1.
Table of pheromones
Pheromone
Origin
Organ
Mouse female urine
VNO
2-Sec-butyl-4,5dihydrothiazole (BT)
Mouse male urine
VNO
2,3-Dehydro-exobrevicomin (DB)
Mouse male urine
VNO
␣- and ␤-Farnesene
Mouse preputial glands
VNO
6-Hydroxyl-6-methyl-3heptanone
Mouse male urine
VNO
2-Heptanone
Mouse urine
VNO
n-Pentyl acetate
Mouse urine
VNO
Isobutylamine
Mouse male urine
VNO
3-Amino-s-triazole (similar
to BT)
4-Ethyl phenol
3-Ethyl-2,7-dimethyl octane
(similar to DB)
Mouse male urine
?
Mouse male urine
Mouse male urine
?
?
3-Cyclohexene-1-methanol
Major urinary proteins
(MUPs)
MHC class I peptides
Mouse male urine
Mouse urine
?
VNO
Mouse urine
VNO
Behavior
V1R neurons
IP3 production in the VNO (377) Suppression of estrus cycle (Lee Booteffect) (219, 281)
Electrical response and calcium
influx in VNO neurons (187)
V1R neurons
Electrical response and calcium Estrus induction and synchronization
influx in VNO neurons (187)
(Whitten effect) (138)
Protooncogene expression in
Puberty acceleration (Vandenbergh effect)
the AOB (32, 95)
(266, 283)
Intermale aggressiveness (280)
Attractiveness to females (139)
V1R neurons
Electrical response and calcium Estrus induction and synchronization
influx in VNO neurons (187)
(Whitten effect) (138)
Protooncogene expression in
Puberty acceleration (Vandenbergh effect)
the AOB (32, 95)
(266, 283)
Intermale aggressiveness (280)
Attractiveness to females (139)
V1R neurons
Electrical response and calcium Puberty acceleration (Vandenbergh effect)
influx in VNO neurons (187)
(266, 283)
Territorial marking (279)
Attractiveness to females (139)
Suppression of estrus cycle (Lee Booteffect) (219)
V1ra and V1rb (V1Rs) Electrical response and calcium Puberty acceleration (Vandenbergh effect)
(266, 282)
neurons
influx in VNO neurons (61,
187)
V1R neurons
Electrical response and calcium Estrus extension (137)
influx in VNO neurons (187)
IP3 production in the VNO (377)
V1ra and V1rb (V1Rs) Electrical response in VNO
Suppression of estrus cycle (Lee Bootneurons
neurons (61)
effect) (281)
V1ra and V1rb (V1Rs) Electrical response in VNO
Estrus acceleration and vaginal opening
neurons
neurons (61)
(276, 301)
?
?
Attractiveness to females (2)
?
?
?
V2R neurons
V2R neurons
(SYFPEITHE;
AAPDNRETF)
Exocrine gland secreting
peptides
␣2-Urinary proteins (␣2u)
MOE
Extraorbital lacrimal glands
VNO
V2R neurons
Rat urine
VNO
V2R neurons
Dodecyl-propionate
Aphrodisin
Rat preputial gland
Hamster vaginal glands
?
VNO
?
?
?
?
?
?
?
?
2-Methylbut-2-enal
Rabbit milk
(Z)-7-Dodecen-1-yl acetate Elephant female urine
Male asian elephant temporal
(⫹)(⫺)1,5-Dimethyl-6,8gland, urine, and breath
dioxabicyclo[3.2.1]octane
(frontaline)
Cellular/Biochemical Response
?
?
?
Electrical response and calcium
influx in VNO neurons (47,
163)
In vitro survival and
proliferation of VNO neurons
(408)
Electrical response and calcium
influx in MOE and VNO
neurons (364)
Single-cell recording in VNO
neurons (156, 185)
Electrical response in VNO
neurons (162, 163)
G protein activation and IP3
production in the VNO (176,
321)
?
IP3 production in the VNO and
c-fos expression in AOB
neurons (136, 177)
?
?
?
Attractiveness to females (2)
Attractiveness to females (2)
Repulsion to mouse (territorial marking?)
(2)
Attractiveness to females (2)
Puberty acceleration (Vandenbergh effect)
(266, 283)
Individual recognition (49, 123, 375)
Intermale aggressiveness
Attractiveness to individuals of a different
mouse strain (358)
Pregnancy block (Bruce effect)
?
?
Ano-genital licking (36)
Copulatory behavior in male (356)
Attraction and oral grasping in pups (335)
Mating-associated behaviors in males (306)
Repulsion, attractiveness, and sexual
behavior depending on the racemic ratio
(92, 304)
VNO, vomeronasal organ; MOE, main olfactory epithelium; AOB, accessory olfactory bulb. Reference numbers are given in parentheses.
3. Soricomorpha (Insectivores)
In musk shrew (Suncus murinus), virgin females exposed to cages recently vacated by an adult male receive
Physiol Rev • VOL
mounts from males significantly sooner than females
exposed to cages soiled by a castrated male, by another
female, or to a clean cage (310).
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
2,5-Dimethyl pyrazine
Target
924
TIRINDELLI ET AL.
In tree shrew (Tupaia belangeri), both males and
females mark their surroundings with urine and skin
gland secretions. Males mark more frequently than females in the absence of conspecific scent; moreover, male
scent stimulates marking in females. Interestingly, increasing amounts of male scent result in a corresponding
increase in marking by females (114), suggesting that, in
the wild, pheromones may not exert an all-or-none effect
but rather act in a dose-specific fashion.
4. Carnivora and Ungulata
5. Primata
Primates scarcely rely on chemical communication;
however, the ring-tailed lemur (Lemur catta) exhibits the
most highly developed olfactory system among all species
belonging to this order (75, 83, 338). Lemur catta possesses a suite of specialized scent glands and a variety of
scent-marking displays (147, 258, 338). Both sexes have
apocrine, located on the wrists, and sebaceous gland
fields in their genital regions and adopt distinctive handstand postures to deposit glandular secretions on substrates. Males mix the glandular secretions and then deposit this mixture via “wrist marking.” They also impregnate their tail fur with their own secretions and then waft
their tail at opponents during characteristic “stink fights”
(147).
There is nowadays a general consensus for the existence of a chemical communication in humans, although
there is still a great controversy about the existence,
extent, and nature of human pheromones. Among the best
and first known effects attributed to human pheromones
is the synchronization of menstrual cycle (237, 238, 367)
and the dimorphic activation of brain areas (332). ExpoPhysiol Rev • VOL
6. Rodentia
The paradigmatic model for mammalian pheromone
investigations involves laboratory species, among which
rodents and especially mice are the most employed, because they are easy to handle, have extremely well-developed olfactory systems, and offer an almost unlimited
possibility to act on behavioral responses by genetic manipulation. In these species, several effects have been
carefully described, and the mechanisms of their action
have even been dissected at a molecular level.
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
Both carnivores (dogs and cats) and ungulates commonly mark their territory with urine and display a specific behavior called “flehmen” (67, 88, 391). In the flehmen reaction, animals, after the physical contact with a
conspecific or with scents emanating from its excrements
(urine or feces), lift the head, draw back their lips, and
push the tongue towards the anterior region of the palate
in a manner that makes them appear to be “grimacing.” It
is commonly believed that this attitude would allow a
faster transfer of pheromones into the olfactory organs.
The saliva of the adult male pig contains a mixture of
steroids (boar odor). When a boar becomes aggressive or
sexually aroused, it produces copious amount of saliva
whose pungent odor is attractive to estrus females (287)
and facilitates the display of the estrous female’s mating
posture (352). The major component of the boar odor is
androstenone, a sexually dimorphic steroid that elicits
both behavioral responses (65). Curiously, androstenone
has been hypothesized to be a human pheromone as well
(18, 96).
sure to odorless axillary’s secretions from females in the
late follicular and ovulatory phase of the menstrual cycle,
respectively, shortens and lengthens the menstrual cycles
of recipient females, accelerating or delaying the luteinizing hormone (LH) surge (367). Women smelling an androgen-like compound activate the preoptic and ventromedial nuclei of the hypothalamus in contrast to men who
activate the paraventricular and dorsomedial nuclei hypothalamus when smelling an estrogen-like substance (332).
Although a robust hypothalamic response is commonly
observed with ordinary odorants, such a dramatic dimorphic difference can be interpreted as a hormonal response
to pheromone stimulation.
The apocrine gland secretion of the axilla in combination with the microbial flora is probably the most relevant source of candidate compounds for human pheromones, which consists of a blend of molecules characterizing the individual body odor (14). However, salivary and
seminal secretions and urine represent other reservoirs of
putative human pheromones. The male sweat and salivary
progesterone derivative androstadienone (4,16-androstadien-3-one) (AND) has been indicated as a putative human pheromone influencing brain activity (20, 331, 332),
endocrine levels (406), physiological arousal, context-dependent mood, and sexual orientation (132, 133, 212, 232,
300). Woman urine contains the estrogen like steroid
estra-1,3,5(10),16-tetraen-3-ol (EST) that also appears to
be a good candidate compound for human pheromones
(20, 132, 331, 332).
In studies of brain activity by positron emission tomography, as above reported, smelling AND and EST
activates regions primarily incorporating the sexually dimorphic nuclei of the anterior hypothalamus (332), and
this activation is differentiated with respect to sex and
compound. Very interestingly, homosexual males process
AND similarly to heterosexual women rather than to heterosexual men (20). Conversely, lesbian women process
AND stimuli as common odorants and, when smelling
EST, they partly share activation of the anterior hypothalamus with heterosexual men (20). These observations
indicate the involvement of the anterior hypothalamus in
physiological processes related to sexual orientation/preference in humans.
FROM PHEROMONES TO BEHAVIOR
Physiol Rev • VOL
signal alert to the presence of a potential danger, promoting dispersion and the adoption of defensive actions in the
group (400). Usually, mice escape from places in which
the odors from the urine of a physically stressed conspecific are present (418). Among other ketones, 2-heptanone
is detectable in the mouse and rat urine (97, 342). In rat,
2-heptanone was found to be increased in urine of
stressed subjects and to induce immobility in a recipient
conspecific forced to swim (97).
Several other urinary compounds have been indicated as activators of distinct anatomical regions of the
pheromonal pathway in rats (176, 321, 372); however,
none of the effects exerted by these substances is reportedly linked to obvious behavioral responses.
Mouse ecology is largely triggered by pheromonal
signals, which have been investigated for a long time, and
therefore, mice represent the world-wide animal model to
study intraspecies communication. In the last decade, an
enormous burden of data has been collected and, how it
is often the case, there is yet an open and sometimes
contrasting debate on the specific properties of the different signaling molecules.
Early studies have demonstrated a number of incontrovertible effects of primer pheromones in mice. It is
known that grouped female mice modify or suppress their
estrous cycle (Lee-Boot effect) (387), while male urine
can restore and synchronize the estrus cycle of noncycling females (Whitten effect) (401) or accelerate puberty
onset in females (Vandenbergh effect) (388). In addition,
the exposure of a recently mated female mouse to a male,
different from the stud, prevents implantation of fertilized
eggs (Bruce effect) (38), implying that the stud or its
individual odor must be memorized at the moment of
mating to be recognized later. These pheromonal effects
are commonly believed to be mediated by stimuli present
in urine that act via the vomeronasal system.
Signaling pheromones also account for many behavioral responses in the mouse. For example, male as well
as lactating female aggressiveness towards an intruder
male are both believed to be triggered by molecules that
are present in male urine.
B. Major Urinary Proteins and Major
Histocompatibility Complex Molecules
Male mouse urine contains an unusually high quantity of proteins, called major urinary proteins (MUPs).
MUPs form a large family of highly homologous androgendependent proteins that are synthesized in the liver and
excreted with urine (46, 208). MUPs and the rat hortologs
␣2 urinary globulins belong to the lipocalin superfamily,
and accordingly, they bind and release small (volatile)
pheromones that are also produced in an androgen-dependent fashion (9, 23). Although MUPs are also ex-
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
The vaginal secretion and saliva of female golden
hamsters (Mesocricetus auratus) contain an aphrodisiac
substance that facilitates the mating behavior of the male
(223, 356). This pheromonal effect has been clearly demonstrated in assays that measure the mounting behavior
of a male in the presence of an anesthetized male hamster
(defined as surrogate female) scented with the stimuli
(222). Evidence shows that biologically active compounds
were present in the proteinaceous fraction, whereas aqueous fractions, organic extracts, and distillates containing
volatile compounds did not exert detectable effects. After
purification, pheromonal activity was attributed to a 17kDa soluble glycosylated protein appropriately termed
aphrodisin (356). The complete primary structure of aphrodisin shows that this protein belongs to the lipocalin
superfamily. A prominent feature of almost all lipocalins
is the presence of a hydrophobic pocket that accommodates endogenous molecules with a relatively high lipophilic coefficient. The biochemically purified aphrodisin carries natural molecules that are contained in the
vaginal secretions (33, 34) and that may be responsible for
the pheromonal effect observed in males (136). Biochemical and immunohistochemical data suggest that aphrodisin may act through G protein-coupled receptors that are
expressed in the vomeronasal system (136, 177).
Fossorial mammals, like the mole-rat, are noted for
sensory specializations, such as exquisite tactile senses
(45, 60). Although suppression of mole-rat reproduction
in females is thought to be meditated by behavioral,
rather than pheromonal, cues from the breeding females
(360), the chemosensory systems retain demonstrable
functions of scent marking in common nesting and toilet
loci (359) that have been associated with pheromonal
communication.
In rats, several pheromones are used in different
contexts as territory marking (309), alarm (1), and maternal behaviors (188, 254). The chemically characterized rat
pheromones include compounds that are produced by
pup’s preputial glands and urine. A particular pattern of
rat maternal behavior is the specific licking and cleaning
of pup’s anogenital area, an action crucial to pup survival,
as nonlicked pups cannot defecate and run into death.
Thus dodecyl propionate, a compound isolated from
pups’ preputial glands, can sustain pups’ anogenital licking by dams and therefore pups’ survival (36).
The odors emanating from physically stressed rodents are recognized by conspecifics (220, 386). Once the
chemical signal is perceived, the receiver undergoes significant changes. For example, lymphocyte T-killers are
suppressed, whereas B-cells proliferate (57), rats develop
hyperthermia (160), c-Fos protein expression is increased
in the mitral cell layer of the accessory olfactory bulb
(164), reproductive behavior and sexual maturation initiate (389a), and social hierarchy may be modified (140).
These bodily compounds, defined as alarm pheromones,
925
926
TIRINDELLI ET AL.
Physiol Rev • VOL
Thom et al. (375) indicated that female mice use
MUPs as a specific genetic marker to identify and preferentially associate with heterozygous males. The primary
function of MUPs is in signaling social information
through scent, including individual genetic identity. Mice
are highly sensitive to differences in the fixed patterns of
multiple MUP isoforms expressed by each individual and
also avoid inbreeding with close relatives that share the
same MUP type as themselves. Heterozygosis assessment,
in contrast, most likely involves recognition of the greater
diversity of MUP isoforms expressed by heterozygous
animals.
Other biological effects of MUPs include attractiveness to females and repulsion to males (265), modification
of light-avoidance behavior in both males and females
(263, 264), the aggressive reactions of males towards
adult and newborn mice (264, 267), and intermale aggressiveness (47). Mate choice and parent-offspring interactions have been demonstrated to be influenced by MHC
genotype B (411). For example, house mice avoid mating
with individuals that are genetically similar at the MHC
locus. Mice are able to recognize MHC-similar individuals
through specific odorant cues. Moreover, females avoid
mating with males carrying MHC genes of their foster
family, supporting a familial imprinting hypothesis (288).
Mate preference for MHC-dissimilar individuals can be
adaptive as it would increase offspring MHC heterozygosity, with beneficial influences on offspring viability through
increased resistance to infectious disease or avoidance of
inbreeding effects (180, 297).
The polymorphism of MHC molecules is reflected
into structurally diverse binding sites, such that different
MHC binds to distinct peptides. Like other membranebound proteins, these MHC molecules are anchored in the
lipid bilayer. However, it has been demonstrated that the
peptide-binding pocket releases its peptide in the extracellular fluid when the MHC is cleaved from the cell. Thus
peptides or their fragments are constitutively excreted in
urine and other bodily secretions (358) that in turn can be
used for interindividual communication (357).
C. Exocrine Gland-Secreting Peptides
Mice secrete a family of peptides of different length
called exocrine gland-secreting peptides (ESPs) from extraorbital, Harderian and submaxillary glands into tear,
nasal mucus, or saliva. The ESP family consists of 38
members in mice and 10 in rats, while it is absent from the
human genome.
Moreover, at least two of these peptides, produced by
the extraorbital lacrimar gland, are gender specific: ESP1
is exclusively secreted by males (162), while ESP36 is
female specific (163). Despite the fact that these peptides
are recognized by systems dedicated to pheromone per-
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
pressed in exocrine glands as mammary, parotid, sublingual, submaxillary, lachrymal, nasal, and in modified sebaceous glands like preputial and perianal glands (344),
their biological effects have been exclusively demonstrated when these proteins were purified from urine or
added to urine that does not possess pheromonal activity
(i.e., of a prepubertal or castrated mouse) (123, 266).
MUPs have been reported to act either as primer or
releaser pheromones; however, some of their effects appear contradictory. This is in part due to the tight binding
of MUPs with volatile pheromones that have been proven
to be directly implicated in estrus synchronization (138),
attractiveness to females (139), intermale aggression
(280), induction of estrus (␣- and ␤-farnesene) (219), puberty acceleration in females (283), and territory marking
(21, 122). The advantage of such a pheromone-protein
complex may include concentration and slow release of
chemosignals, stabilization of labile pheromones, or delivery to the reception sites. Common chromatographic
procedures for purification of MUPs are ineffectual to
remove these pheromones, and solvent extraction of volatiles appears the only way to obtain the native apoproteins. Therefore, in literature, the term of MUPs commonly defines the protein with the bound pheromones.
While a general consensus exists about the source,
male urine, that releases the pheromonal stimuli responsible for the puberty acceleration (Vandenbergh effect), a
debate is currently open about the molecules that determine this behavioral response. Evidence suggests that
MUPs and not the bound pheromones are effective in
accelerating the puberty onset in females. These results
are supported by the biological activity shown by MUPs
when stripped of its pheromones and by a short synthetic
hexapeptide homologous to the NH2-terminal sequence of
many MUPs variants (266). Conclusions from a different
laboratory, performing similar but not identical experiments, seem, however, to argue against this hypothesis
(283).
It has been also demonstrated that the combinatorial
diversity of urinary MUPs among wild mice (a male can
produce up to 12 variants from the polymorphic MUP
family, that includes a total of 21 functional isoforms) is
virtually as great as for molecules of the major histocompatibility complex (MHC) (see below) (48, 123). However,
although MHC has been long regarded as the primary
determinant for individual recognition via bodily odors,
this evidence has now been challenged by recent studies
in which it appears that recognition, at least in the wild, is
instead strictly and exclusively dependent on the MUPs
variant profile contained in each individual urine (49). Very
intriguing is the observation that recognition requires the physical contact with the donor urine, thus discarding the
hypothesis of this effect being exclusively attributed to
the volatile pheromones released from MUPs (49).
FROM PHEROMONES TO BEHAVIOR
ception, until now there were no clues regarding their
physiological function.
III. CHEMOSENSORY SYSTEMS
DETECTING PHEROMONES
A. Vomeronasal Epithelium
The vomeronasal epithelium is part of the vomeronasal organ (VNO) of Jacobson, a tubular structure en-
cased in a protective bony capsule located at the base of
the nasal septum (68, 134) (Fig. 2). Given its recessed
nature, the vomeronasal epithelium cannot be reached by
the airstream that regularly flows through the nasal cavity. Thus, to target the vomeronasal epithelium, molecules
dissolved in the nasal mucus must be sucked into the
lumen of the organ. Lateral to the lumen are blood vessels
and sinuses, innervated by the autonomic nervous system, that induces vasodilation and vasoconstriction,
thereby producing a pumplike action for stimulus access
to the lumen (72, 109, 244, 248, 249, 328). In ungulates,
as described in section IIA4, the “flehmen” behavior is
thought to be associated with promoting stimulus access
to the vomeronasal organ (268).
Norepinephrine accumulation has been measured in
the vomeronasal organ of mice after exposure to pheromonal cues (419). It is possible that this hormonal release
alters the vascular tone, or changes the glandular secretions, thus increasing the receptivity of the sensory epithelium in particular behavioral contexts. The functioning
of the vomeronasal organ is thus not simply a passive
event, but it can be actively modulated.
The medial, concave side of the vomeronasal lumen
is lined by a pseudostratified epithelium composed of
three types of cells: sensory neurons, supporting cells,
and basal cells. The basal stem cells are located both
along the basal membrane of the sensory epithelium and
at the boundary with the nonsensory epithelium (85).
Supporting cells are found in the uppermost superficial
layer of the sensory epithelium, while vomeronasal neurons form two overlapping populations, basal and apical,
that are further described in detail in sections IV and V.
In the mouse, the vomeronasal sensory epithelium
increases from birth to puberty and becomes fully developed and completely active at 2 mo after birth (117). The
sensory neurons, on the mucosal side, bear dendrites and
apical microvilli, whereas, on the opposite side, their
axons cross the basal membrane and merge together,
forming vomeronasal nerves that run between the paired
olfactory bulbs and enter the accessory olfactory bulb at
the posterior dorsal side of the main olfactory bulb (see
Figs. 2 and 3).
The vomeronasal sensory neurons take origin from
the olfactory placode together with gonadotropin-releasing hormone (GnRH, also known as LH-RH or luteinizing
hormone-releasing hormone) neurons and ␥-aminobutyric acid (GABA)-containing neurons (405). The GnRH
FIG. 1. A schematic diagram showing
the anatomical pathways of the rodent
vomeronasal and main olfactory systems.
Physiol Rev • VOL
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
It has been believed for a long time that two chemosensory systems, the main olfactory and the vomeronasal
system, were responsible for different functions. The
main olfactory system was considered to be responsible
for recognizing the conventional volatile odorant molecules, whereas the vomeronasal system was thought to be
tuned for sensing pheromones. Recent studies have demonstrated that both chemosensory systems, together with
additional olfactory organs, are involved in pheromone
detection. In these systems, peripheral chemosensory
neurons located in the nasal cavity express distinct families of receptors that are believed to bind pheromones
and trigger a cascade of molecular and electrical events
that, ultimately, influence some aspects of the social behavior of the individual.
Although the main olfactory and vomeronasal systems share a similar histological organization, they also
display relevant differences with regard to the receptor
repertoire that they express and to the connections to
specific central areas (Fig. 1). Each system possesses
primary sensory neurons that send axons to second-order
neurons (mitral cells) in specific regions of the main
olfactory bulb (main olfactory system) or of the accessory
olfactory bulb (vomeronasal system). The mitral cells of
the main olfactory bulb project to several higher centers
including the piriform cortex and the cortical amygdala.
Instead, projections from the accessory olfactory bulb
only reach the medial amygdala and the posterior-medial
part of the cortical amygdala. From here fibers terminate
in the hypothalamus either directly or via the bed nucleus
of the stria terminalis.
In the following sections we summarize how the
various sensory epithelia are functionally organized. For
more exhaustive anatomical description, readers may
consult reviews specifically dedicated to these subjects
(66, 229).
927
928
TIRINDELLI ET AL.
neurons traverse the developing accessory olfactory bulb
and migrate to the mediobasal hypothalamus. The vomeronasal sensory neurons maintain a functional relationship with the hypothalamic areas in the adult mammal,
influencing neuroendocrine function and behavior.
Do humans have a vomeronasal organ? The presence
of a vomeronasal organ in human embryos, containing
bipolar neurons similar to developing vomeronasal neurons of other species, was clearly demonstrated (24, 25,
166, 240). However, the vomeronasal structure becomes
more simplified later in development (24), and several
Physiol Rev • VOL
reports showed that in human adults it becomes a blindended diverticulum in the septal mucosa (246, 383). Moreover, the epithelium lining the vomeronasal organ in human adults is different from that of other species and
from the olfactory or respiratory epithelium (259, 366).
Several studies showed that vomeronasal cells were not
stained by an antibody against the olfactory marker protein (259, 366, 383, 403), a protein abundantly expressed
in mature neurons in the olfactory and vomeronasal epithelia of other species (41, 144). In addition, Trotier et al.
(383) did not identify vomeronasal nerve bundles using a
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
FIG. 2. Chemosensory systems in rodents. A simplified sagittal section of a rodent head showing the location of sensory epithelia: the main
olfactory epithelium (MOE), the vomeronasal organ (VNO), the septal organ of Masera (SO), the Grueneberg ganglion (GG), and the location of the
main olfactory bulb (MOB) with necklace glomeruli (NG) and of the accessory olfactory bulb (AOB). Insets show images obtained from the intrinsic
green fluorescent protein (GFP) fluorescence from olfactory marker protein (OMP)-GFP gene-targeted mice. Top left: the Grueneberg ganglion has
an arrowlike shape with neurons clustered in small groups. Axons fasciculate and form a single nerve bundle. [Modified from Fuss et al. (82).] Top
right: the septal organ of Masera is an island of sensory epithelium. Axons from sensory neurons form two bundles. [Modified from Levai and
Strotmann (191).] Bottom left: mature vomeronasal sensory neurons in a coronal section of the vomeronasal organ. Bottom right: mature olfactory
sensory neurons in a coronal section of the main olfactory epithelium.
FROM PHEROMONES TO BEHAVIOR
specific antibody against the S-100 protein, a characteristic axonal marker (7). Therefore, even if vomeronasal
cells were chemosensitive, they would not have any obvious way to communicate with the brain. Therefore, the
development of human vomeronasal structures seems to
be limited to the embryonic stage, when they possibly
play a role in the migration of GnRH-secreting neurons
toward the brain (383; for review, see Ref. 246).
B. Olfactory Epithelium
Physiol Rev • VOL
C. Other Olfactory Organs
1. Septal organ of Masera
The septal organ of Masera (SO) (35, 311) is a small
island of olfactory epithelium within the nasal mucosa
and is located on each side of the septum, posterior to the
nasopalatine duct that connects the oral to the nasal
cavity (Fig. 2). The anatomical organization of the septal
organ appears similar to that of the main olfactory epithelium, although its neuroepithelium appears thinner
(218, 396). Furthermore, sensory neurons of the septal
organ are provided with cilia that bear shorter dendrites
and larger dendritic knobs than those of the main olfactory epithelium (218); their axons project to a subset of
glomeruli in the posterior, ventromedial main olfactory
bulb (Fig. 3). Interestingly, some of these glomeruli exclusively receive input from the septal organ sensory
neurons, while others are also innervated by fibers originating from the main olfactory epithelium (191). The anterior position of the septal organ, its vicinity to the
nasopalatine duct, and the wide distribution in many
mammalian species suggest that this organ may be functionally specialized for the early detection of biologically
relevant molecules as those, for example, resulting from
licking behavior.
2. Grueneberg ganglion
The Grueneberg ganglion is located bilaterally in the
anterodorsal area of the nasal cavity in the corners
formed by the septum and the nasal roof of several mammalian species (93a) (Fig. 2). Differently from other chemosensory organs, neurons of the Grueneberg ganglion
cluster in small groups (30, 79, 80, 317). Only two types of
FIG. 3. Projections of sensory neurons to the olfactory bulbs. Each
olfactory sensory neuron sends a single axon to the main olfactory bulb
(MOB). The apical and basal sensory neurons in the vomeronasal organ
(VNO) send axons, respectively, to the anterior and posterior accessory
olfactory bulb (AOB), a posterior dorsal region of the main olfactory
bulb. Neurons of the septal organ of Masera (SO) project to the posterior
ventromedial main olfactory bulb. Neurons of the Grueneberg ganglion
(GG) project to necklace glomeruli (NG) in the main olfactory bulb.
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
New evidence suggests that some pheromones are
detected by the main olfactory epithelium, a structure
specifically dedicated to sensing volatile molecules, usually named general odorants. The main olfactory epithelium lines cartilaginous protuberances in the posterior
nasal cavity (Fig. 2), called turbinates, and is a pseudostratified epithelium composed, similarly to the vomeronasal
epithelium, of three main types of cells: olfactory sensory
neurons, supporting cells, and basal cells. The olfactory
sensory neurons are bipolar neurons with a single ciliated
dendrite extending to the mucosal surface. Cilia are the
site of primary transduction for odorants and, possibly,
for pheromones. Although most olfactory sensory neurons have the same type of transduction cascade, it has
been recently discovered that subsets of olfactory neurons express different types of transduction molecules
(see sect. VB2) (reviewed in Refs. 31, 217, 269). Axons of
sensory neurons are segregated in the epithelium and,
after crossing the basal membrane, bundle together to
form the fascicles of the olfactory nerves. Axons project
to distinct domains of the main olfactory bulb as further
described in section V.
In the mouse, the olfactory system is functionally
established during the prenatal phase as the first axons
contact the telencephalic vesicle around embryonic day
13 (59, 320). Conversely, axonal targeting to specific domains within the main olfactory bulb occurs as early as
embryonic day 15.5 (59).
In the mouse, both vomeronasal and olfactory sensory neurons undergo a continuous renewal process
throughout adult life, and their rate of neurogenesis is
also regulated by environmental factors (11, 59, 90, 91,
402). Readers interested in the neurogenesis of olfactory
and vomeronasal sensory neurons may consult some recent reviews on the subject (101, 205).
The human olfactory epithelial region is a small area
(1–5 cm2) covering part of the superior turbinate, septum,
and roof of the nasal cavity. The transition boundary
between olfactory and respiratory epithelium is not as
sharp as in other mammals; instead, there is an irregular
zone where olfactory and respiratory cells intermingle.
Moreover, patches of respiratory epithelium have been
identified in purely olfactory areas (262).
929
930
TIRINDELLI ET AL.
IV. PUTATIVE PHEROMONE RECEPTORS
A. Vomeronasal Receptors
In the vomeronasal organ of rodents, detection of
pheromones is mediated by two anatomically independent pathways that originate from chemosensory neurons
of distinct compartments, basal and apical, of the neuroepithelium. The organization of the vomeronasal epithelium of rodents in apical and basal neurons was initially
based on the specific expression pattern of marker molecules (261, 340, 341). More recently, exhaustive data
established that also distinct transduction components
characterize these two subsets of chemosensory neurons.
The G protein subunit G␣i2 was found to enrich the apical
neurons, whereas G␣o was identified in the basal ones
(19, 143). These important findings are credited to have
addressed the efforts towards the search of receptors that
could specifically couple to G proteins. Thus two large
families of vomeronasal G protein-coupled receptors,
named V1R and V2R, have been discovered, and their
expression pattern was extensively analyzed. Moreover,
other putative receptors as the major histocompatibility
complex class Ib molecules (H2-Mv) have been identified
in the vomeronasal organ, and their expression was demonstrated to be tightly linked to that of V2Rs.
ily compared with other species (313, 346, 415, 422). The
gene tree of functional V1Rs in placental mammals, based
on the mouse repertoire, is split into 12 clades (a-l) (Fig. 4)
(313). All rat V1R receptor genes, except for one, are
distributed within 10 of these 12 clades (a-g and j-l). Thus
the mouse h and i clades represent an instructive example
of species specificity, exclusively including mouse sequences. Given that one rat pseudogene is placed in each
of the h and i subfamilies, it is likely that the absence of
rat functional genes in these clades is the result of deletion in this species and of postspeciation expansion in the
mouse lineage.
V1Rs in the mouse account for almost 200 potentially
functional receptors and are highly expressed in vomeronasal neurons, at least at mRNA level. The mouse has a
larger V1R repertoire than the rat (Fig. 5).
In the mouse V1R phylogenetic tree, clades share
from 15 to 40% of their amino acid residues between each
other, with intrafamily identities varying between 40 and
99%. High sequence variability is found in transmembrane
domain 2, in intracellular loop 1, and in extracellular
loops 2 and 3, while transmembrane domain 3 is highly
conserved. A few amino acid residues mostly located
between transmembrane domains 5 and 6, and a potential
glycosylation in extracellular loop 2, are common features
to virtually all members of the 12 V1R clades, while specific amino acid signatures are characteristic of each of
the 12 families (313).
1. V1Rs
The first family of vomeronasal receptors, V1Rs, was
discovered with elegant experiments of differential
screening, based on the assumption that the genetic profile of two vomeronasal neurons expressing G␣i2 only
differs for the receptor RNA that they express (71, 285). It
was also evident that V1R expression is restricted to the
apical neurons of the vomeronasal organ which express
G␣i2 (71, 285). From the structural point of view, V1Rs are
related to group I of G protein-coupled receptors and, in
the mouse, represent a strikingly expanded receptor famPhysiol Rev • VOL
FIG. 4. The mouse vomeronasal receptor families. Top: phylogenetic tree of V1Rs showing 12 phylogenetically isolated subfamilies.
Bottom: phylogenetic tree of the 4 V2R families showing their clades.
Numbers of receptors for each clade and family are shown in red.
[Modified from Silvotti et al. (355).]
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
cells have been identified in the Grueneberg ganglion:
glial cells and ciliated neurons (30). Each neuron has a
single axon that fasciculates with others into nerve bundles that project to ⬃10 glomeruli in the main olfactory
bulb (Fig. 3). These glomeruli are a subpopulation of the
necklace glomeruli. Since necklace glomeruli are active
during the suckling behavior of the pup, it was suggested
that the Grueneberg ganglion may sense maternal pheromones (82, 173, 317). However, a very recent study has
rejected this hypothesis, suggesting instead that alarm
pheromones, volatile molecules released by animals to
alert conspecifics about danger, are the likely substrates
of the Grueneberg ganglion neurons. These results are
discussed in more detail in section VB.
FROM PHEROMONES TO BEHAVIOR
931
V1R genes are scattered in regions of many chromosomes, where members of a given family are found almost
invariably clustered. It has been recently reported that
V1R clusterization may not simply reflect gene expansion
but may be actively linked to the mechanisms underlying
the tight regulation of receptors’ expression. V1R expression is, indeed, monoallelic and monogenic, both of these
being tools that achieve the expression of one receptor
type per cell, thus contributing to the narrow tuning of the
vomeronasal neurons (318).
At present, some vertebrate genomes of species of
different orders have been extensively scrutinized for the
Physiol Rev • VOL
presence of V1Rs (Fig. 5). Interestingly, V1R genes were
found in all species except in chicken and chimpanzee, in
which the presence of a functional vomeronasal organ has
not been consistently demonstrated. The number of functional V1R genes drops dramatically from rodents to other
species. The human genome contains several V1R pseudogenes, only five human V1R genes that have an intact
open reading frame, and at least one V1R that is transcribed in the olfactory epithelium. The chimpanzee genome has no obvious functional V1R genes but a large
collection of pseudogenes. This is surprising if we consider that the massive loss of V1R genes in chimpanzee
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
FIG. 5. Evolution of nasal chemosensory receptor gene repertoires in vertebrates. The tree shows the phylogeny of 13 species of vertebrates:
8 mammals (platypus, chimpanzee, human, mouse, rat, dog, cow, and opossum), 1 bird (chicken), 1 amphibian (western clawed frog), and 3 teleost
fishes (green spotted pufferfish, fugu, and zebrafish). The numbers of intact genes for V1R, V2R, OR, and TAAR receptors are given together with
the numbers of nonintact genes shown in parentheses. Numbers of receptor genes in each species were taken from Grus et al. (94), except for
chimpanzee which were obtained from Nei et al. 2008; V1Rs and V2Rs in green spotted pufferfish, fugu, and zebrafish from Shi and Zhang (346); ORs
in fugu from Niimura and Nei (274); and TAARs in fugu and zebrafish from Hashiguchi and Nishida (105).
932
TIRINDELLI ET AL.
2. V2Rs
A second family of vomeronasal receptors, V2Rs, was
discovered in 1997 (110, 234, 323) and, in mouse, it accounts for ⬃120 potentially functional members. V2Rs
belong to group III G protein-coupled receptors and are
characterized by a long NH2-terminal domain. This region
has been demonstrated to represent the binding domain
in most receptors of group III (for review, see Ref. 293).
V2Rs are strictly coexpressed with the G protein subunit
G␣o in the basal neurons of the vomeronasal organ.
In the mouse, intact V2Rs can be grouped in four
distinct families (A-D) according to sequence homology
(Fig. 4). Together, receptors of families A, B, and D represent 95% of all V2Rs. Their expression pattern strongly
suggests that these receptors are expressed by mutual
exclusion in the vomeronasal organ. A phylogenetic analysis shows that family C V2Rs (also known as V2R2
family) are very divergent from those of families A, B, and
D (3, 231, 323), and these differences are also functional
(354). This family is also noteworthy for the high sequence conservation between species. Moreover, compared with other V2R pseudogenes, family C pseudogenes
have one or very few inactivating mutations (also in the
Physiol Rev • VOL
human pseudogenes), suggesting a relatively recent loss
of function.
In rodents, this receptor superfamily is largely expanded and, in contrast to V1Rs, it appears completely
degenerated in primates, dogs, and cows (Fig. 5). In the
dog, the lack of a sizable V1R and V2R family does not
probably reflect a reduction in the pheromonal relevance
in triggering intraspecific behavior, but rather a general
decline of the vomeronasal system, a hypothesis that is
supported by anatomical data (63). In contrast, in the
opossum, V2Rs are abundantly represented (Fig. 5). Thus
it appears that the evolution of opossum and rodent V2Rs
genes parallel that of the V1R ones in these species, which
expanded independently. Moreover, most opossum and
rodent V2R and V1R genes arose by duplication (rodent
V2R and V1R genes are found in genomic clusters, and
closely related V2Rs tend to reside near one another in
the genome) after these lineages diverged. This receptor
diversification and the consequent detection of additional
pheromones may have resulted in a selective advantage.
The expression pattern of V2Rs makes them different
from V1Rs and odorant receptors in that each basal neuron expresses two V2Rs, one of which is a member of
family C (231). Moreover, there is evidence for a combinatorial scheme for the detection of chemosensory cues
in the basal neurons of the vomeronasal organ. Recent
data indicate the existence of multimolecular complexes
in individual basal neurons, constituted by MHC molecules, by family A, B, or D and family C V2Rs (355). For
example, neurons expressing family A receptors of the
highly represented clades I, III, and IV almost exclusively
express the family C receptor Vmn2r2, whereas neurons
expressing clade VIII receptors specifically express the
family C receptor Vmn2r1. Interestingly, coexpression of
Vmn2r1 and Vmn2r2 with family A V2Rs seems to involve
whole clades rather than individual receptors. Thus a
peculiar modality of expression subsists in the basal layer
of the vomeronasal organ, where two family receptors are
mutually exclusively expressed in the same basal neurons.
It has been suggested that V2Rs bind MHC peptides
(185) and MUPs (47), as further discussed in section VA.
A V2R receptor, V2R83, is also expressed outside the
vomeronasal organ, in a large subset of neurons of the
Grueneberg ganglion (79), and it is a good candidate
receptor for alarm pheromones (30).
At present, however, direct evidence that V2Rs are
directly involved in pheromone signaling is still missing.
At least 50% of the basal vomeronasal neurons express another multigene family representing nonclassical
class I major histocompatibility complex (H2-Mv) genes.
Each of the nine identified H2-Mv genes is present in a
subset of vomeronasal neurons, and multiple H2-Mv
genes can be coexpressed in the same neuron (127, 207).
H2-Mv molecules are also coexpressed in a combinatorial
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
appears even greater than for the human repertoire, in
which V1R genes have been identified. However, the most
surprising observation is that the dog only possesses eight
functional V1R genes spread among four clades. This
finding is unexpected because the dog is traditionally
renowned to have evolved an extraordinary well-developed olfactory system and a complex social behavior
(346). Although it is plausible that V1Rs emerged as dominant receptors for triggering specific rodent behaviors
(multiple losses of ancestral V1R genes in carnivores and
artiodactyls and gains of many new genes by gene duplication in rodents), it cannot be alternatively ruled out that
a loss of dog V1Rs is due to the breeding and domestication processes of this species. Therefore, diversification
into subfamilies probably occurred prior to the rodentdog-primate split early in the mammalian evolution (415).
The question as to whether V1Rs are exclusively
expressed in the vomeronasal epithelium or also in other
tissues is still controversial. Karunadasa et al. (152) reported the expression of V1Rd RNA in some cells scattered within the main olfactory epithelium of embryonic
and postnatal mouse and suggested that pheromone detection may also take place in this structure. V1R-homolog RNAs have also been identified in the main olfactory epithelium of goats and humans (316, 394). However,
caution must be taken before raising any hypothesis as,
until now, there are no clues that a functional V1R is
expressed at a detectable level in the main olfactory
epithelium.
FROM PHEROMONES TO BEHAVIOR
manner with receptors of specific subfamily of V2Rs (127,
207). It has been initially proposed that H2-Mv may function as indispensable escort molecules in the transport of
V2Rs to the plasma membrane (207). However, this chaperone-like function is not generalized to all V2Rs (127,
354), and moreover, it is likely to be exclusively restricted
to the mouse or rat as the opossum genome contains a
hundred functional V2R genes and no H2-Mv genes (346).
3. Ligands of vomeronasal receptors
Physiol Rev • VOL
To deorphanize V1Rs and study their binding properties, it is necessary to develop an efficient bioassay. At
present, it is only plausible to assume that V1Rs are
vomeronasal receptors narrowly tuned to bind small volatile pheromones, possibly through a hydrophobic pocket
located in their transmembrane region. Nevertheless,
achievements in this direction have been very recently
accomplished by Shirokova and colleagues who managed
to deorphanize the human V1Rs employing NH2-terminally tagged receptors expressed in the cell line HeLa/Olf
(350). It appears that, in sharp contrast to V1Rs, each
human recombinant V1R receptor responds to several
aliphatic alcohols or aldehydes via the olfactory G protein
G␣olf (350). Unfortunately, a pheromonal function of human V1R ligands has yet to be demonstrated.
4. Formyl peptide receptors
Recently, a group of receptors belonging to the family of formyl peptide receptors (FPRs) has been exclusively identified in the vomeronasal organ (310a). FPRs
are G protein-coupled receptors expressed in all mammals, with three genes present in the human genome and
seven in the mouse genome (249a). At least two of the
mouse FPRs have been reportedly involved in immune
and inflammation responses being expressed in granulocytes, monocytes, and macrophages (183a). An in situ
analysis of the vomeronasal organ of the mouse indicates
that five FPRs are expressed in a small subset of chemosensory neurons of the apical layer. Antibody staining
reveals that FPRs are featured by a subcellular distribution similar to that of V2R receptors including, besides the
microvillar surface, the dendrite and somata. Although
Fpr expression is restricted to the apical layer and, as
well as VIRs, adopts monogenic transcription, no coexpression is evident with receptors of this large family,
suggesting that FPRs characterize a new and specific
class of chemosensory neurons. Experiments based on
calcium imaging on the whole vomeronasal epithelium
and in dissociated chemosensory neurons establish that
indeed FPR agonists, such as some microbial, antimicrobial, and viral peptides, stimulate neuronal subsets with a
nonidentical affinity profile. These findings indirectly
prove that these receptors play a role in offering a barrier
against immunological diseases, or perhaps also in pheromonal communication itself, given that formyl peptides
have been detected in urine (52a) and that, in the olfactory and vomeronasal epithelium, two FPR isotypes appear to be expressed in the glandular but not chemosensory compartment (310a).
B. Odorant Receptors
Odorant receptor genes were discovered in 1991 by
Buck and Axel (39). The vast majority of the olfactory
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
Although an initial report provided clues on the possibility that V1Rs could be expressed in a heterologous
system (99), the problem of expression later became the
major barrier to the identification of ligands for vomeronasal receptors, as well as for odorant receptors (for
reviews, see Refs. 224, 256, 312). Some information about
ligands for V1Rs was obtained from physiological approaches to the problem. Results from calcium imaging
experiments on slices of the mouse vomeronasal organ
indicated that some molecules credited as pheromones
stimulated specific vomeronasal neurons that, for their
topographic position in the vomeronasal organ, were
likely to express the V1R repertoire (187). Furthermore,
these authors reasoned that, if neurons responding to
these pheromones indeed reflected a V1R-dependent activity, V1Rs must be narrowly tuned to bind one or very
few pheromones. This contrasts with the broad selectivity
of olfactory sensory neurons that can be stimulated by a
large array of odorants. Further and more direct investigations on the V1R ligand specificity lead to the conclusion that V1Rs are indeed qualified as pheromone receptors: Del Punta et al. (61) succeeded to delete a large
genomic region containing ⬃16 V1R genes of subfamily a
and b, including the receptor V1rb2. The behavior of these
mutant animals was also thoroughly investigated showing
remarkable deficits (see sect. VII). Moreover, electrophysiological responses of vomeronasal neurons to some
pheromones, including 6-hydroxyl-6-methyl-3-heptanone,
n-pentylacetate, isobutylamine, and 2-heptanone, were altered (Table 1). Thus the repertoire of V1R receptor candidates to respond to these pheromones was restricted to
a relatively limited number.
Another approach, developed by Boschat et al. (28),
led to the identification of the first pheromone-V1R pair in
mammals. These authors employed mouse mutant lines
with a GFP-targeted V1rb2 allele, allowing the visualization and the isolation of the V1Rb2-expressing neurons for
calcium imaging based experiments. Thus a direct correlation was established between V1Rb2 expression and the
response of vomeronasal neurons to the mouse pheromone 2-heptanone (Table 1). As a confirmation of the
strong selectivity of V1Rs above reported, 2-heptanonerelated compounds were found to be ineffectual to stimulate V1Rb2-expressing sensory neurons.
933
934
TIRINDELLI ET AL.
same neurons, suggesting that subsets of neurons exist in
the olfactory epithelium that are not exclusively dedicated to the detection of conventional odorants. By the
expression of olfactory TAARs in heterologous cell systems it was demonstrated that at least four of them
(mTAAR3, mTAAR4, mTAAR5, mTAAR7f) bind smallmolecule amines and that each of these receptors recognizes a specific set of ligands, among which are ␤-phenylethylamine, isoamylamine, and trimethylamine that are
natural components of the mouse urine. ␤-Phenylethylamine, a ligand for mTAAR4, is contained in human and
rodent urine, where its concentration rises during stressful situations (93, 286, 361). Isoamylamine and trimethylamine bind to mTAAR3 and mTAAR5, respectively, and
are present at higher concentrations in male compared
with female urine (84, 276). Noteworthy, isoamylamine
has been reported to act as a pheromone accelerating
puberty onset in female mice (276). Therefore, stimulation of mTAARs seems to affect the capacity of discriminating gender and social status of an individual. These
findings strongly contribute to confirm that the main olfactory epithelium is involved in the elicitation of physiological responses and innate behaviors.
In the mouse, TAARs are also expressed in subsets of
neurons of the Grueneberg ganglion (80). As in the main
olfactory epithelium, also in the Grueneberg ganglion,
distinct TAAR subtypes are expressed in nonoverlapping
subsets of neurons that do not coexpress the vomeronasal
receptors V2Rs. Since the number of neurons expressing
TAARs and V2Rs in the Grueneberg ganglion is much
higher during the perinatal stage than in the adult, it was
suggested that these neurons may be important for interactions between pups and their mother (80).
TAARs have also been identified in the genome of
many species including humans and fish, with the number
of putatively functional TAAR genes greatly varying
among species (105) (Fig. 5).
C. Trace Amine-Associated Receptors
D. Olfactory-Specific Guanylyl Cyclase
Type D Receptor
In 2006, Liberles and Buck (194) reported the expression of a second class of chemosensory receptors in the
mouse olfactory epithelium. These receptors are the
“trace amine-associated receptors” (TAARs) (27, 192,
203), a family of G protein-coupled receptors discovered
in 2001 (27, 42) (for review, see Ref. 425). TAARs are
unrelated to odorant receptors but have similarities with
receptors for biogenic amines, such as serotonin and
dopamine receptors. The TAAR gene family includes 15
members in the mouse (Fig. 5), and each of them, except
for TAAR1, is expressed in the olfactory epithelium (194).
Moreover, the level of expression of TAAR genes is similar to that of odorant receptors, although TAARs and
odorant receptors do not appear to be coexpressed in the
Physiol Rev • VOL
A small subset of olfactory sensory neurons localized
in the dorsal region of the main olfactory epithelium
expresses a specific subtype of the transmembrane receptor guanylyl cyclase, named guanylyl cyclase type D
(GC-D) (81, 149). The GC-D receptor is encoded by one of
the seven guanylyl cyclase genes with whom it shares a
similar topology: all receptor guanylyl cyclases retain an
NH2-terminal ligand binding region, a single transmembrane loop, and a COOH terminus that includes the catalytic domain (81, 296). Specific ligands have been identified for some guanyly cyclases, but not for GC-D yet;
however, potential ligands for this receptor have been
recently assessed. In the mouse, for example, neurons
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
neurons express only 1 of the 1,000 odorant receptor
genes (51, 226, 273, 308, 370, 389, 389). Neurons expressing different receptor types are usually found in one of
four partially overlapping zones of the olfactory epithelium, within each of which a given odorant receptor appears to be randomly distributed (128, 251, 308, 369, 389).
An olfactory sensory neuron typically responds to
more than one type of odorant molecules, and a given
odorant can activate neurons with different specificity
(70, 77, 226, 351). For the discrimination and perception
of odorants, the odorant receptor family seems to employ
a combinatorial scheme, with one single odorant molecule activating several types of receptors and each receptor sensing several different types of odorants, thereby
producing a unique combination of receptors activated by
each odorant molecule. This combinatorial scheme allows the olfactory system to recognize an enormous number of odorants (40, 256).
Odorant receptor genes have been identified in regions outside the main olfactory epithelium as the septal
organ and the vomeronasal epithelium. About 50 odorant
receptor genes are expressed in the mouse septal organ
(150, 378). Moreover, a subset of 44 odorant receptor
genes is also expressed in the vomeronasal epithelium
(190), although their functional role has not been established.
At present, we do not possess evidence that odorant
receptors can also function as pheromone receptors.
However, intriguingly, recent experiments linked the expression of a subclass of odorant receptors with stereotyped behavioral expressions in the mouse, thus suggesting that odorant receptors can convey information both
about the quality of an odorant and the individual that has
emitted it (169).
For additional information about odorant receptors,
the reader is referred to several reviews that extensively
discuss this subject (86, 224, 225, 256).
FROM PHEROMONES TO BEHAVIOR
V. PHEROMONE TRANSDUCTION
Distinct neuronal subsets of the olfactory organs express specific types of receptors, second messenger systems, and ion channels to transform the pheromone binding into electrical signals.
The electrical activity of chemosensory neurons in
response to stimuli can be measured with a number of
approaches, including recording of the field potential, of
the firing activity with an array of extracellular electrodes
and by patch-clamp technique at the single-cell level.
Nonetheless, calcium imaging either in dissociated cells
or epithelium slices is often employed as well as the
metabolic activity of stimulated neurons is frequently
monitored upon pheromonal stimulation by evaluating
the expression of the early protooncogenes.
A. Vomeronasal Sensory Neurons
1. Signal transduction molecules
The detailed signal transduction cascade in the
vomeronasal organ is still largely unknown, although it
has been clearly demonstrated that pheromone stimuli
cause membrane depolarization and increase the action
potential firing rate in vomeronasal sensory neurons (28,
53, 61, 68, 115, 124, 125, 185, 187, 211, 363, 374). As
described in previous sections, the sensory epithelium of
the vomeronasal organ presents two subsets of neurons:
apical and basal. Apical neurons express members of the
V1R family of vomeronasal receptor genes together with
G␣i2, whereas basal neurons express V2R receptors and
G␣o (19, 102, 142, 143, 321, 348, 371, 398). The anatomical
and perhaps functional segregation in the vomeronasal
epithelium is likely to be a prerogative of species as rats,
Physiol Rev • VOL
mice, hamsters, opossum, and platypus that possess functional V2R genes (Fig. 5).
Krieger et al. (176) reported that whole male urine
induces inositol 1,4,5-trisphosphate (IP3) production via
the activation of G␣i2 as well as G␣o, suggesting a role for
both G protein subtypes in transducing pheromonal responses mediated by urinary components. Moreover, according to these authors, the activation of G␣i2 and G␣o
subtypes in the vomeronasal epithelium appears to be
caused by two structurally different classes of molecules.
Whereas G␣i2 activation was only observed upon stimulation with lipophilic volatile components, G␣o activation
was elicited by the rat major urinary protein, ␣2-globulin.
A confirmatory role of these G proteins in pheromone
transduction issues from behavioral (see sect. VII) and
anatomical studies performed on G␣i2 and G␣o negative
mutant mice (278, 373): both mouse lines show a remarkable reduction in the size of the neuronal layers where
G␣o and G␣i2 were originally expressed, indicating that G
protein-mediated activity is required for survival of vomeronasal sensory neurons.
Since PLC stimulation and IP3 production appear to
be predominantly mediated by the release of the ␤␥ complex of the heterotrimeric G proteins, Go and Gi (56), it
has been proposed that G␤␥ may play an important role in
the transduction process of the vomeronasal epithelium.
One ␤-subunit (G␤2) and two ␥-subunits (G␥2 and G␥8)
have been described in the vomeronasal epithelium (321,
322, 380). G␥2 immunopositive neurons are exclusively
localized in the apical layer, whereas G␥8 neurons are
preferentially restricted to the basal one. G␤2, on the
other hand, seems to be active in both neuronal subsets
(Fig. 6). Biochemical studies reveal that IP3 formation
induced by stimulation of membrane preparations with
volatile urinary compounds was selectively blocked by
an anti-G␤2 antibody, whereas second messenger production induced by stimulation with major urinary proteins was inhibited by preincubation of the vomeronasal membranes with an antibody recognizing G␥8 (321).
Thus it appears that the G␣o␤2␥8 complex may control
PLC activation by proteinaceous pheromones, whereas
G␣i2␤2␥2 neurons respond to urinary volatile compounds.
Other lines of evidence indicate that, at least in some
vomeronasal sensory neurons, transduction is mediated
by a phosphatidylinositide signaling pathway. Some studies showed that electrical responses, IP3 production, and
calcium entry upon pheromonal stimulation (53, 176, 177,
215, 330, 397) are specifically blocked by the phospholipase C inhibitor U73122 (115, 363). Recently, Thompson
et al. (377) demonstrated that compounds known to induce pregnancy block (Bruce effect), such as MHC peptides and male urine, also increased the IP3 formation,
confirming a major role of the phosphatidylinositide pathway in the vomeronasal signal transduction.
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
expressing GC-D respond to dilute urine and to the intestinal peptide hormones uroguanylyn and guanylin, while
neurons in GC-D knockouts are unresponsive to these
compounds (186). Given their localization in the enteric
tract, uroguanylyn and guanylin are supposed to contribute to the maintenance of salt and water homeostasis or
to the detection of cues related to hunger, satiety, or thirst
(186). GC-D neurons also express carbonic anhydrase II
and respond to CO2 stimuli (118). It is not clear if there is
a possible interplay between these two chemosensory
responses.
Despite its exquisite expression in the main olfactory
epithelium, GC-D appears to be a pseudogene in species
lacking a functional vomeronasal organ such as humans,
apes, Old World and New World monkeys, and tarsier.
This suggests that a functional GC-D gene was lost early
in primate evolution and that chemical detection in most
primate species is unlikely to involve GC-D (416).
935
936
TIRINDELLI ET AL.
The ion channel TRPC2, a member of the superfamily
of transient receptor potential channels (390, 412), is
expressed in the microvilli of both apical and basal neurons (197, 241, 271) and plays an important role in the
transduction process of some, but not all, vomeronasal
sensory neurons. Interesting to note, TRPC2 is unique
among TRP channels in that a functional gene has been
lost from the Old World monkey and human genomes
(198, 420). Proposed mechanisms for the activation of the
TRPC2 channel in the vomeronasal neurons are similar to
those involved in the signaling cascade of Drosophila
photoreceptors (103, 250): pheromones trigger the activa-
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
Physiol Rev • VOL
89 • JULY 2009 •
www.prv.org
FROM PHEROMONES TO BEHAVIOR
level in the gartner snake (215) and in rodents (215, 319,
424). However, the decrease in cAMP occurs with a longer
delay with respect to pheromone stimulation (319). Possible targets for modulatory actions by cAMP include the
Ca2⫹-activated cation channel (195) and the recently
characterized hyperpolarization-activated cyclic nucleotide-gated (HCN) cation channels (64).
It is intriguing that a subpopulation of apical sensory
neurons expresses members of the odorant receptor gene
family, together with G␣i2 and TRPC2 (190). Since some
conventional odorants are detected by the vomeronasal
epithelium (329, 382, 409), it was suggested that vomeronasal neurons expressing odorant receptors might be
those that respond to odorants (190).
Following the initial transduction events, when the
membrane potential of vomeronasal sensory neurons
reaches threshold, action potentials are generated and
propagate along the axons to the accessory olfactory
bulb. Interestingly, most vomeronasal neurons fire tonically for several seconds, without sign of adaptation,
when small current injections (64, 196, 347, 384) or dilute
urine are applied for up to 2–3 s (211, 421). However, in
one study, longer stimulations of 10 – 60 s caused spikefrequency adaptation (384), and in addition, a recent work
suggested that sensory adaptation in vomeronasal sensory neurons requires the influx of Ca2⫹ and is mediated
by calmodulin (362).
2. Responses to urinary pheromones
Using an array of extracellular electrodes, Holy et al.
(115) simultaneously recorded the activity of a large number of mouse vomeronasal neurons in response to dilute
urine. Neurons responded to urine with an increase in the
firing frequency that remained approximately unchanged
if the stimulus was applied for as long as 100 s, indicating
that the response of these neurons do not adapt to prolonged stimulus exposure. Furthermore, they identified
distinct and nonoverlapping subsets of neurons that were
selectively activated by substances contained in female or
male urine; in contrast, about half of the vomeronasal
neurons detected stimuli that were independent of sex.
FIG. 6. Transduction mechanisms in vomeronasal and olfactory sensory neurons. A: a scanning electron micrograph of the knob of a rat
vomeronasal sensory neuron showing several microvilli. Scale bar, 1 ␮m. [From K. Doving and D. Trotier, unpublished data; see also Trotier et al.
(382a).] B: binding of pheromone molecules to vomeronasal receptors in the microvilli activates a transduction cascade involving V1R, G␣i2␤2␥2
in apical neurons, or V2R, G␣o␤2␥8 in basal neurons. The ␤␥ complex activates the PLC that in turn may cause an elevation of IP3 and/or DAG. The
gating of the TRPC2 channel allows an influx of Na⫹ and Ca2⫹, causing a membrane depolarization. An additional cation-selective channel activated
by Ca2⫹ may be involved in the transduction cascade. The TRPC2 channel appears to be the principal transduction channel in V1R-expressing
neurons, whereas its role in V2R-expressing neurons in unclear. C: a scanning electron micrograph of the knob of a human olfactory sensory neuron
showing the protrusion of several cilia. Scale bar, 1 ␮m. [From Morrison and Costanzo (262).] D: olfactory transduction takes place in the cilia, where
ligands bind to odorant receptors (OR) located in the ciliary membrane, thus activating a G protein (G␣olf) that stimulates adenylyl cyclase type 3
(ACIII), producing an increase in the generation of cAMP from ATP. cAMP directly gates ion channels (CNG), causing an inward current carried
by Na⫹ and Ca2⫹. Ca2⫹ entry amplifies the signal by causing the efflux of Cl⫺ through Ca2⫹-activated Cl⫺ channels (Cl-Ca). These ion fluxes cause
a depolarization. Inhibitory actions are played by the complex Ca2⫹-calmodulin (CaM) by increasing the activity of the phosphodiesterase (PDE)
that hydrolyzes cAMP and by reducing the affinity for cAMP of the CNG channels. Ca2⫹ is extruded by the Na⫹/Ca2⫹ exchanger (Na/Caex).
Physiol Rev • VOL
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
tion of G protein-coupled receptors that in turn stimulate
phospholipase C and the production of the lipid messenger diacylglycerol (DAG) (211) and possibly of polyunsaturated fatty acids such as arachidonic acid (363, 421) and
linolenic acid (Fig. 6) (363).
Genetic ablation of the TRPC2 channel results in a
strong, but not complete, reduction of electrophysiological responses of the vomeronasal neurons to urine or its
volatile components (156, 193, 368). Moreover, male
TRPC2-knockout mice showed severe behavioral abnormalities (see sect. VII) (157, 193, 236, 368), indicating that
the TRPC2 channel may be an important component of
the transduction of pheromone signals.
Nevertheless, the genetic ablation of TRPC2 does not
completely abolish the response of the vomeronasal organ to pheromones, since electrophysiological recordings
have demonstrated that basal neurons of mutant mice still
respond to MHC peptide ligands as controls (156). In
addition, TRPC2-knockout mice exhibit the Bruce effect
(156), a neuroendocrine response that requires a functional vomeronasal organ. All together, these studies suggest that at least a subset of basal neurons uses a different
channel from TRPC2 for the transduction of some pheromonal stimuli.
Interestingly, Liman (195) showed that a cation channel activated by a high concentration of Ca2⫹ is expressed
in hamster vomeronasal sensory neurons. This channel is
permeant to both Na⫹ and K⫹ and is blocked by ATP and
cAMP, sharing many features with the TRPM4 ion channel
(275). These properties make this channel a good candidate to be directly involved in sensory transduction or in
amplifying a primary sensory response. At present, the
molecular nature of this channel and its relation with
TRPC2 or with IP3 production are not yet resolved.
Some components of the cAMP signaling cascade
have been identified in vomeronasal sensory neurons.
These include adenylyl cyclase types 2 (19), 4, 5 (184), and
6 (184, 319); the subunit CNGA4 of the olfactory cyclic
nucleotide-gated channel (19, 184); and phosphodiesterase types 1 and 4 (50, 183), suggesting that cAMP may play
a physiological role in vomeronasal sensory neurons. Indeed, some pheromones induce a reduction of the cAMP
937
938
TIRINDELLI ET AL.
Physiol Rev • VOL
an exceptionally low-noise imaging of large neuronal populations, unexpectedly found that responses to dilute
urine are predominantly, and perhaps exclusively, generated by neurons in the apical layer, where V1Rs are expressed.
Leinders-Zufall et al. (187) showed that some volatile
hydrophobic pheromones, known to be tightly bound to
the hydrophobic pocket of MUPs, produced a calcium
increase in neurons of slice preparations of the vomeronasal organ. The distribution pattern of the activated
neurons by any of these molecules reflects a localization
that can be associated with the expression of the vomeronasal receptors V1Rs (187), and the percentage of neurons responding to an individual pheromone (0.2–3%)
matches the proportion of neurons expressing one type of
V1R. Therefore, when volatile molecules are released by
MUPs, in the external environment or in the nasal cavity,
they are likely to bind and activate V1Rs.
Urine contains compounds whose physiological role
is largely unknown. Nodari et al. (277) applied chromatographic fractions of female mouse urine to isolated vomeronasal epithelia mounted on a multielectrode array and
measured the spiking activity of neurons (277). Increase
in the firing rate was associated with compounds with
homogeneous chemical properties that, after further purifications, appeared to correspond to sulfated steroids.
Sulfatase-treated urine extracts lost ⬎80% of their activity, indicating that sulfated compounds are the predominant ligands for vomeronasal neurons in female urine.
Interestingly, a collection of 31 synthetic sulfated steroids
triggered responses 30-fold more frequently than did a
similarly sized stimulus set containing the majority of all
previously reported ligands. Vomeronasal neurons seemed
to have a high tuning specificity for sulfated steroids, since
each neuron responded only to one or to a few structurally similar compounds, but collectively, neurons detected all major classes of sulfated steroids. Some of
these compounds may be indicators of a status of stress.
Indeed, the concentration of two sulfated glucocorticoids
in urine increased about threefold in animals subjected to
stress compared with unstressed controls, indicating that
information about the status of stress of these mice is
encoded in urine and could be detected by vomeronasal
sensory neurons (277).
3. Responses to secretory peptides
Kimoto et al. (163) showed that mouse recombinant
exocrine gland-secreting peptides (see sect. IIC) evoked
electrical activity in the vomeronasal but not in the olfactory epithelium. Interestingly, when the electrical responses of individual neurons to ESP1 were recorded
with a multielectrode array, only a small percentage of
neurons, 1.6%, responded to ESP1 (163). This result is consistent with the possibility that ESP1 is recognized by one
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
Thus these findings establish the capacity of the vomeronasal organ to identify differences between sexes.
A recent study by He et al. (106) further stressed the
concept of the vomeronasal organ being subtly dedicated
to the identification of conspecifics. These authors quantified the number of neurons responding to individual
urine of either sex and of different strains in slices of the
vomeronasal epithelium of mice genetically modified to
express a calcium indicator. It was observed that only a
very small subpopulation of neurons (1–3%) exclusively
responded to urine samples from all individuals of the
same sex. The gender-specific neurons were evenly distributed both in the apical and basal layers of the epithelium and did not differ in males or females, consistent
with the little dimorphism generally observed at the level
of receptor expression in the vomeronasal epithelium.
Interestingly, neurons specific to female urine showed
different activation patterns depending on the phase of
the estrus cycle; in contrast, urine from castrated mice
failed to elicit a response in male-specific neurons. Furthermore, information about individuals and strains is not
encoded by specialized neurons but, rather, by the combinatorial activation of vomeronasal neurons, as no urine
samples from different mice of the same sex elicited the
same pattern of neuronal activity. Thus the information
contained in the mouse urine about gender is encoded by
dedicated neurons; strains and individuals are recognized
by a combinatorial code of neuronal activation (106), with
the obvious advantage to discriminate a large population
of individuals with a restricted repertoire of receptors.
He et al. (106) also compared responses elicited by
synthetic strain-specific MHC class I peptides with those
induced by urine of the same strain and found that neurons activated by urine and by the peptides did not overlap. Furthermore, the same authors (106) pointed out that
only fragments, but not the peptides, of the MHC complexes are detectable in urine (358). This result strongly
contrasts with what was observed by Leinders-Zufall et al.
(185), who identified a specific subset of basal neurons
highly sensitive to MHC class I peptides and to urine from
mice of the relevant haplotype.
Other studies investigated the response of the vomeronasal epithelium to specific urinary components, such
as MUPs, that represent another class of candidates for
strain recognition, as well as for heterozygosity signals
(375). Kimoto et al. (163) found that a recombinant isoform of MUPs elicits electrical responses in the vomeronasal epithelium, and Chamero et al. (47) showed that
native and recombinant MUPs evoke calcium influx in
sensory neurons expressing the G protein subunit G␣o.
Since G␣o coexpresses with V2Rs, it is likely that MUPs
directly activate a response in vomeronasal sensory neurons by binding to a subset of these receptors (47). In
sharp contrast to these observation, Holekamp et al.
(113), by using a new microscopical technique that allows
FROM PHEROMONES TO BEHAVIOR
type of receptor only, most likely by a specific V2R, as ESP1
induced c-fos production in vomeronasal neurons expressing V2Rp5, a V2R receptor of the family A, clade III (98).
B. Olfactory Sensory Neurons
1. The cAMP transduction cascade
2. Neuronal subsets
The subset of olfactory sensory neurons expressing
TAARs also express G␣olf, and therefore, it is possible
that they use a signal transduction pathway coupled to
adenylyl cyclase and cAMP (194), although additional
components of the transduction pathway in these neurons are still unknown.
The ion channel TRPM5 is expressed in another subset of olfactory sensory neurons that coexpress the
CNGA2 channel, and some components of the PLC pathway, such as PLC-␤2 and the G protein ␥13 subunit (202).
Physiol Rev • VOL
Since TRPM5 channels are directly activated by intracellular Ca2⫹ (112, 204, 299, 423), it is tempting to speculate
that Ca2⫹ entry through CNG channels could gate TRPM5
channels.
Neurons expressing TRPM5 are located in the ventrolateral zones of the olfactory epithelium and project to
domains of the ventral region of the main olfactory bulb,
an area responsive to urine and other putative pheromones (199 –201, 336, 409). Unfortunately, electrical responses from individual olfactory sensory neurons expressing TRPM5 have not yet been investigated.
As previously mentioned (see sect. IVD), another subset of neurons in the main olfactory epithelium expresses
components typical of a cGMP instead than a cAMP second messenger system, namely, the receptor guanylyl
cyclase type D (GC-D), the cGMP phosphodiesterase
PDE2 (149), and the cyclic nucleotide-gated channel
CNGA3, that was originally discovered in cone photoreceptors (417). These neurons are located in the central
area of the olfactory epithelium and project to an anatomically distinct group of interconnected glomeruli, the
necklace glomeruli, that have been implicated in the suckling response of mammals (149). Leinders-Zufall et al.
(186) showed that neurons expressing GC-D can be activated by the peptide hormones uroguanylin and guanylin
and by natural urine stimuli, while Hu et al. (118) demonstrated that these neurons may be involved in the detection of carbon dioxide. In both studies, GC-D neurons
were shown to use a cGMP signaling cascade (118, 186).
A subset of neurons in the main olfactory epithelium
of humans and mice expresses some vomeronasal receptors of the V1R family (152, 313, 315). Ligands corresponding to aliphatic alcohols or aldehydes have been identified
for human V1Rs (350); however, it is still unresolved
whether neurons expressing these receptors project their
axons to the main or the accessory olfactory bulb.
A small population of neurons in the main olfactory
epithelium expresses the TRPC2 channel. At present,
there are no clues whether these neurons have chemosensory properties, but this finding is certainly relevant
and must be seriously taken into account when comparing the behavioral effects produced by the genetic deletions of the TRPC2 channel with those of the surgical
removal of the vomeronasal organ (P. Mombaert, personal communication).
Early microscopical and immunological studies have
revealed the existence of a novel cell type in the olfactory
epithelium that is morphologically different from chemosensory neurons or supporting cells (44, 148, 260). The
main feature of these cells is represented by the microvilli
that they bear on the mucosal side. It now appears that
olfactory microvillar cells do not represent a homogeneous population of cells but rather they include different
cellular subsets expressing distinct signal transduction
molecules (6, 74, 201). It is very unlikely that these cells
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
In contrast to a common belief, some sensory neurons of the main olfactory epithelium respond to pheromones (364, 395). A large majority of olfactory sensory
neurons express a member of the odorant receptor family
that is coupled to a signal transduction pathway mediated
by cAMP (Fig. 6). The binding of ligands to the odorant
receptor activates the G protein subunit G␣olf, which
stimulates adenylyl cyclase-3 activity producing an increase in cAMP concentration. cAMP causes the opening
of the ciliary cyclic nucleotide-gated channels composed
of three subunits: CNGA2, CNGA4, and CNGB1b (for
reviews, see Refs. 29, 153, 292). These channels allow an
influx of Na⫹ and Ca2⫹ inside the cilia. Ca2⫹-activated Cl⫺
channels are then activated in turn and, due to the unusually high intracellular Cl⫺ concentration, allow an outflow
of Cl⫺, contributing to the inward current (167, 179). As a
result of the odorant binding to receptors, the olfactory
sensory neuron depolarizes. The CNG channel allows
Ca2⫹ entry not only for excitatory but also for inhibitory
purposes. The complex Ca2⫹-calmodulin activates a phosphodiesterase (PDE1C2) that hydrolyzes cAMP and, importantly, produces a negative-feedback effect on the
CNG channel itself, that has been shown to mediate olfactory adaptation (178, 242).
When the depolarization following the activation of
the transduction cascade reaches the threshold for activation of action potentials, they will travel along the axon
to the main olfactory bulb. Differently from vomeronasal
sensory neurons, a maintained current injection will only
elicit a short burst of action potentials in olfactory sensory neurons, instead of a tonic firing (196).
For more details about the cAMP transduction cascade
in olfactory sensory neurons, the reader may consult previous comprehensive reviews (40, 168, 235, 242, 292, 337).
939
940
TIRINDELLI ET AL.
are true neurons as they do not seem to project their long
foot process to the olfactory bulb (74, 201). However,
although their role is enigmatic, a possible chemosensory
function for microvillar cells has been reported (74).
3. Responses of the main olfactory epithelium
to pheromones
C. Grueneberg Ganglion
The transduction pathways in the Grueneberg ganglion
are still largely uncharacterized. Fleischer and colleagues
showed that the majority of neurons express the vomeronasal receptor V2R83 (V2R2 family), together with the G protein G␣i (79, 80), while a distinct subpopulation of G␣iexpressing neurons also expresses TAAR receptors (80).
Another study investigated whether neuronal subsets of the
Grueneberg ganglion express transduction components similar to those of GC-D neurons, based on the observation that
both types of neurons project to the necklace domain of the
main olfactory bulb (Fig. 3). Fleischer et al. (78) reported
that a subset of Grueneberg ganglion neurons expressing the
V2R83 receptor expresses also the GC-D receptor and the
cGMP-stimulated phosphodiesterase PDE2A.
Physiol Rev • VOL
VI. SIGNALING FROM SENSORY EPITHELIA TO
THE BRAIN
A. Main and Accessory Olfactory Bulbs
1. Anatomical organization
The main olfactory bulb is the primary target region of
the olfactory neurons, whereas vomeronasal sensory neurons project to a well-defined dorsal and posterior region of
the olfactory bulb, namely, the accessory olfactory bulb. The
main olfactory bulb exists in all vertebrates, except for
aquatic mammals. In contrast, the accessory olfactory bulb
is absent in adult Old World monkeys, apes, and humans (for
more information, see Table 1 in Ref. 240). In rodents, both
the main and the accessory olfactory bulbs have a largely
similar laminar organization consisting of a superficial nerve
layer formed by the axonal projections of the chemosensory
neurons, the glomerular layer that represents the first-order
synaptic region between sensory neurons and mitral cell
dendrites, and the external and internal plexiform layers that
are regions where soma of, respectively, mitral/tufted and
granule cells reside (for review, see Ref. 240; see also Ref.
182 for suggested changes to the nomenclature of the accessory olfactory bulb). In the main olfactory bulb, glomeruli
are quite well anatomically separated, encapsulated by periglomerular cells, and rather uniform in size (⬃50 ␮m diameter in mice). Conversely, in the accessory olfactory bulb,
these synaptic structures are very diffusely organized, surrounded by a small number of periglomerular cells, and
highly variable in size (10 –30 ␮m diameter in mice). In the
main olfactory bulb, mitral/tufted cells form a distinct monolayer, whereas in the accessory olfactory bulb they are much
less organized. An important difference between mitral/
tufted cells of the two olfactory systems is their arborization:
in the main olfactory bulb these neurons bear only one
apical dendrite, whereas in the accessory olfactory bulb
mitral/tufted cells are equipped with up to six apical dendrites, each entering a different glomerulus (Fig. 7; for review, see Ref. 240).
2. Topographic projections
Axons of all the olfactory sensory neurons expressing a
particular odorant receptor converge to only two glomeruli
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
Wang et al. (395) showed that milk and urinary pheromones that are responsible for puberty delay aggressiveness
in the mouse (279, 281) and evoke an electrical response
from the main olfactory epithelium, as measured by field
potential recordings. Moreover, when the same experimental approach was adapted to mice in which the adenylyl
cyclase type 3 gene had been knocked out, no electrical
response was recorded, indicating that receptors responding
to these compounds are likely to be coupled to the cAMP
signaling pathway. Since behavioral deficits are also evident
in adenylyl cyclase type 3 mutant mice, it can be envisaged
that a cAMP-mediated pheromone signaling may play an
important role in the olfactory epithelium.
MHC peptides can also be detected by sensory neurons of the main olfactory epithelium via the cAMP signaling cascade, as shown by Spehr et al. (364). Two MHC
peptide ligands, SYFPEITHI and AAPDNRETF, elicited an
electrical response, measured by field potential recordings in the main olfactory epithelium at a concentration
near 10⫺11 M (364). Mutant mice lacking the CNGA2 gene
failed to display electrical responses when stimulated
with both peptides, suggesting that a subset of olfactory
neurons expressing CNGA2 are actively involved in the
detection of pheromones besides conventional odorants.
In contrast to that, Lin et al. (200) have demonstrated
that at least two pheromones, 2-heptanone and dimethylpyrazine, stimulate the main olfactory epithelium without the intervention of the CNGA2 channel, suggesting
that pheromone detection in the olfactory epithelium relies on multiple transduction pathways.
Brechbuhl et al. (30) recently uncovered at least one
of the functional roles of the Grueneberg ganglion. Indeed, a subpopulation of neurons was shown to respond
to the yet unidentified “alarm pheromones,” a group of
volatile compounds that are released by animals subjected to stress. Lesions of this organ completely abolish
the so-called freezing reaction, a typical fear response
elicited by alarm pheromones. Curiously, in this study,
lesions of the Grueneberg ganglion did not affect motherpup recognition (30), a behavior that was believed to be
mediated by this organ (118, 173, 317).
FROM PHEROMONES TO BEHAVIOR
941
in the main olfactory bulb (257, 308, 389). This has been
experimentally demonstrated at a single-neuron resolution
by using transgenic mice in which the expression of a given
odorant receptor was linked to that of the green fluorescent
protein, thus allowing the visualization of the axonal netPhysiol Rev • VOL
work (255, 256, 381). In mice there are ⬃2,000 glomeruli, and
their localization is roughly conserved among individuals.
Therefore, the olfactory bulb is topographically organized,
with each glomerulus representing a single type of odorant
receptor (Fig. 7C).
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
FIG. 7. Main and accessory olfactory bulbs. A and B show the general organization of the main olfactory bulb (MOB) and of the accessory
olfactory bulb (AOB) in the adult rat from Nissl-stained specimens. A: each mitral cell (MC) of the main olfactory bulb has a single apical dendrite
projecting to a single glomerulus and several lateral dendrites that contact granule cell dendrites. B: each mitral cell (here indicated as large principal
cell, LPC) of the accessory olfactory bulb has several apical dendrites projecting to multiple glomeruli. EPL, external plexiform layer; MC, mitral
cells; EGC, external granule cell; asterisk, tufted cell axon; IGC, internal granule cell dendrite; S, superficial centrifugal fiber; LOT, lateral olfactory
tract. [From Larriva-Sahd (182).] C and D: schematic topographical representations of sensory projections in the main olfactory (C) and vomeronasal
(D) systems. C: in the main olfactory system, olfactory sensory neurons expressing a given odorant receptor project to the same glomerulus (or
glomerular pair). Three populations of olfactory sensory neurons, each expressing one different type of odorant receptor, are represented in
different colors. D: in the vomeronasal system, sensory neurons that express the same vomeronasal receptor project to multiple small glomeruli.
Six populations of vomeronasal sensory neurons, each expressing one different type of vomeronasal receptor, are represented in different colors.
942
TIRINDELLI ET AL.
Physiol Rev • VOL
through glutamatergic synapses, but the chemical information is further processed by the activity of inhibitory
interneurons as well as periglomerular and granule cells
(141, 210, 339) that contact mitral cells via glutamatergic
dendrodendritic synapses. These consist of an excitatory
glutamatergic input from mitral cells to granule cells,
inducing the release of GABA from granule cells, that in
turn provides inhibition to the mitral cells (141, 182, 303).
For more information, the reader may consult previous
comprehensive reviews (339, 345, 385).
3. Encoding pheromonal information in the main and
accessory olfactory bulbs
To simultaneously examine the responses to pheromones and odorants in the mouse main and accessory
olfactory bulbs, Xu et al. (409) used high-resolution functional magnetic resonance imaging (fMRI), a very powerful technique that can show dynamic responses in both
entire olfactory bulbs to various stimuli in the same animals. Mouse urine activated restricted areas in both olfactory bulbs, eliciting signals mainly in the ventral region of the
main olfactory bulb and in the anterior part of the main
accessory bulb. Since the anterior region of the mouse accessory olfactory bulb receives input from V1R-expressing
vomeronasal sensory neurons, this study is in agreement
with recent results from Holekamp et al. (113) who suggested that urine mainly activates V1Rs (see sect. VA2).
The encoding of pheromonal information in the mitral
cells of the accessory olfactory bulb has been investigated
by recording responses in awake mice using a miniature
motorized drive to move microelectrodes and record the
activity of single neurons (213, 214). Male mice were allowed
to investigate conspecifics differing by sex, genetic makeup,
and hormonal status. Interestingly, an increase in mitral cell
activity occurred after test animals directly contacted and
actively investigated the head and the anogenital regions of
the stimulus animal. The recorded mitral cell responses
were found to be specific for sex and strain of the stimulus
mouse. In addition, mitral cells exhibited both highly specific excitatory and inhibitory responses, these latter likely
originating from the inhibitory interneuron activity. Therefore, the activity of mitral cells appears to be tuned by the
combination of sex and genetic makeup (213, 214), and
therefore, information about sex and strain is already integrated in the accessory olfactory bulb.
Several studies have revealed that the exposure to
urinary components from opposite-sex conspecifics increased the number of c-fos-immunoreactive mitral and
granule and periglomerular cells in the accessory olfactory bulb (116, 126, 228, 295, 410). Volatile urinary odors
from male as well as female mice also stimulate c-fos
expression in distinct clusters of glomeruli in the main
olfactory bulb in both sexes (228, 270).
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
In rodents, V1Rs and V2Rs are individually expressed
in vomeronasal sensory neurons, and each neuron expressing a given vomeronasal receptor (or vomeronasal
receptor pair) sends its axonal projection toward the
anterior (V1Rs) or posterior (V2Rs) accessory olfactory
bulb (100, 143, 414). Gene targeting approaches were used
to visualize the topographical map of each targeted receptor in the mouse accessory olfactory bulb (16, 62, 314).
When a fluorescent marker protein was inserted downstream of a vomeronasal receptor gene, the resulting
visualized axonal projections of neurons expressing that
specific receptor showed that several glomeruli in subregions of the anterior or posterior accessory olfactory bulb
received inputs. A more accurate analysis indicated that
axons from neurons expressing the same receptor coalesce into 10 –30 glomeruli and that a single glomerulus
receives input from neurons expressing different receptor
types (Fig. 7D). Moreover, the capacity to converge and to
form these glomeruli depends on the possibility to express a specific vomeronasal receptor gene, suggesting
that the receptor proteins themselves play a role in axonal
wiring to the bulb (16, 314). In fact, vomeronasal neurons
expressing a nonfunctional (truncated) V1R gene no
longer converge in the accessory olfactory bulb. This
phenotype, also observed for odorant receptors (343),
could be possibly due to the inability of the truncated
gene to prevent the expression of other V1Rs, resulting in
a population of fluorescently tagged neurons randomly
expressing different V1Rs . Indeed, Roppolo et al. (318)
showed that the expression of a nonfunctional V1R allele
allows the coexpression of other V1R genes. Although no
V1R expression has been reported in other subcellular
compartments besides the projecting dendrite, it is possible to envisage that V1Rs may direct axon guidance,
perhaps through a mechanism that involves their expression in the growth cones.
Wagner et al. (393) generated transgenic mouse lines
in which projections from various populations of neurons
expressing distinct V1Rs were differentially labeled with
fluorescent proteins. They found a glomerular map divided in multiple nonoverlapping domains, with each domain clustering glomeruli formed by neurons expressing
V1Rs of the same subfamily, randomly intermingled (393).
Moreover, these authors (393) found that a mitral cell is
not only connected to glomeruli innervated by neurons
expressing the same V1R (62) but also to those associated
with receptors of the same subfamily. Therefore, in the
anterior accessory olfactory bulb, the spatial map originated from vomeronasal organ inputs seems to rely on
receptor subfamilies, rather than individual receptors as
in the main olfactory bulb. This wiring scheme may help
to discriminate molecules with highly similar chemical
structure at different ratios in a pheromonal blend.
In both the main and accessory olfactory systems,
mitral cells are activated by primary sensory neurons
FROM PHEROMONES TO BEHAVIOR
B. Signaling Beyond the Olfactory Bulbs
In the main olfactory bulb, mitral/tufted cells directly
transmit signals from glomeruli to pyramidal neurons in the
olfactory cortex, without a thalamic relay. The primary olfactory cortex, defined as the ensemble of brain regions that
receive direct input from the olfactory bulb, is composed of
several anatomically distinct areas: the piriform cortex, the
olfactory tubercle, the anterior olfactory nucleus, the cortical amygdala, and the entorhinal cortex. In turn, most of
these cortical regions project back to the main olfactory
bulb and to other areas of the brain, including the thalamus,
the hypothalamus, the hippocampus, the frontal cortex, and
the orbitofrontal cortex (Fig. 8; reviewed in Ref. 349). Anatomical studies and genetic approaches indicate that the
topographical organization of odorant information in the
olfactory bulb is not reflected in the olfactory cortex.
Instead, odorant receptors seem to be mapped to multiple, discrete clusters of cortical neurons.
Mitral cells of the accessory olfactory bulb project to
areas of the limbic system: the medial amygdala and the
posteromedial cortical amygdala, which are also called
“vomeronasal amygdala,” the accessory olfactory tract, and
the bed nucleus of the stria terminalis (Fig. 8) (272, 334, 392).
FIG. 8. Central pathways involved in processing chemical information by the vomeronasal and main olfactory systems. [Based on data
from Martinez-Marcos (229) and Yoon et al.
(413).]
Physiol Rev • VOL
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
Lin et al. (199) recorded the electrical activity in the
accessory olfactory bulb in response to urine of different
mouse strains. A subset of mitral cells was found to be
very selective, responding only to male urine. Moreover,
by combining electrophysiology and gas chromatographic
separation of male mouse urine components, these authors identified a male specific compound, methyl-thiomethanethiol (MTMT), that was indeed responsible for
this effect. Interestingly, mitral cells responding to MTMT
were not stimulated by any other urinary compound,
allowing the appellation of “specialist” cells. Furthermore, behavioral experiments also established the biological activity of this molecule, as female mice developed
interest and attraction towards castrated mouse urine
after the addition of synthetic MTMT.
943
944
TIRINDELLI ET AL.
C. Pheromone-Activated Hormonal Systems
Recent experiments have revolutionized the common
beliefs about the control of hormonal responses by pheromones (26, 413). The effects of pheromones on the neuroendocrine status are mediated by a group of neurons in
the hypothalamus, which constitutes the endocrine control center (for reviews, see Refs. 89, 245, 353). These
neurons secrete the GnRH, also known as LH-RH, which
in turn stimulates the anterior pituitary gland to release
two gonadotrophins: the luteinizing hormone (LH) and
the follicle-stimulating hormone (FSH). Both hormones
control the development and the function of the gonads in
males and females, although with different effects.
A genetic approach has been used to visualize the
GnRH neuronal network and, surprisingly, GnRH neurons
were found to be connected with areas relaying information from the main olfactory system, in contrast to the
traditional view that only ascribes connections with
GnRH neurons from the vomeronasal system (26, 413).
Boehm et al. (26) identified neurons that synapse
with GnRH by using a genetic transneuronal tracer, barley
lectin, a glycoprotein that is transferred across synapses.
The barley lectin gene was positioned under the control of
the GnRH promoter so that only GnRH neurons would
express this peptide. In another study, Yoon et al. (413)
used a different experimental approach to anatomically
trace the afferent pathways to GnRH neurons in the hypothalamus. Only neurons expressing GnRH were infected with a modified pseudo-rabies virus and green
fluorescent protein reporter. The genetically controlled
Physiol Rev • VOL
viral tracing allowed the identification of a major projection pathway from the main olfactory system.
Both studies (26, 413) came to the conclusion that
the small pool of GnRH cells (800 neurons) was surprisingly synaptically connected to some tens of thousand
neurons in ⬃50 diverse brain areas, including the main
olfactory ones, suggesting that GnRH neurons may influence a large variety of brain functions.
Interestingly, Yoon et al. (413) reported the absence of synaptic connections between the vomeronasal pathway and GnRH neurons. This could be due to
different experimental approaches used in the two
studies or, perhaps, presynaptic inputs from the vomeronasal pathway identified in one study (26) might derive from a subset of GnRH neurons distinct from those
targeted in the other study (413). However, the observations of Yoon et al. (413) do not refute previous
neuroanatomical tracing data indicating that the vomeronasal pathway projects to the hypothalamic area
where GnRH neurons are located. Indeed, Yoon et al.
(413) confirmed that neurons in this area receive inputs
from the vomeronasal pathway. It could be speculated
that the connections with GnRH are, at least in part,
mediated by nonsynaptic (e.g., hormonal or neuromodulatory) mechanisms.
As a confirmation of these observations, stimulation of
male mice with urinary pheromones induced the activation
of neurons in the anterior cortical amygdala as well as in the
dorsal and ventral endopiriform nucleus. Therefore, signals
in response to pheromonal cues also occur via the main
olfactory epithelium and are transmitted to GnRH neurons
through the olfactory cortex. Moreover, feedback loops
have been identified so that GnRH neurons could influence
both odorant and pheromone signaling.
VII. PHEROMONE-MEDIATED BEHAVIORS
From results presented in the previous sections, it
now appears evident that both the vomeronasal and the
main olfactory systems may be activated by chemicals
that are known to produce pheromonal effects. The
involvement of the vomeronasal and main olfactory
systems on specific behaviors is usually investigated by
inactivating one or both systems. In the past, lesions
were made only by surgical removal of the vomeronasal
organ (VNOx) or by chemical ablation of the main
olfactory epithelium (MOEx) (Table 2). The introduction of a molecular genetic approach allowed the deletion of specific transduction molecules, although it
must be taken into account that the knowledge of genes
involved in transduction in each olfactory organ is still
incomplete, and therefore, results of behavioral experiments from knockout animals must be interpreted very
cautiously.
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
Although vomeronasal sensory neurons expressing
V1Rs or V2Rs, respectively, project to distinct areas of the
accessory olfactory bulb, this segregation, at least in some
species, does not seem to be completely maintained in the
amygdalic regions, where clusters of neurons may combine signal information originating from both families of
vomeronasal receptors (230, 252, 253, 327, 392).
In essence, the majority of these studies provide
evidence that, excluding areas of convergence, the two
vomeronasal pathways project to specific areas of the
amygdala and hypothalamus, thus suggesting a partial
functional separation (252, 253).
In the vomeronasal system, projections to the hypothalamus from the limbic regions are directed to the ventromedial and the medial preoptic areas, which are involved in
reproductive and social behaviors (158, 159, 289).
The medial amygdala receives direct input from the
accessory olfactory bulb, but also from the main olfactory
bulb. However, other rostral basal telencephalic regions
seem to play a relevant role in the integration of signals
originating from both the vomeronasal and the main olfactory epithelia.
945
FROM PHEROMONES TO BEHAVIOR
TABLE
2.
Effects of surgical and genetic ablations of olfactory and vomeronasal systems
VNX
Hamster
Rat
VGENX
OEX
2Scent marking (290)
2Male copulatory
behavior (55, 76, 298)
2Lordosis behavior (221)
2Individual recognition
(365)
2Partner preference (10)
2Scent marking (145)
2Sexual arousal (146)
2Individual recognition (146)
2Male copulatory behavior (146)
2Penile erection (172)
Cat
2Lordosis behavior (154)
2Sex discrimination (154)
(73, 376)
2Copulatory behavior (ac3) (395)
2Intermale aggressiveness
(ac-3) (395)
2Individual recognition
(cnga2) (364)
2Suckling behavior
1Maternal behavior (52)
2Catnip reaction (104)
Reference numbers are given in parentheses.
Here, we will discuss experimental results mainly
related to sexual and aggressive behaviors and will try
to interpret to what extent the vomeronasal and/or the
main olfactory systems appear to be involved. It is very
important to note that behavioral expressions largely
vary among species; moreover, responses to pheromones cannot be simply described as innate or stereotyped, but may also depend on contextual situations
and can be modified by some forms of learning and
memory.
Physiol Rev • VOL
A. Male Behaviors
1. Male sexual behavior
Sexual behavior has been intensively approached by
many studies in the last 20 years, particularly in rodents. In
male mice, the typical mating sequence consists of olfactory
exploration of estrus females followed by multiple episodes
of ultrasound vocalization, mounting and intromission, and
eventually ejaculation. All these parameters can be evaluated in a typical test experiment (120, 121).
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
2Stress-induced
hyperthermia (165)
Maternal aggression (170)
2Male induced sexual
receptivity (302)
2Nursing behavior (36)
2Infanticide (243)
2Sexual arousal (325)
2Lordosis behavior (326)
2Lordosis behavior to
male mount (155)
1Sperm motility (175)
2Sex discrimination (trpc2) (193,
284, 368)
2Intermale aggressiveness (trpc2;
g␣i2) (278, 368)
2Urine marking (181, 233) 2Maternal aggression (trpc2; v1rav1rb; g␣i2) (61, 161, 193, 278)
2Intermale aggressiveness 2Lactating behavior (trpc2) (161)
(54, 233)
2Marking
Male-like sexual behavior in females
(trpc2) (161)
2Copulatory behavior (54) 2Copulatory behavior (v1ra-v1rb)
(61)
1Puberty onset (209)
2⌴arking (trpc2)(193)
2Strain recognition (206)
2Pregnancy block (156,
206, 216)
2Sexual investigation
(135)
Male-like sexual behavior
in females (161)
Ferret
2Sexual investigation
(404)
Lemur
2Copulatory behavior (8)
2Intermale aggressiveness
(8)
Opossum
2Male-mediated induction
of estrus (131)
Prairie vole 2Intermale aggressiveness
(399)
2Pairing (189)
Rabbit
1Maternal behavior (87)
Mouse
OGENX
946
TIRINDELLI ET AL.
Physiol Rev • VOL
cation as scent or flank marking are likely to be primarily
mediated by the main olfactory system (145).
2. Male aggressiveness
In mice, male to male territorial aggression is a remarkable innate social behavior that is triggered by compounds present in urine (see sect. II, Table 1). In a common assay employed to test male aggressiveness, a male
intruder is introduced in a cage containing a singly housed
male resident. After a period of olfactory investigation, to
allow individual recognition, the resident mouse initiates
a series of attacks against the intruder. The number and
duration of the attacks and the latency to the first attack
represent standard parameters to evaluate the aggressive
behavior (58, 233).
The surgical ablation of the vomeronasal organ has
been long known to nearly abolish aggressiveness in male
mice and prairie voles (12, 54, 55, 233, 399, 407), an effect
that is recapitulated in mutant mice bearing the deletion
of the gene encoding TRPC2 (161, 193, 368). Indeed, male
mice bearing the genetic ablation of the TRPC2 channel
fail to display aggression toward male intruders (193,
368), a behavioral effect that, in an attenuated form, is
also evident in mice lacking G␣i2 (278).
The main olfactory system also plays a very important role in male aggressiveness. Indeed, mice that either
have undergone the chemical ablation of the olfactory
epithelium (via nasal irrigation with zinc sulfate) (154) or
have been genetically mutated in the genes encoding the
odorant transduction molecules adenylyl cyclase type 3
and CNGA2 (227, 395) show abnormal sexual and aggressive behaviors (227).
Thus both the olfactory and the vomeronasal systems
appear to be necessary for eliciting male aggressiveness.
B. Female Behaviors
1. Female sexual behavior
In females of different species, sexual behavior is
characterized by ultrasonic vocalization and lordosis elicited by lumbosacral tactile stimulation (174, 294).
In mice, the genetic ablation of the TRPC2 channel
did not alter female sexual receptivity to males but, surprisingly, caused a female sexual behavior similar to that
of males, i.e., females attempted to copulate as males
(161). Kimchi et al. (161) also showed that similar sexual
behaviors were displayed by females in which the vomeronasal organ was surgically removed, in contrast to a
previous study that described a reduced sexual receptivity in VNOx females (155). The chemical ablation of the
main olfactory epithelium (with zinc sulfate) reduces sexual behavior in female mice (154).
In female rats, both chemical ablation of the main
olfactory epithelium and vomeronasal organ removal af-
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
The genetic ablation of the TRPC2 channel greatly
impairs some features of the sexual behavior. Indeed,
although male mice mate normally with females, they also
show an indiscriminate courtship and mounting behavior
toward females (161, 193, 368). A similar situation was
observed in VNOx male mice (161). Furthermore, individual recognition, through the discrimination of nonvolatile
odorants, is also affected in VNOx mice (155, 284). On the
other hand, the simultaneous deletion of a group of V1R
receptors results in a significant decrease of male copulatory behavior (61), although this latter appears normal
in G␣i2 mutants, suggesting that apical neurons are likely
not to be of relevant importance for mating motivation. All
these findings reveal that the mouse vomeronasal organ is
indeed involved in individual recognition and reproductive
activity, but that signaling through the vomeronasal organ is
not required for initiating sexual behavior.
The chemical ablation of the main olfactory system
(via nasal irrigation with zinc sulfate) produces a striking reduction in male reproductive behavior (227). Furthermore, the genetic deletion of the CNGA2 channel
(227), or of adenylyl cyclase type 3 (395), entirely abolishes male mating performances producing severe deficiencies in each step involved in mating: olfactory
exploration, mounting, and intromission, compared
with normal mice. Thus the main olfactory system
seems to play an essential role for male mouse mating
behavior.
In rats, individual recognition and sexual arousal is
only temporary or moderately depressed by lesions of the
vomeronasal organ (22, 325) as well as the overall copulatory behaviors appear to be slightly or partially affected
(171, 324, 325).
Sexual behavior in male hamsters is totally abolished
by bilateral removal of the olfactory bulbs, an operation
that eliminates sensory input from both the olfactory and
the vomeronasal systems (298). The destruction of the
main olfactory epithelium did not impair male hamster
mating behavior, whereas the peripheral deafferentation
of the vomeronasal system produces severe sexual behavior deficits in approximately one-third of the treated animals. Combined deafferentation of both the vomeronasal
and the olfactory systems completely abolishes copulation in all experimental animals (298). Mating behavior
alterations appear after removal of the vomeronasal organ in sexually inexperienced males, and this effect is
reversed by intracerebroventricular injections of GnRH
(76, 247), indicating that GnRH acts as an intermediate in
the facilitation of mating behavior by vomeronasal sensory input. However, sexual experience prior to the removal of the vomeronasal organ mitigates the effect of the
lesion on sexual behavior, indicating that behavioral responses are influenced both by VNOx and by sexual experience (291). In this species, other ways of communi-
FROM PHEROMONES TO BEHAVIOR
fect sexual receptivity and the male-induced release of
GnRH (17, 302).
In female hamsters, the removal of the vomeronasal
organ disrupts the elicitation of lordosis by lumbosacral
tactile stimulation, an effect reversed by injection of
GnRH (221).
2. Maternal aggressive behavior
VIII. CONCLUSIONS AND
FUTURE CHALLENGES
Although we are only beginning to understand the
overlapping role of the main olfactory and vomeronasal
systems for pheromone sensing, there is now overwhelming evidence that the main olfactory system is involved in
a variety of important physiological mechanisms underlying pheromonal communication.
In the early 1970s, different studies developed a
model that led to the so-called dual olfactory hypothesis,
sustained by the observation that efferent and afferent
connections of the main and accessory olfactory bulb
project in nonoverlapping areas of the brain, according to
which the olfactory and vomeronasal systems were organized as parallel anatomical axes subserving different
functions. Recent anatomical and functional studies have
indicated that this distinction is not as strict as commonly
thought (229). There seems to be, in fact, areas of large
convergence in the anterior medial amygdala. Moreover,
in the rostral basal telencephalon, both inputs converged
in areas that were classically categorized as exclusively
vomeronasal or olfactory. As for the anatomical observation, behavioral and biochemical studies seem to point to
a complementary role of the two systems. For example,
the functional ablation of the olfactory and vomeronasal
organs leads to similar behavioral changes. Chemosensory neurons sensitive to pheromone have been identified
in both subsystems. Not last, humans as well as other
species in which the vomeronasal system is absent can
probably detect pheromones via the main olfactory epithelium. It is therefore possible that the two systems may
Physiol Rev • VOL
have developed independently for the detection of pheromones, and only the main olfactory epithelium has further specialized also for the perception of conventional
odorants. Proof of this hypothesis will require additional
studies especially in species in which olfaction is not a
primary sensory process as humans, for example; however, it is interesting to note that part of the main olfactory central pathways seem to converge in regions that
control the sexual status of the individual.
ACKNOWLEDGMENTS
We thank Didier Trotier for kindly providing the original
micrograph for figure 6A.
Address for reprint requests and other correspondence: A.
Menini, Scuola Internazionale Superiore di Studi Avanzati, Via
Beirut 2, 34014 Trieste, Italy (e-mail: [email protected]).
GRANTS
This work was supported by grants from the Italian Ministry of Research and by the Italian Institute of Technology.
REFERENCES
1. Abel EL. Alarm substance emitted by rats in the forced-swim test
is a low volatile pheromone. Physiol Behav 50: 723–727, 1991.
2. Achiraman S, Archunan G. Characterization of urinary volatiles
in Swiss male mice (Mus musculus): bioassay of identified compounds. J Biosci 27: 679 – 686, 2002.
3. Alioto TS, Ngai J. The odorant receptor repertoire of teleost fish.
BMC Genomics 6: 173, 2005.
4. Amrein H. Pheromone perception and behavior in Drosophila.
Curr Opin Neurobiol 14: 435– 442, 2004.
5. Apfelbach R, Blanchard CD, Blanchard RJ, Hayes RA, McGregor
IS. The effects of predator odors in mammalian prey species: a review
of field and laboratory studies. Neurosci Biobehav Rev 29: 1123–1144,
2005.
6. Asan E, Drenckhahn D. Immunocytochemical characterization of
two types of microvillar cells in rodent olfactory epithelium. Histochem Cell Biol 123: 157–168, 2005.
7. Astic L, Pellier-Monnin V, Godinot F. Spatio-temporal patterns
of ensheathing cell differentiation in the rat olfactory system during
development. Neuroscience 84: 295–307, 1998.
8. Aujard F. Effect of vomeronasal organ removal on male sociosexual responses to female in a prosimian primate (Microcebus
murinus). Physiol Behav 62: 1003–1008, 1997.
9. Bacchini A, Gaetani E, Cavaggioni A. Pheromone binding proteins of the mouse, Mus musculus. Experientia 48: 419 – 421, 1992.
10. Ballard CL, Wood RI. Partner preference in male hamsters: steroids, sexual experience and chemosensory cues. Physiol Behav
91: 1– 8, 2007.
11. Barber PC, Raisman G. Replacement of receptor neurones after
section of the vomeronasal nerves in the adult mouse. Brain Res
147: 297–313, 1978.
12. Bean NJ. Modulation of agonistic behavior by the dual olfactory
system in male mice. Physiol Behav 29: 433– 437, 1982.
13. Bean NJ, Wysocki CJ. Vomeronasal organ removal and female
mouse aggression: the role of experience. Physiol Behav 45: 875–
882, 1989.
14. Beier K, Ginez I, Schaller H. Localization of steroid hormone
receptors in the apocrine sweat glands of the human axilla. Histochem Cell Biol 123: 61– 65, 2005.
15. Bellringer JF, Pratt HP, Keverne EB. Involvement of the vomeronasal organ and prolactin in pheromonal induction of delayed
implantation in mice. J Reprod Fertil 59: 223–228, 1980.
16. Belluscio L, Koentges G, Axel R, Dulac C. A map of pheromone
receptor activation in the mammalian brain. Cell 97: 209 –220, 1999.
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
Female mice are normally not aggressive towards
intruders, but during lactation, they violently attack nonresident males. This type of maternal aggressive behavior
has been shown to be dependent on chemosensory cues
(13). The surgical removal of the vomeronasal organ affects aggression of lactating females (13, 161), whereas it
has little or no effect on other maternal behavior parameters as nest building or pup retrieval (13). The genetic
ablation of the TRPC2 channel (161, 193), the G␣i2 subunit
(278), or a subfamily of V1R receptors (61), also results in
a low expression of maternal aggression, suggesting that
this type of behavior is possibly linked to the apical subset
of chemosensory neurons.
947
948
TIRINDELLI ET AL.
Physiol Rev • VOL
42. Bunzow JR, Sonders MS, Arttamangkul S, Harrison LM,
Zhang G, Quigley DI, Darland T, Suchland KL, Pasumamula S,
Kennedy JL, Olson SB, Magenis RE, Amara SG, Grandy DK.
Amphetamine, 3,4-methylenedioxymethamphetamine, lysergic acid
diethylamide, and metabolites of the catecholamine neurotransmitters are agonists of a rat trace amine receptor. Mol Pharmacol 60:
1181–1188, 2001.
43. Butenandt A, Beckmann R, Stamm D, Hecker E. über den
Sexuallockstoff des Seidenspinners Bombyx mori. Reindarstellung
und Konstitution. Naturforsch 14: 283–284, 1959.
44. Carr VM, Farbman AI, Colletti LM, Morgan JI. Identification of
a new non-neuronal cell type in rat olfactory epithelium. Neuroscience 45: 433– 449, 1991.
45. Catania KC. Evolution of sensory specializations in insectivores.
Anat Rec A Discov Mol Cell Evol Biol 287: 1038 –1050, 2005.
46. Cavaggioni A, Mucignat-Caretta C. Major urinary proteins,
alpha(2U)-globulins and aphrodisin. Biochim Biophys Acta 1482:
218 –228, 2000.
47. Chamero P, Marton TF, Logan DW, Flanagan K, Cruz JR,
Saghatelian A, Cravatt BF, Stowers L. Identification of protein
pheromones that promote aggressive behaviour. Nature 450: 899 –
902, 2007.
48. Cheetham SA, Smith AL, Armstrong SD, Beynon RJ, Hurst
JL. Limited variation in the major urinary proteins of laboratory
mice. Physiol Behav 96: 253–261, 2009.
49. Cheetham SA, Thom MD, Jury F, Ollier WE, Beynon RJ, Hurst
JL. The genetic basis of individual-recognition signals in the
mouse. Curr Biol 17: 1771–1777, 2007.
50. Cherry JA, Pho V. Characterization of cAMP degradation by
phosphodiesterases in the accessory olfactory system. Chem
Senses 27: 643– 652, 2002.
51. Chess A, Simon I, Cedar H, Axel R. Allelic inactivation regulates
olfactory receptor gene expression. Cell 78: 823– 834, 1994.
52. Chirino R, Beyer C, Gonzalez-Mariscal G. Lesion of the main
olfactory epithelium facilitates maternal behavior in virgin rabbits.
Behav Brain Res 180: 127–132, 2007.
52a.Chromek M, Slamová Z, Bergman P, Kovács L, Podracká L,
Ehrén I, Hökfelt T, Gudmundsson GH, Gallo RL, Agerberth B,
Brauner A. The antimicrobial peptide cathelicidin protects the
urinary tract against invasive bacterial infection. Nat Med 12: 636 –
641, 2006.
53. Cinelli AR, Wang D, Chen P, Liu W, Halpern M. Calcium transients in the garter snake vomeronasal organ. J Neurophysiol 87:
1449 –1472, 2002.
54. Clancy AN, Coquelin A, Macrides F, Gorski RA, Noble EP.
Sexual behavior and aggression in male mice: involvement of the
vomeronasal system. J Neurosci 4: 2222–2229, 1984.
55. Clancy AN, Macrides F, Singer AG, Agosta WC. Male hamster
copulatory responses to a high molecular weight fraction of vaginal
discharge: effects of vomeronasal organ removal. Physiol Behav
33: 653– 660, 1984.
56. Clapham DE, Neer EJ. G protein beta gamma subunits. Annu Rev
Pharmacol Toxicol 37: 167–203, 1997.
57. Cocke R, Moynihan JA, Cohen N, Grota LJ, Ader R. Exposure
to conspecific alarm chemosignals alters immune responses in
BALB/c mice. Brain Behav Immun 7: 36 – 46, 1993.
58. Connor JL, Lynds PG. Mouse aggression and the intruder-familiarity effect: evidence for multiple-factor determination c57bl.
J Comp Physiol Psychol 91: 270 –280, 1977.
59. Conzelmann S, Malun D, Breer H, Strotmann J. Brain targeting
and glomerulus formation of two olfactory neuron populations
expressing related receptor types. Eur J Neurosci 14: 1623–1632,
2001.
60. Crish SD, Rice FL, Park TJ, Comer CM. Somatosensory organization and behavior in naked mole-rats I: vibrissa-like body hairs
comprise a sensory array that mediates orientation to tactile stimuli. Brain Behav Evol 62: 141–151, 2003.
61. Del Punta K, Leinders-Zufall T, Rodriguez I, Jukam D,
Wysocki CJ, Ogawa S, Zufall F, Mombaerts P. Deficient pheromone responses in mice lacking a cluster of vomeronasal receptor
genes. Nature 419: 70 –74, 2002.
62. Del Punta K, Puche A, Adams NC, Rodriguez I, Mombaerts P.
A divergent pattern of sensory axonal projections is rendered
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
17. Beltramino C, Taleisnik S. Release of LH in the female rat by
olfactory stimuli. Effect of the removal of the vomeronasal organs
or lesioning of the accessory olfactory bulbs. Neuroendocrinology
36: 53–58, 1983.
18. Benton D, Wastell V. Effects of androstenol on human sexual
arousal. Biol Psychol 22: 141–147, 1986.
19. Berghard A, Buck LB. Sensory transduction in vomeronasal neurons: evidence for G alpha o, G alpha i2, and adenylyl cyclase II as
major components of a pheromone signaling cascade. J Neurosci
16: 909 –918, 1996.
20. Berglund H, Lindstrom P, Savic I. Brain response to putative
pheromones in lesbian women. Proc Natl Acad Sci USA 103: 8269 –
8274, 2006.
21. Beynon RJ, Hurst JL. Urinary proteins and the modulation of
chemical scents in mice and rats. Peptides 25: 1553–1563, 2004.
22. Bluthe RM, Dantzer R. Role of the vomeronasal system in vasopressinergic modulation of social recognition in rats. Brain Res
604: 205–210, 1993.
23. Bocskei Z, Groom CR, Flower DR, Wright CE, Phillips SE,
Cavaggioni A, Findlay JB, North AC. Pheromone binding to two
rodent urinary proteins revealed by X-ray crystallography. Nature
360: 186 –188, 1992.
24. Boehm N, Gasser B. Sensory receptor-like cells in the human
foetal vomeronasal organ. Neuroreport 4: 867– 870, 1993.
25. Boehm N, Roos J, Gasser B. Luteinizing hormone-releasing hormone (LHRH)-expressing cells in the nasal septum of human fetuses. Brain Res 82: 175–180, 1994.
26. Boehm U, Zou Z, Buck LB. Feedback loops link odor and pheromone signaling with reproduction. Cell 123: 683– 695, 2005.
27. Borowsky B, Adham N, Jones KA, Raddatz R, Artymyshyn R,
Ogozalek KL, Durkin MM, Lakhlani PP, Bonini JA, Pathirana
S, Boyle N, Pu X, Kouranova E, Lichtblau H, Ochoa FY,
Branchek TA, Gerald C. Trace amines: identification of a family
of mammalian G protein-coupled receptors. Proc Natl Acad Sci
USA 98: 8966 – 8971, 2001.
28. Boschat C, Pelofi C, Randin O, Roppolo D, Luscher C, Broillet
MC, Rodriguez I. Pheromone detection mediated by a V1r vomeronasal receptor. Nat Neurosci 5: 1261–1262, 2002.
29. Bradley J, Reisert J, Frings S. Regulation of cyclic nucleotidegated channels. Curr Opin Neurobiol 15: 343–349, 2005.
30. Brechbuhl J, Klaey M, Broillet MC. Grueneberg ganglion cells
mediate alarm pheromone detection in mice. Science 321: 1092–
1095, 2008.
31. Breer H, Fleischer J, Strotmann J. The sense of smell: multiple
olfactory subsystems. Cell Mol Life Sci 63: 1465–1475, 2006.
32. Brennan PA, Schellinck HM, Keverne EB. Patterns of expression of the immediate-early gene egr-1 in the accessory olfactory
bulb of female mice exposed to pheromonal constituents of male
urine. Neuroscience 90: 1463–1470, 1999.
33. Briand L, Blon F, Trotier D, Pernollet JC. Natural ligands of
hamster aphrodisin. Chem Senses 29: 425– 430, 2004.
34. Briand L, Trotier D, Pernollet JC. Aphrodisin, an aphrodisiac
lipocalin secreted in hamster vaginal secretions. Peptides 25: 1545–
1552, 2004.
35. Broman I. Uber die Entwickelung der konstanten grosseren Nasenhohlendr usen der Nagetiere. Zeitschriftf ur Anatomie und Entwicklungs-Geschichte 60: 439 –586, 1921.
36. Brouette-Lahlou I, Godinot F, Vernet-Maury E. The mother
rat’s vomeronasal organ is involved in detection of dodecyl propionate, the pup’s preputial gland pheromone. Physiol Behav 66:
427– 436, 1999.
37. Bruce HM. An exteroceptive block to pregnancy in the mouse.
Nature 184: 105, 1959.
38. Bruce HM. Pheromones and behavior in mice. Acta Neurol Psychiatr Belg 69: 529 –538, 1969.
39. Buck L, Axel R. A novel multigene family may encode odorant
receptors: a molecular basis for odor recognition. Cell 65: 175–187,
1991.
40. Buck LB. The molecular architecture of odor and pheromone
sensing in mammals. Cell 100: 611– 618, 2000.
41. Buiakova OI, Krishna NS, Getchell TV, Margolis FL. Human
and rodent OMP genes: conservation of structural and regulatory
motifs and cellular localization. Genomics 20: 452– 462, 1994.
FROM PHEROMONES TO BEHAVIOR
63.
64.
65.
66.
67.
70.
71.
72.
73.
74.
75.
76.
77.
78.
79.
80.
81.
82.
83.
84.
85.
86.
87.
88.
Physiol Rev • VOL
89. Gore AC. GnRH: The Master Molecule of Reproduction. Norwell,
MA: Kluwer Academic, 2002.
90. Graziadei PP, Graziadei GA. Neurogenesis and neuron regeneration in the olfactory system of mammals. I. Morphological aspects
of differentiation and structural organization of the olfactory sensory neurons. J Neurocytol 8: 1–18, 1979.
91. Graziadei PP, Monti A, Graziadei G. Neurogenesis and neuron
regeneration in the olfactory system of mammals. III. Deafferentation and reinnervation of the olfactory bulb following section of the
fila olfactoria in rat. J Neurocytol 9: 145–162, 1980.
92. Greenwood DR, Comeskey D, Hunt MB, Rasmussen LE. Chemical communication: chirality in elephant pheromones. Nature 438:
1097–1098, 2005.
93. Grimsby J, Toth M, Chen K, Kumazawa T, Klaidman L, Adams
JD, Karoum F, Gal J, Shih JC. Increased stress response and
beta-phenylethylamine in MAOB-deficient mice. Nat Genet 17: 206 –
210, 1997.
93a.Grüneberg H. A ganglion probably belonging to the N. terminals
system in the nasal mucosa of the mouse. Z Anat Entwicklungsgesch 140: 39 –52, 1973.
94. Grus WE, Shi P, Zhang J. Largest vertebrate vomeronasal type 1
receptor gene repertoire in the semiaquatic platypus. Mol Biol Evol
24: 2153–2157, 2007.
95. Guo J, Zhou A, Moss RL. Urine and urine-derived compounds
induce c-fos mRNA expression in accessory olfactory bulb. Neuroreport 8: 1679 –1683, 1997.
96. Gustavson AR, Dawson ME, Bonett DG. Androstenol, a putative
human pheromone, affects human (Homo sapiens) male choice
performance. J Comp Psychol 101: 210 –212, 1987.
97. Gutierrez-Garcia AG, Contreras CM, Mendoza-Lopez MR,
Garcia-Barradas O, Cruz-Sanchez JS. Urine from stressed rats
increases immobility in receptor rats forced to swim: role of 2-heptanone. Physiol Behav 91: 166 –172, 2007.
98. Haga S, Yoshihara Y, Touhara K. Male-specific exocrine peptide
ESP1 activates a selective neural pathway via V2Rp5 receptor in
the mouse vomeronasal system. Chem Senses 33: S90, 2008.
99. Hagino-Yamagishi K, Matsuoka M, Ichikawa M, Wakabayashi
Y, Mori Y, Yazaki K. The mouse putative pheromone receptor was
specifically activated by stimulation with male mouse urine. J Biochem 129: 509 –512, 2001.
100. Halem HA, Baum MJ, Cherry JA. Sex difference and steroid
modulation of pheromone-induced immediate early genes in the
two zones of the mouse accessory olfactory system. J Neurosci
21: 2474 –2480, 2001.
101. Halpern M, Martinez-Marcos A. Structure and function of the
vomeronasal system: an update. Prog Neurobiol 70: 245–318, 2003.
102. Halpern M, Shapiro LS, Jia C. Differential localization of G
proteins in the opossum vomeronasal system. Brain Res 677: 157–
161, 1995.
103. Hardie RC. Vision: dynamic platforms. Nature 450: 37–39, 2007.
104. Hart BL, Leedy MG. Analysis of the catnip reaction: mediation by
olfactory system, not vomeronasal organ. Behav Neural Biol 44:
38 – 46, 1985.
105. Hashiguchi Y, Nishida M. Evolution of trace amine associated
receptor (TAAR) gene family in vertebrates: lineage-specific expansions and degradations of a second class of vertebrate chemosensory receptors expressed in the olfactory epithelium. Mol Biol Evol
24: 2099 –2107, 2007.
106. He J, Ma L, Kim S, Nakai J, Yu CR. Encoding gender and
individual information in the mouse vomeronasal organ. Science
320: 535–538, 2008.
107. Hebb AL, Zacharko RM, Dominguez H, Trudel F, Laforest S,
Drolet G. Odor-induced variation in anxiety-like behavior in mice
is associated with discrete and differential effects on mesocorticolimbic cholecystokinin mRNA expression. Neuropsychopharmacology 27: 744 –755, 2002.
108. Hebb AL, Zacharko RM, Gauthier M, Trudel F, Laforest S,
Drolet G. Brief exposure to predator odor and resultant anxiety
enhances mesocorticolimbic activity and enkephalin expression in
CD-1 mice. Eur J Neurosci 20: 2415–2429, 2004.
109. Hedgewig R. [Comparative anatomical studies on the Jacobson
organs of Nycticebus coucang Boddaert, 1785 (Prosimiae, Lorisidae) and Galago crassicaudatus E. Geoffroy, 1812 (Prosimiae,
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
68.
convergent by second-order neurons in the accessory olfactory
bulb. Neuron 35: 1057–1066, 2002.
Dennis JC, Allgier JG, Desouza LS, Eward WC, Morrison EE.
Immunohistochemistry of the canine vomeronasal organ. J Anat
203: 329 –338, 2003.
Dibattista M, Mazzatenta A, Grassi F, Tirindelli R, Menini A.
Hyperpolarization-activated cyclic nucleotide-gated channels in
mouse vomeronasal sensory neurons. J Neurophysiol 100: 576 –
586, 2008.
Dorries KM, Adkins-Regan E, Halpern BP. Olfactory sensitivity
to the pheromone, androstenone, is sexually dimorphic in the pig.
Physiol Behav 57: 255–259, 1995.
Doty R. Handbook of Olfaction and Gustation. New York: Dekker,
2003.
Doty RL, Dunbar I. Attraction of beagles to conspecific urine,
vaginal and anal sac secretion odors. Physiol Behav 12: 825– 833,
1974.
Doving KB, Trotier D. Structure and function of the vomeronasal
organ. J Exp Biol 201: 2913–2925, 1998.
Duchamp-Viret P, Chaput MA, Duchamp A. Odor response
properties of rat olfactory receptor neurons. Science 284: 2171–
2174, 1999.
Dulac C, Axel R. A novel family of genes encoding putative
pheromone receptors in mammals. Cell 83: 195–206, 1995.
Eccles R. Autonomic innervation of the vomeronasal organ of the
cat. Physiol Behav 28: 1011–1015, 1982.
Edwards DA, Burge KG. Olfactory control of the sexual behavior
of male and female mice. Physiol Behav 11: 867– 872, 1973.
Elsaesser R, Montani G, Tirindelli R, Paysan J. Phosphatidylinositide signalling proteins in a novel class of sensory cells in the
mammalian olfactory epithelium. Eur J Neurosci 21: 2692–2700,
2005.
Epple G. Communication by chemical signals. In: Comparative
Primate Biology: Behavior, Conservation and Ecology, edited by
Mitchell G and Erwin J. New York: Liss, 1986, vol. 2A, p. 531–580.
Fernandez-Fewell GD, Meredith M. Facilitation of mating behavior in male hamsters by LHRH and AcLHRH5–10: interaction
with the vomeronasal system. Physiol Behav 57: 213–221, 1995.
Firestein S, Picco C, Menini A. The relation between stimulus
and response in olfactory receptor cells of the tiger salamander.
J Physiol 468: 1–10, 1993.
Fleischer J, Mamasuew K, Breer H. Expression of cGMP signaling elements in the Grueneberg ganglion. Histochem Cell Biol 131:
75– 88, 2009.
Fleischer J, Schwarzenbacher K, Besser S, Hass N, Breer H.
Olfactory receptors and signalling elements in the Grueneberg
ganglion. J Neurochem 98: 543–554, 2006.
Fleischer J, Schwarzenbacher K, Breer H. Expression of trace
amine-associated receptors in the Grueneberg ganglion. Chem
Senses 32: 623– 631, 2007.
Fulle HJ, Vassar R, Foster DC, Yang RB, Axel R, Garbers DL.
A receptor guanylyl cyclase expressed specifically in olfactory
sensory neurons. Proc Natl Acad Sci USA 92: 3571–3575, 1995.
Fuss SH, Omura M, Mombaerts P. The Grueneberg ganglion of
the mouse projects axons to glomeruli in the olfactory bulb. Eur
J Neurosci 22: 2649 –2654, 2005.
Ganzhorn JU, Kappeler PM. [Lemurs of Madagascar. Tests on
evolution of primate communities.] Naturwissenschaften 80: 195–
208, 1993.
Gavaghan McKee CL, Wilson ID, Nicholson JK. Metabolic phenotyping of nude and normal (Alpk:ApfCD, C57BL10J) mice. J
Proteome Res 5: 378 –384, 2006.
Giacobini P, Benedetto A, Tirindelli R, Fasolo A. Proliferation
and migration of receptor neurons in the vomeronasal organ of the
adult mouse. Brain Res 123: 33– 40, 2000.
Godfrey PA, Malnic B, Buck LB. The mouse olfactory receptor
gene family. Proc Natl Acad Sci USA 101: 2156 –2161, 2004.
Gonzalez-Mariscal G, Chirino R, Beyer C, Rosenblatt JS. Removal of the accessory olfactory bulbs promotes maternal behavior in virgin rabbits. Behav Brain Res 152: 89 –95, 2004.
Goodwin M, Gooding KM, Regnier F. Sex pheromone in the dog.
Science 203: 559 –561, 1979.
949
950
110.
111.
112.
113.
114.
116.
117.
118.
119.
120.
121.
122.
123.
124.
125.
126.
127.
128.
130.
131.
132.
133.
Lorisidae). I. Nycticebus coucang.] Gegenbaurs Morphol Jahrb 126:
543–593, 1980.
Herrada G, Dulac C. A novel family of putative pheromone receptors in mammals with a topographically organized and sexually
dimorphic distribution. Cell 90: 763–773, 1997.
Hoffmann KH, Dettner K, Tomaschko KH. Chemical signals in
insects and other arthropods: from molecular structure to physiological functions. Physiol Biochem Zool 79: 344 –356, 2006.
Hofmann T, Chubanov V, Gudermann T, Montell C. TRPM5 is
a voltage-modulated and Ca2⫹-activated monovalent selective cation channel. Curr Biol 13: 1153–1158, 2003.
Holekamp TF, Turaga D, Holy TE. Fast three-dimensional fluorescence imaging of activity in neural populations by objectivecoupled planar illumination microscopy. Neuron 57: 661– 672, 2008.
Holst DV, Eichmann F. Sex-specific regulation of marking behavior by sex hormones and conspecifics scent in tree shrews (Tupaia
belangeri). Physiol Behav 63: 157–164, 1998.
Holy TE, Dulac C, Meister M. Responses of vomeronasal neurons to natural stimuli. Science 289: 1569 –1572, 2000.
Honda N, Sakamoto H, Inamura K, Kashiwayanagi M. Changes
in Fos expression in the accessory olfactory bulb of sexually
experienced male rats after exposure to female urinary pheromones. Eur J Neurosci 27: 1980 –1988, 2008.
Horowitz LF, Montmayeur JP, Echelard Y, Buck LB. A genetic
approach to trace neural circuits. Proc Natl Acad Sci USA 96:
3194 –3199, 1999.
Hu J, Zhong C, Ding C, Chi Q, Walz A, Mombaerts P, Matsunami H, Luo M. Detection of near-atmospheric concentrations of
CO2 by an olfactory subsystem in the mouse. Science 317: 953–957,
2007.
Hudson R, Distel H. Pheromonal release of suckling in rabbits
does not depend on the vomeronasal organ. Physiol Behav 37:
123–128, 1986.
Hull EM, Dominguez JM. Getting his act together: roles of glutamate, nitric oxide, and dopamine in the medial preoptic area.
Brain Res 1126: 66 –75, 2006.
Hull EM, Dominguez JM. Sexual behavior in male rodents. Horm
Behav 52: 45–55, 2007.
Hurst JL, Beynon RJ. Scent wars: the chemobiology of competitive signalling in mice. Bioessays 26: 1288 –1298, 2004.
Hurst JL, Payne CE, Nevison CM, Marie AD, Humphries RE,
Robertson DH, Cavaggioni A, Beynon RJ. Individual recognition in mice mediated by major urinary proteins. Nature 414:
631– 634, 2001.
Inamura K, Kashiwayanagi M. Inward current responses to urinary substances in rat vomeronasal sensory neurons. Eur J Neurosci 12: 3529 –3536, 2000.
Inamura K, Kashiwayanagi M, Kurihara K. Inositol-1,4,5trisphosphate induces responses in receptor neurons in rat vomeronasal sensory slices. Chem Senses 22: 93–103, 1997.
Inamura K, Kashiwayanagi M, Kurihara K. Regionalization of
Fos immunostaining in rat accessory olfactory bulb when the
vomeronasal organ was exposed to urine. Eur J Neurosci 11:
2254 –2260, 1999.
Ishii T, Mombaerts P. Expression of nonclassical class I major
histocompatibility genes defines a tripartite organization of the
mouse vomeronasal system. J Neurosci 28: 2332–2341, 2008.
Iwema CL, Fang H, Kurtz DB, Youngentob SL, Schwob JE.
Odorant receptor expression patterns are restored in lesion-recovered rat olfactory epithelium. J Neurosci 24: 356 –369, 2004.
Jackson DE, Ratnieks FL. Communication in ants. Curr Biol 16:
R570 –R574, 2006.
Jackson LM, Harder JD. Vomeronasal organ removal blocks
pheromonal induction of estrus in gray short-tailed opossums (Monodelphis domestica). Biol Reprod 54: 506 –512, 1996.
Jacob S, Kinnunen LH, Metz J, Cooper M, McClintock MK.
Sustained human chemosignal unconsciously alters brain function.
Neuroreport 12: 2391–2394, 2001.
Jacob S, Zelano B, Gungor A, Abbott D, Naclerio R, McClintock
MK. Location and gross morphology of the nasopalatine duct in
human adults. Arch Otolaryngol Head Neck Surg 126: 741–748, 2000.
Physiol Rev • VOL
134. Jacobson L, Trotier D, Doving KB. Anatomical description of a
new organ in the nose of domesticated animals by Ludvig Jacobson
(1813). Chem Senses 23: 743–754, 1998.
135. Jakupovic J, Kang N, Baum MJ. Effect of bilateral accessory
olfactory bulb lesions on volatile urinary odor discrimination and
investigation as well as mating behavior in male mice. Physiol
Behav 93: 467– 473, 2008.
136. Jang T, Singer AG, O’Connell RJ. Induction of c-fos in hamster
accessory olfactory bulbs by natural and cloned aphrodisin. Neuroreport 12: 449 – 452, 2001.
137. Jemiolo B, Andreolini F, Xie TM, Wiesler D, Novotny M.
Puberty-affecting synthetic analogs of urinary chemosignals in the
house mouse, Mus domesticus. Physiol Behav 46: 293–298, 1989.
138. Jemiolo B, Harvey S, Novotny M. Promotion of the Whitten
effect in female mice by synthetic analogs of male urinary constituents. Proc Natl Acad Sci USA 83: 4576 – 4579, 1986.
139. Jemiolo B, Xie TM, Novotny M. Socio-sexual olfactory preference in female mice: attractiveness of synthetic chemosignals.
Physiol Behav 50: 1119 –1122, 1991.
140. Jemiolo B, Xie TM, Novotny M. Urine marking in male mice:
responses to natural and synthetic chemosignals. Physiol Behav
52: 521–526, 1992.
141. Jia C, Chen WR, Shepherd GM. Synaptic organization and neurotransmitters in the rat accessory olfactory bulb. J Neurophysiol
81: 345–355, 1999.
142. Jia C, Goldman G, Halpern M. Development of vomeronasal
receptor neuron subclasses and establishment of topographic projections to the accessory olfactory bulb. Brain Res 102: 209 –216,
1997.
143. Jia C, Halpern M. Subclasses of vomeronasal receptor neurons:
differential expression of G proteins [Gi alpha 2 and G(o alpha)]
and segregated projections to the accessory olfactory bulb. Brain
Res 719: 117–128, 1996.
144. Johnson EW, Eller PM, Jafek BW. An immuno-electron microscopic comparison of olfactory marker protein localization in the
supranuclear regions of the rat olfactory epithelium and vomeronasal organ neuroepithelium. Acta Otolaryngol 113: 766 –771, 1993.
145. Johnston RE. Vomeronasal and/or olfactory mediation of ultrasonic calling and scent marking by female golden hamsters.
Physiol Behav 51: 437– 448, 1992.
146. Johnston RE, Rasmussen K. Individual recognition of female
hamsters by males: role of chemical cues and of the olfactory and
vomeronasal systems. Physiol Behav 33: 95–104, 1984.
147. Jolly A. Lemur social behavior and primate intelligence. Science
153: 501–506, 1966.
148. Jourdan F. [Ultrastructure of the olfactory epithelium of the rat:
polymorphism of the receptors]. C R Acad Sci Hebd Seances Acad
Sci D 280: 443– 446, 1975.
149. Juilfs DM, Fulle HJ, Zhao AZ, Houslay MD, Garbers DL,
Beavo JA. A subset of olfactory neurons that selectively express
cGMP-stimulated phosphodiesterase (PDE2) and guanylyl cyclase-D define a unique olfactory signal transduction pathway. Proc
Natl Acad Sci USA 94: 3388 –3395, 1997.
150. Kaluza JF, Gussing F, Bohm S, Breer H, Strotmann J. Olfactory receptors in the mouse septal organ. J Neurosci Res 76:
442– 452, 2004.
151. Karlson P, Luscher M. Pheromones’: a new term for a class of
biologically active substances. Nature 183: 55–56, 1959.
152. Karunadasa DK, Chapman C, Bicknell RJ. Expression of pheromone receptor gene families during olfactory development in the
mouse: expression of a V1 receptor in the main olfactory epithelium. Eur J Neurosci 23: 2563–2572, 2006.
153. Kaupp UB, Seifert R. Cyclic nucleotide-gated ion channels.
Physiol Rev 82: 769 – 824, 2002.
154. Keller M, Douhard Q, Baum MJ, Bakker J. Destruction of the
main olfactory epithelium reduces female sexual behavior and
olfactory investigation in female mice. Chem Senses 31: 315–323,
2006.
155. Keller M, Pierman S, Douhard Q, Baum MJ, Bakker J. The
vomeronasal organ is required for the expression of lordosis behaviour, but not sex discrimination in female mice. Eur J Neurosci
23: 521–530, 2006.
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
115.
TIRINDELLI ET AL.
FROM PHEROMONES TO BEHAVIOR
Physiol Rev • VOL
lebacks: an immuno-ecological approach. Proc Biol Sci 273: 1407–
1414, 2006.
181. Labov JB, Wysocki CJ. Vomeronasal organ and social factors
affect urine marking by male mice. Physiol Behav 45: 443– 447,
1989.
182. Larriva-Sahd J. The accessory olfactory bulb in the adult rat: a
cytological study of its cell types, neuropil, neuronal modules, and
interactions with the main olfactory system. J Comp Neurol 510:
309 –350, 2008.
183. Lau YE, Cherry JA. Distribution of PDE4A and G(o) alpha immunoreactivity in the accessory olfactory system of the mouse. Neuroreport 11: 27–32, 2000.
183a.Le Y, Murphy PM, Wang JM. Formyl-peptide receptors revisited.
Trends Immunol 23: 541–548, 2002.
184. Lee SJ, Mammen A, Kim EJ, Kim SY, Park YJ, Park M, Han
HS, Bae YC, Ronnett GV, Moon C. The vomeronasal organ and
adjacent glands express components of signaling cascades found in
sensory neurons in the main olfactory system. Mol Cells 26: 2008.
185. Leinders-Zufall T, Brennan P, Widmayer P, PCS, Maul-Pavicic
A, Jager M, Li XH, Breer H, Zufall F, Boehm T. MHC class I
peptides as chemosensory signals in the vomeronasal organ. Science 306: 1033–1037, 2004.
186. Leinders-Zufall T, Cockerham RE, Michalakis S, Biel M, Garbers DL, Reed RR, Zufall F, Munger SD. Contribution of the
receptor guanylyl cyclase GC-D to chemosensory function in the
olfactory epithelium. Proc Natl Acad Sci USA 104: 14507–14512,
2007.
187. Leinders-Zufall T, Lane AP, Puche AC, Ma W, Novotny MV,
Shipley MT, Zufall F. Ultrasensitive pheromone detection by
mammalian vomeronasal neurons. Nature 405: 792–796, 2000.
188. Leon M, Moltz H. Endocrine control of the maternal pheromone
in the postpartum female rat. Physiol Behav 10: 65– 67, 1973.
189. Lepri JJ, Wysocki CJ. Removal of the vomeronasal organ disrupts the activation of reproduction in female voles. Physiol Behav
40: 349 –355, 1987.
190. Levai O, Feistel T, Breer H, Strotmann J. Cells in the vomeronasal organ express odorant receptors but project to the accessory
olfactory bulb. J Comp Neurol 498: 476 – 490, 2006.
191. Levai O, Strotmann J. Projection pattern of nerve fibers from the
septal organ: DiI-tracing studies with transgenic OMP mice. Histochem Cell Biol 120: 483– 492, 2003.
192. Lewin AH. Receptors of mammalian trace amines. AAPS J 8:
E138 –E145, 2006.
193. Leypold BG, Yu CR, Leinders-Zufall T, Kim MM, Zufall F, Axel
R. Altered sexual and social behaviors in trp2 mutant mice. Proc
Natl Acad Sci USA 99: 6376 – 6381, 2002.
194. Liberles SD, Buck LB. A second class of chemosensory receptors
in the olfactory epithelium. Nature 442: 645– 650, 2006.
195. Liman ER. Regulation by voltage and adenine nucleotides of a
Ca2⫹-activated cation channel from hamster vomeronasal sensory
neurons. J Physiol 548: 777–787, 2003.
196. Liman ER, Corey DP. Electrophysiological characterization of
chemosensory neurons from the mouse vomeronasal organ. J Neurosci 16: 4625– 4637, 1996.
197. Liman ER, Corey DP, Dulac C. TRP2: a candidate transduction
channel for mammalian pheromone sensory signaling. Proc Natl
Acad Sci USA 96: 5791–5796, 1999.
198. Liman ER, Innan H. Relaxed selective pressure on an essential
component of pheromone transduction in primate evolution. Proc
Natl Acad Sci USA 100: 3328 –3332, 2003.
199. Lin DY, Zhang SZ, Block E, Katz LC. Encoding social signals in
the mouse main olfactory bulb. Nature 434: 470 – 477, 2005.
200. Lin W, Arellano J, Slotnick B, Restrepo D. Odors detected by
mice deficient in cyclic nucleotide-gated channel subunit A2 stimulate the main olfactory system. J Neurosci 24: 3703–3710, 2004.
201. Lin W, Margolskee R, Donnert G, Hell SW, Restrepo D. Olfactory neurons expressing transient receptor potential channel M5
(TRPM5) are involved in sensing semiochemicals. Proc Natl Acad
Sci USA 104: 2471–2476, 2007.
202. Lin W, Ogura T, Margolskee RF, Finger TE, Restrepo D.
TRPM5-expressing solitary chemosensory cells respond to odorous
irritants. J Neurophysiol 99: 1451–1460, 2008.
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
156. Kelliher KR, Spehr M, Li XH, Zufall F, Leinders-Zufall T.
Pheromonal recognition memory induced by TRPC2-independent
vomeronasal sensing. Eur J Neurosci 23: 3385–3390, 2006.
157. Keverne EB. Pheromones, vomeronasal function, and genderspecific behavior. Cell 108: 735–738, 2002.
158. Kevetter GA, Winans SS. Connections of the corticomedial
amygdala in the golden hamster. I. Efferents of the “vomeronasal
amygdala.” J Comp Neurol 197: 81–98, 1981.
159. Kevetter GA, Winans SS. Connections of the corticomedial
amygdala in the golden hamster. II. Efferents of the “olfactory
amygdala.” J Comp Neurol 197: 99 –111, 1981.
160. Kikusui T, Takigami S, Takeuchi Y, Mori Y. Alarm pheromone
enhances stress-induced hyperthermia in rats. Physiol Behav 72:
45–50, 2001.
161. Kimchi T, Xu J, Dulac C. A functional circuit underlying male
sexual behaviour in the female mouse brain. Nature 448: 1009 –
1014, 2007.
162. Kimoto H, Haga S, Sato K, Touhara K. Sex-specific peptides
from exocrine glands stimulate mouse vomeronasal sensory neurons. Nature 437: 898 –901, 2005.
163. Kimoto H, Sato K, Nodari F, Haga S, Holy TE, Touhara K. Sexand strain-specific expression and vomeronasal activity of mouse
ESP family peptides. Curr Biol 17: 1879 –1884, 2007.
164. Kiyokawa Y, Kikusui T, Takeuchi Y, Mori Y. Partner’s stress
status influences social buffering effects in rats. Behav Neurosci
118: 798 – 804, 2004.
165. Kiyokawa Y, Takeuchi Y, Mori Y. Two types of social buffering
differentially mitigate conditioned fear responses. Eur J Neurosci
26: 3606 –3613, 2007.
166. Kjaer I, Fischer HB. The human vomeronasal organ: prenatal
developmental stages and distribution of luteinizing hormone-releasing hormone. Eur J Oral Sci 104: 34 – 40, 1996.
167. Kleene SJ. Origin of the chloride current in olfactory transduction. Neuron 11: 123–132, 1993.
168. Kleene SJ. The electrochemical basis of odor transduction in
vertebrate olfactory cilia. Chem Senses 33: 839 – 859, 2008.
169. Kobayakawa K, Kobayakawa R, Matsumoto H, Oka Y, Imai T,
Ikawa M, Okabe M, Ikeda T, Itohara S, Kikusui T, Mori K,
Sakano H. Innate versus learned odour processing in the mouse
olfactory bulb. Nature 450: 503–508, 2007.
170. Kolunie JM, Stern JM. Maternal aggression in rats: effects of
olfactory bulbectomy, ZnSO4-induced anosmia, and vomeronasal
organ removal. Horm Behav 29: 492–518, 1995.
171. Kondo Y, Sudo T, Tomihara K, Sakuma Y. Activation of accessory olfactory bulb neurons during copulatory behavior after deprivation of vomeronasal inputs in male rats. Brain Res 962: 232–
236, 2003.
172. Kondo Y, Tomihara K, Sakuma Y. Sensory requirements for
noncontact penile erection in the rat. Behav Neurosci 113: 1062–
1070, 1999.
173. Koos DS, Fraser SE. The Grueneberg ganglion projects to the
olfactory bulb. Neuroreport 16: 1929 –1932, 2005.
174. Kow LM, Florea C, Schwanzel-Fukuda M, Devidze N, Kami
KH, Lee A, Zhou J, Maclaughlin D, Donahoe P, Pfaff D. Development of a sexually differentiated behavior and its underlying
CNS arousal functions. Curr Top Dev Biol 79: 37–59, 2007.
175. Koyama S, Kamimura S. Effects of vomeronasal organ removal
on the sperm motility in male mice. Zool Sci 20: 1355–1358, 2003.
176. Krieger J, Schmitt A, Lobel D, Gudermann T, Schultz G,
Breer H, Boekhoff I. Selective activation of G protein subtypes in
the vomeronasal organ upon stimulation with urine-derived compounds. J Biol Chem 274: 4655– 4662, 1999.
177. Kroner C, Breer H, Singer AG, O’Connell RJ. Pheromoneinduced second messenger signaling in the hamster vomeronasal
organ. Neuroreport 7: 2989 –2992, 1996.
178. Kurahashi T, Menini A. Mechanism of odorant adaptation in the
olfactory receptor cell. Nature 385: 725–729, 1997.
179. Kurahashi T, Yau KW. Co-existence of cationic and chloride
components in odorant-induced current of vertebrate olfactory
receptor cells. Nature 363: 71–74, 1993.
180. Kurtz J, Wegner KM, Kalbe M, Reusch TB, Schaschl H, Hasselquist D, Milinski M. MHC genes and oxidative stress in stick-
951
952
TIRINDELLI ET AL.
Physiol Rev • VOL
226. Malnic B, Hirono J, Sato T, Buck LB. Combinatorial receptor
codes for odors. Cell 96: 713–723, 1999.
227. Mandiyan VS, Coats JK, Shah NM. Deficits in sexual and aggressive behaviors in Cnga2 mutant mice. Nat Neurosci 8: 1660 –1662,
2005.
228. Martel KL, Baum MJ. Sexually dimorphic activation of the accessory, but not the main, olfactory bulb in mice by urinary volatiles.
Eur J Neurosci 26: 463– 475, 2007.
229. Martinez-Marcos A. On the organization of olfactory and vomeronasal cortices. Prog Neurobiol 87: 21–30, 2009.
230. Martinez-Marcos A, Halpern M. Differential projections from
the anterior and posterior divisions of the accessory olfactory bulb
to the medial amygdala in the opossum, Monodelphis domestica.
Eur J Neurosci 11: 3789 –3799, 1999.
231. Martini S, Silvotti L, Shirazi A, Ryba NJ, Tirindelli R. Coexpression of putative pheromone receptors in the sensory neurons of the vomeronasal organ. J Neurosci 21: 843– 848, 2001.
232. Martins Y, Preti G, Crabtree CR, Runyan T, Vainius AA,
Wysocki CJ. Preference for human body odors is influenced by
gender and sexual orientation. Psychol Sci 16: 694 –701, 2005.
233. Maruniak JA, Wysocki CJ, Taylor JA. Mediation of male mouse
urine marking and aggression by the vomeronasal organ. Physiol
Behav 37: 655– 657, 1986.
234. Matsunami H, Buck LB. A multigene family encoding a diverse
array of putative pheromone receptors in mammals. Cell 90: 775–
784, 1997.
235. Matthews HR, Reisert J. Calcium, the two-faced messenger of
olfactory transduction and adaptation. Curr Opin Neurobiol 13:
469 – 475, 2003.
236. McCarthy MM, Auger AP. He’s a lover, not a fighter–smell, sex
and civility. Trends Endocrinol Metab 13: 183–184, 2002.
237. McClintock MK. Menstrual synchorony and suppression. Nature
229: 244 –245, 1971.
238. McClintock MK. Estrous synchrony: modulation of ovarian cycle
length by female pheromones. Physiol Behav 32: 701–705, 1984.
239. McClintock MK. On the nature of mammalian and human pheromones. Ann NY Acad Sci 855: 390 –392, 1998.
240. Meisami E, Bhatnagar KP. Structure and diversity in mammalian
accessory olfactory bulb. Microsc Res Tech 43: 476 – 499, 1998.
241. Menco BP, Carr VM, Ezeh PI, Liman ER, Yankova MP. Ultrastructural localization of G-proteins and the channel protein TRP2
to microvilli of rat vomeronasal receptor cells. J Comp Neurol 438:
468 – 489, 2001.
242. Menini A. Calcium signalling and regulation in olfactory neurons.
Curr Opin Neurobiol 9: 419 – 426, 1999.
243. Mennella JA, Moltz H. Infanticide in the male rat: the role of the
vomeronasal organ. Physiol Behav 42: 303–306, 1988.
244. Meredith M. Chronic recording of vomeronasal pump activation
in awake behaving hamsters. Physiol Behav 56: 345–354, 1994.
245. Meredith M. Vomeronasal function. Chem Senses 23: 463– 466,
1998.
246. Meredith M. Human vomeronasal organ function: a critical review
of best and worst cases. Chem Senses 26: 433– 445, 2001.
247. Meredith M, Howard G. Intracerebroventricular LHRH relieves
behavioral deficits due to vomeronasal organ removal. Brain Res
Bull 29: 75–79, 1992.
248. Meredith M, Marques DM, O’Connell RO, Stern FL. Vomeronasal pump: significance for male hamster sexual behavior. Science
207: 1224 –1226, 1980.
249. Meredith M, O’Connell RJ. Efferent control of stimulus access to
the hamster vomeronasal organ. J Physiol 286: 301–316, 1979.
249a.Migeotte I, Communi D, Parmentier M. Formyl peptide receptors: a promiscuous subfamily of G protein-coupled receptors controlling immune responses. Cytokine Growth Factor Rev 17: 501–
519, 2006.
250. Minke B, Parnas M. Insights on TRP channels from in vivo studies
in Drosophila. Annu Rev Physiol 68: 649 – 684, 2006.
251. Miyamichi K, Serizawa S, Kimura HM, Sakano H. Continuous
and overlapping expression domains of odorant receptor genes in
the olfactory epithelium determine the dorsal/ventral positioning of
glomeruli in the olfactory bulb. J Neurosci 25: 3586 –3592, 2005.
252. Mohedano-Moriano A, Pro-Sistiaga P, Ubeda-Banon I, Crespo
C, Insausti R, Martinez-Marcos A. Segregated pathways to the
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
203. Lindemann L, Hoener MC. A renaissance in trace amines inspired by a novel GPCR family. Trends Pharmacol Sci 26: 274 –281,
2005.
204. Liu D, Liman ER. Intracellular Ca2⫹ and the phospholipid PIP2
regulate the taste transduction ion channel TRPM5. Proc Natl Acad
Sci USA 100: 15160 –15165, 2003.
205. Lledo PM, Saghatelyan A. Integrating new neurons into the adult
olfactory bulb: joining the network, life-death decisions, and the
effects of sensory experience. Trends Neurosci 28: 248 –254, 2005.
206. Lloyd-Thomas A, Keverne EB. Role of the brain and accessory
olfactory system in the block to pregnancy in mice. Neuroscience 7:
907–913, 1982.
207. Loconto J, Papes F, Chang E, Stowers L, Jones EP, Takada T,
Kumanovics A, Fischer LK, Dulac C. Functional expression of
murine V2R pheromone receptors involves selective association
with the M10 and M1 families of MHC class Ib molecules. Cell 112:
607– 618, 2003.
208. Logan DW, Marton TF, Stowers L. Species specificity in major
urinary proteins by parallel evolution. PLoS ONE 3: e3280, 2008.
209. Lomas DE, Keverne EB. Role of the vomeronasal organ and
prolactin in the acceleration of puberty in female mice. J Reprod
Fertil 66: 101–107, 1982.
210. Lowe G. Flash photolysis reveals a diversity of ionotropic glutamate receptors on the mitral cell somatodendritic membrane.
J Neurophysiol 90: 1737–1746, 2003.
211. Lucas P, Ukhanov K, Leinders-Zufall T, Zufall F. A diacylglycerol-gated cation channel in vomeronasal neuron dendrites is impaired in TRPC2 mutant mice: mechanism of pheromone transduction. Neuron 40: 551–561, 2003.
212. Lundstrom JN, Hummel T, Olsson MJ. Individual differences in
sensitivity to the odor of 4,16-androstadien-3-one. Chem Senses 28:
643– 650, 2003.
213. Luo M, Fee MS, Katz LC. Encoding pheromonal signals in the
accessory olfactory bulb of behaving mice. Science 299: 1196 –1201,
2003.
214. Luo M, Katz LC. Encoding pheromonal signals in the mammalian
vomeronasal system. Curr Opin Neurobiol 14: 428 – 434, 2004.
215. Luo Y, Lu S, Chen P, Wang D, Halpern M. Identification of
chemoattractant receptors and G-proteins in the vomeronasal system of garter snakes. J Biol Chem 269: 16867–16877, 1994.
216. Ma D, Allen ND, Van Bergen YC, Jones CM, Baum MJ,
Keverne EB, Brennan PA. Selective ablation of olfactory receptor neurons without functional impairment of vomeronasal receptor neurons in OMP-ntr transgenic mice. Eur J Neurosci 16: 2317–
2323, 2002.
217. Ma M. Encoding olfactory signals via multiple chemosensory systems. Crit Rev Biochem Mol Biol 42: 463– 480, 2007.
218. Ma M, Grosmaitre X, Iwema CL, Baker H, Greer CA, Shepherd GM. Olfactory signal transduction in the mouse septal organ.
J Neurosci 23: 317–324, 2003.
219. Ma W, Miao Z, Novotny MV. Role of the adrenal gland and
adrenal-mediated chemosignals in suppression of estrus in the
house mouse: the lee-boot effect revisited. Biol Reprod 59: 1317–
1320, 1998.
220. Mackay-Sim A, Laing DG. Discrimination of odors from stressed
rats by non-stressed rats. Physiol Behav 24: 699 –704, 1980.
221. Mackay-Sim A, Rose JD. Removal of the vomeronasal organ
impairs lordosis in female hamsters: effect is reversed by luteinising hormone-releasing hormone. Neuroendocrinology 42: 489 – 493,
1986.
222. Macrides F, Singer AG, Clancy AN, Goldman BD, Agosta WC.
Male hamster investigatory and copulatory responses to vaginal
discharge: relationship to the endocrine status of females. Physiol
Behav 33: 633– 637, 1984.
223. Magert HJ, Cieslak A, Alkan O, Luscher B, Kauffels W, Forssmann WG. The golden hamster aphrodisin gene. Structure, expression in parotid glands of female animals, and comparison with a
similar murine gene. J Biol Chem 274: 444 – 450, 1999.
224. Malnic B. Searching for the ligands of odorant receptors. Mol
Neurobiol 35: 175–181, 2007.
225. Malnic B, Godfrey PA, Buck LB. The human olfactory receptor
gene family. Proc Natl Acad Sci USA 101: 2584 –2589, 2004.
FROM PHEROMONES TO BEHAVIOR
253.
254.
255.
256.
257.
259.
260.
261.
262.
263.
264.
265.
266.
267.
268.
269.
270.
271.
272.
273.
274.
275.
276.
Physiol Rev • VOL
277. Nodari F, Hsu FF, Fu X, Holekamp TF, Kao LF, Turk J, Holy
TE. Sulfated steroids as natural ligands of mouse pheromonesensing neurons. J Neurosci 28: 6407– 6418, 2008.
278. Norlin EM, Gussing F, Berghard A. Vomeronasal phenotype and
behavioral alterations in G alpha i2 mutant mice. Curr Biol 13:
1214 –1219, 2003.
279. Novotny M, Harvey S, Jemiolo B. Chemistry of male dominance
in the house mouse, Mus domesticus. Experientia 46: 109 –113,
1990.
280. Novotny M, Harvey S, Jemiolo B, Alberts J. Synthetic pheromones that promote inter-male aggression in mice. Proc Natl Acad
Sci USA 82: 2059 –2061, 1985.
281. Novotny M, Jemiolo B, Harvey S, Wiesler D, Marchlewska-Koj
A. Adrenal-mediated endogenous metabolites inhibit puberty in
female mice. Science 231: 722–725, 1986.
282. Novotny MV, Jemiolo B, Wiesler D, Ma W, Harvey S, Xu F, Xie
TM, Carmack M. A unique urinary constituent, 6-hydroxy-6-methyl3-heptanone, is a pheromone that accelerates puberty in female mice.
Chem Biol 6: 377–383, 1999.
283. Novotny MV, Ma W, Wiesler D, Zidek L. Positive identification of
the puberty-accelerating pheromone of the house mouse: the volatile ligands associating with the major urinary protein. Proc Biol
Sci 266: 2017–2022, 1999.
284. Pankevich DE, Baum MJ, Cherry JA. Olfactory sex discrimination persists, whereas the preference for urinary odorants from
estrous females disappears in male mice after vomeronasal organ
removal. J Neurosci 24: 9451–9457, 2004.
285. Pantages E, Dulac C. A novel family of candidate pheromone
receptors in mammals. Neuron 28: 835– 845, 2000.
286. Paulos MA, Tessel RE. Excretion of beta-phenethylamine is elevated in humans after profound stress. Science 215: 1127–1129,
1982.
287. Pearce GP, Hughes PE. The influence of male contact on plasma
cortisol concentrations in the prepubertal gilt. J Reprod Fertil 80:
417– 424, 1987.
288. Penn D, Potts W. MHC-disassortative mating preferences reversed by cross-fostering. Proc Biol Sci 265: 1299 –1306, 1998.
289. Petrovich GD, Canteras NS, Swanson LW. Combinatorial amygdalar inputs to hippocampal domains and hypothalamic behavior
systems. Brain Res 38: 247–289, 2001.
290. Petrulis A, Peng M, Johnston RE. Effects of vomeronasal organ
removal on individual odor discrimination, sex-odor preference,
and scent marking by female hamsters. Physiol Behav 66: 73– 83,
1999.
291. Pfeiffer CA, Johnston RE. Hormonal and behavioral responses
of male hamsters to females and female odors: roles of olfaction,
the vomeronasal system, and sexual experience. Physiol Behav 55:
129 –138, 1994.
292. Pifferi S, Boccaccio A, Menini A. Cyclic nucleotide-gated ion
channels in sensory transduction. FEBS Lett 580: 2853–2859, 2006.
293. Pin JP, Kniazeff J, Goudet C, Bessis AS, Liu J, Galvez T,
Acher F, Rondard P, Prezeau L. The activation mechanism of
class-C G-protein coupled receptors. Biol Cell 96: 335–342, 2004.
294. Portfors CV. Types and functions of ultrasonic vocalizations in
laboratory rats and mice. J Am Assoc Lab Anim Sci 46: 28 –34,
2007.
295. Portillo W, Diaz NF, Retana-Marquez S, Paredes RG. Olfactory, partner preference and Fos expression in the vomeronasal
projection pathway of sexually sluggish male rats. Physiol Behav
88: 389 –397, 2006.
296. Potthast R, Potter LR. Phosphorylation-dependent regulation of
the guanylyl cyclase-linked natriuretic peptide receptors. Peptides
26: 1001–1008, 2005.
297. Potts WK, Wakeland EK. Evolution of MHC genetic diversity: a
tale of incest, pestilence and sexual preference. Trends Genet 9:
408 – 412, 1993.
298. Powers JB, Winans SS. Vomeronasal organ: critical role in mediating sexual behavior of the male hamster. Science 187: 961–963,
1975.
299. Prawitt D, Monteilh-Zoller MK, Brixel L, Spangenberg C,
Zabel B, Fleig A, Penner R. TRPM5 is a transient Ca2⫹-activated
cation channel responding to rapid changes in [Ca2⫹]i. Proc Natl
Acad Sci USA 100: 15166 –15171, 2003.
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
258.
vomeronasal amygdala: differential projections from the anterior
and posterior divisions of the accessory olfactory bulb. Eur J Neurosci 25: 2065–2080, 2007.
Mohedano-Moriano A, Pro-Sistiaga P, Ubeda-Banon I, RosaPrieto C, Saiz-Sanchez D, Martinez-Marcos A. V1R and V2R
segregated vomeronasal pathways to the hypothalamus. Neuroreport 19: 1623–1626, 2008.
Moltz H, Lee TM. The maternal pheromone of the rat: identity and
functional significance. Physiol Behav 26: 301–306, 1981.
Mombaerts P. Seven-transmembrane proteins as odorant and chemosensory receptors. Science 286: 707–711, 1999.
Mombaerts P. Genes and ligands for odorant, vomeronasal and
taste receptors. Nat Rev Neurosci 5: 263–278, 2004.
Mombaerts P, Wang F, Dulac C, Chao SK, Nemes A, Mendelsohn M, Edmondson J, Axel R. Visualizing an olfactory sensory
map. Cell 87: 675– 686, 1996.
Montagna W, Yun JS. The skin of primates. X. The skin of the
ring-tailed lemur (Lemur catta). Am J Phys Anthropol 20: 95–117,
1962.
Moran DT, Jafek BW, Rowley JC III. The vomeronasal (Jacobson’s) organ in man: ultrastructure and frequency of occurrence. J
Steroid Biochem Mol Biol 39: 545–552, 1991.
Moran DT, Rowley JC III, Jafek BW. Electron microscopy of
human olfactory epithelium reveals a new cell type: the microvillar
cell. Brain Res 253: 39 – 46, 1982.
Mori K. Monoclonal antibodies (2C5 and 4C9) against lactoseries
carbohydrates identify subsets of olfactory and vomeronasal receptor cells and their axons in the rabbit. Brain Res 408: 215–221, 1987.
Morrison EE, Costanzo RM. Morphology of the human olfactory
epithelium. J Comp Neurol 297: 1–13, 1990.
Mucignat-Caretta C. Modulation of exploratory behavior in female mice by protein-borne male urinary molecules. J Chem Ecol
28: 1853–1863, 2002.
Mucignat-Caretta C, Caretta A. Urinary chemical cues affect
light avoidance behaviour in male laboratory mice, Mus musculus.
Anim Behav 57: 765–769, 1999.
Mucignat-Caretta C, Caretta A, Baldini E. Protein-bound male
urinary pheromones: differential responses according to age and
gender. Chem Senses 23: 67–70, 1998.
Mucignat-Caretta C, Caretta A, Cavaggioni A. Acceleration of
puberty onset in female mice by male urinary proteins. J Physiol
486: 517–522, 1995.
Mucignat-Caretta C, Cavaggioni A, Caretta A. Male urinary
chemosignals differentially affect aggressive behavior in male
mice. J Chem Ecol 30: 777–791, 2004.
Muller-Schwarze D. Chemical Ecology: Odor Communication in
Animals. New York: Elsevier, 1979.
Munger SD, Leinders-Zufall T, Zufall F. Subsystem organization
of the mammalian sense of smell. Annu Rev Physiol 71: 115–140,
2008.
Muroi Y, Ishii T, Komori S, Kitamura N, Nishimura M. Volatile
female odors activate the accessory olfactory system of male mice
without physical contact. Neuroscience 141: 551–558, 2006.
Murphy FA, Tucker K, Fadool DA. Sexual dimorphism and
developmental expression of signal-transduction machinery in the
vomeronasal organ. J Comp Neurol 432: 61–74, 2001.
Newman R, Winans SS. An experimental study of the ventral
striatum of the golden hamster. II. Neuronal connections of the
olfactory tubercle. J Comp Neurol 191: 193–212, 1980.
Ngai J, Dowling MM, Buck L, Axel R, Chess A. The family of
genes encoding odorant receptors in the channel catfish. Cell 72:
657– 666, 1993.
Niimura Y, Nei M. Evolutionary dynamics of olfactory receptor
genes in fishes and tetrapods. Proc Natl Acad Sci USA 102: 6039 –
6044, 2005.
Nilius B, Vennekens R. From cardiac cation channels to the
molecular dissection of the transient receptor potential channel
TRPM4. Pflügers Arch 453: 313–321, 2006.
Nishimura K, Utsumi K, Yuhara M, Fujitani Y, Iritani A. Identification of puberty-accelerating pheromones in male mouse urine.
J Exp Zool 251: 300 –305, 1989.
953
954
TIRINDELLI ET AL.
Physiol Rev • VOL
323. Ryba NJ, Tirindelli R. A new multigene family of putative pheromone receptors. Neuron 19: 371–379, 1997.
324. Saito TR, Hokao R, Imamichi T. Incidence of lordosis behavior
in the female rat following removal of the vomeronasal organ.
Jikken Dobutsu 37: 93–95, 1988.
325. Saito TR, Moltz H. Copulatory behavior of sexually naive and
sexually experienced male rats following removal of the vomeronasal organ. Physiol Behav 37: 507–510, 1986.
326. Saito TR, Moltz H. Sexual behavior in the female rat following
removal of the vomeronasal organ. Physiol Behav 38: 81– 87, 1986.
327. Salazar I, Brennan PA. Retrograde labelling of mitral/tufted cells
in the mouse accessory olfactory bulb following local injections of
the lipophilic tracer DiI into the vomeronasal amygdala. Brain Res
896: 198 –203, 2001.
328. Salazar I, Sanchez-Quinteiro P, Aleman N, Prieto D. Anatomical, immnunohistochemical and physiological characteristics of
the vomeronasal vessels in cows and their possible role in vomeronasal reception. J Anat 212: 686 – 696, 2008.
329. Sam M, Vora S, Malnic B, Ma W, Novotny MV, Buck Neuropharmacology LB. Odorants may arouse instinctive behaviours.
Nature 412: 142, 2001.
330. Sasaki K, Okamoto K, Inamura K, Tokumitsu Y, Kashiwayanagi M. Inositol-1,4,5-trisphosphate accumulation induced by urinary pheromones in female rat vomeronasal epithelium. Brain Res
823: 161–168, 1999.
331. Savic I. Brain imaging studies of the functional organization of
human olfaction. Chem Senses 30 Suppl 1: i222–i223, 2005.
332. Savic I, Berglund H, Gulyas B, Roland P. Smelling of odorous
sex hormone-like compounds causes sex-differentiated hypothalamic activations in humans. Neuron 31: 661– 668, 2001.
333. Sbarbati A, Osculati F. Allelochemical communication in vertebrates: kairomones, allomones and synomones. Cells Tissues Organs 183: 206 –219, 2006.
334. Scalia F, Winans SS. The differential projections of the olfactory
bulb and accessory olfactory bulb in mammals. J Comp Neurol 161:
31–55, 1975.
335. Schaal B, Coureaud G, Langlois D, Ginies C, Semon E, Perrier
G. Chemical and behavioural characterization of the rabbit mammary pheromone. Nature 424: 68 –72, 2003.
336. Schaefer ML, Yamazaki K, Osada K, Restrepo D, Beauchamp
GK. Olfactory fingerprints for major histocompatibility complexdetermined body odors II: relationship among odor maps, genetics,
odor composition, and behavior. J Neurosci 22: 9513–9521, 2002.
337. Schild D, Restrepo D. Transduction mechanisms in vertebrate
olfactory receptor cells. Physiol Rev 78: 429 – 466, 1998.
338. Schilling A. Olfactory communication in prosimians. In: The Study
of Prosimian Behavior, edited by Doyle GA and Martin RD. New
York: Academic, 1979.
339. Schoppa NE, Urban NN. Dendritic processing within olfactory
bulb circuits. Trends Neurosci 26: 501–506, 2003.
340. Schwarting GA, Crandall JE. Subsets of olfactory and vomeronasal sensory epithelial cells and axons revealed by monoclonal
antibodies to carbohydrate antigens. Brain Res 547: 239 –248, 1991.
341. Schwarting GA, Deutsch G, Gattey DM, Crandall JE. Glycoconjugates are stage- and position-specific cell surface molecules
in the developing olfactory system. 1: The CC1 immunoreactive
glycolipid defines a rostrocaudal gradient in the rat vomeronasal
system. J Neurobiol 23: 120 –129, 1992.
342. Schwende FJ, Wiesler D, Novotny M. Volatile compounds associated with estrus in mouse urine: potential pheromones. Experientia 40: 213–215, 1984.
343. Serizawa S, Miyamichi K, Nakatani H, Suzuki M, Saito M,
Yoshihara Y, Sakano H. Negative feedback regulation ensures the
one receptor-one olfactory neuron rule in mouse. Science 302:
2088 –2094, 2003.
344. Shahan K, Denaro M, Gilmartin M, Shi Y, Derman E. Expression of six mouse major urinary protein genes in the mammary,
parotid, sublingual, submaxillary, and lachrymal glands and in the
liver. Mol Cell Biol 7: 1947–1954, 1987.
345. Shepherd GM, Chen WR, Willhite D, Migliore M, Greer CA.
The olfactory granule cell: from classical enigma to central role in
olfactory processing. Brain Res Rev 55: 373–382, 2007.
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
300. Preti G, Wysocki CJ, Barnhart KT, Sondheimer SJ, Leyden
JJ. Male axillary extracts contain pheromones that affect pulsatile
secretion of luteinizing hormone and mood in women recipients.
Biol Reprod 68: 2107–2113, 2003.
301. Price MA, Vandenbergh JG. Analysis of puberty-accelerating
pheromones. J Exp Zool 264: 42– 45, 1992.
302. Rajendren G, Dudley CA, Moss RL. Role of the vomeronasal
organ in the male-induced enhancement of sexual receptivity in
female rats. Neuroendocrinology 52: 368 –372, 1990.
303. Rall W, Shepherd GM, Reese TS, Brightman MW. Dendrodendritic synaptic pathway for inhibition in the olfactory bulb. Exp
Neurol 14: 44 –56, 1966.
304. Rasmussen LE, Greenwood DR. Frontalin: a chemical message
of musth in Asian elephants (Elephas maximus). Chem Senses 28:
433– 446, 2003.
305. Rasmussen LE, Lee TD, Roelofs WL, Zhang A, Daves GD Jr.
Insect pheromone in elephants. Nature 379: 684, 1996.
306. Rasmussen LE, Lee TD, Zhang A, Roelofs WL, Daves GD Jr.
Purification, identification, concentration and bioactivity of (Z)-7dodecen-1-yl acetate: sex pheromone of the female Asian elephant,
Elephas maximus. Chem Senses 22: 417– 437, 1997.
307. Rasmussen LE, Schulte BA. Chemical signals in the reproduction
of Asian (Elephas maximus) and African (Loxodonta africana)
elephants. Anim Reprod Sci 53: 19 –34, 1998.
308. Ressler KJ, Sullivan SL, Buck LB. A zonal organization of odorant receptor gene expression in the olfactory epithelium. Cell 73:
597– 609, 1993.
309. Richards DB, Stevens DA. Evidence for marking with urine by
rats. Behav Biol 12: 517–523, 1974.
310. Rissman EF, Crews D. Effect of castration on epididymal sperm
storage in male musk shrews (Suncus murinus) and mice (Mus
musculus). J Reprod Fertil 86: 219 –222, 1989.
310a.Rivière S, Challet L, Fluegge D, Spehr M, Rodriguez I. Formyl
peptide receptor-like proteins are a novel family of vomeronasal
chemosensors. Nature 2009 Apr 22. [Epub ahead of print] doi:
10.1038/nature08029.
311. Rodolfo-Masera T. Sur l’esistenza di un particulare organo olfattivo nel setto nasale della cavia e di altri roditori. Arch Ital Anat
Embryol 48: 157–212, 1943.
312. Rodriguez I. Pheromone receptors in mammals. Horm Behav 46:
219 –230, 2004.
313. Rodriguez I, Del Punta K, Rothman A, Ishii T, Mombaerts P.
Multiple new and isolated families within the mouse superfamily of
V1r vomeronasal receptors. Nat Neurosci 5: 134 –140, 2002.
314. Rodriguez I, Feinstein P, Mombaerts P. Variable patterns of
axonal projections of sensory neurons in the mouse vomeronasal
system. Cell 97: 199 –208, 1999.
315. Rodriguez I, Greer CA, Mok MY, Mombaerts P. A putative
pheromone receptor gene expressed in human olfactory mucosa.
Nat Genet 26: 18 –19, 2000.
316. Rodriguez I, Mombaerts P. Novel human vomeronasal receptorlike genes reveal species-specific families. Curr Biol 12: R409 –
R411, 2002.
317. Roppolo D, Ribaud V, Jungo VP, Luscher C, Rodriguez I.
Projection of the Gruneberg ganglion to the mouse olfactory bulb.
Eur J Neurosci 23: 2887–2894, 2006.
318. Roppolo D, Vollery S, Kan CD, Luscher C, Broillet MC, Rodriguez I. Gene cluster lock after pheromone receptor gene
choice. EMBO J 26: 3423–3430, 2007.
319. Rossler P, Kroner C, Krieger J, Lobel D, Breer H, Boekhoff I.
Cyclic adenosine monophosphate signaling in the rat vomeronasal
organ: role of an adenylyl cyclase type VI. Chem Senses 25: 313–
322, 2000.
320. Royal SJ, Key B. Development of P2 olfactory glomeruli in P2internal ribosome entry site-tau-LacZ transgenic mice. J Neurosci
19: 9856 –9864, 1999.
321. Runnenburger K, Breer H, Boekhoff I. Selective G protein beta
gamma-subunit compositions mediate phospholipase C activation
in the vomeronasal organ. Eur J Cell Biol 81: 539 –547, 2002.
322. Ryba NJ, Tirindelli R. A novel GTP-binding protein gammasubunit, G gamma 8, is expressed during neurogenesis in the
olfactory and vomeronasal neuroepithelia. J Biol Chem 270: 6757–
6767, 1995.
FROM PHEROMONES TO BEHAVIOR
Physiol Rev • VOL
receptor subtypes are spatially segregated in the nasal neuroepithelium. Cell Tissue Res 276: 429 – 438, 1994.
370. Strotmann J, Wanner I, Krieger J, Raming K, Breer H. Expression
of odorant receptors in spatially restricted subsets of chemosensory neurones. Neuroreport 3: 1053–1056, 1992.
371. Sugai T, Sugitani M, Onoda N. Subdivisions of the guinea-pig
accessory olfactory bulb revealed by the combined method with
immunohistochemistry, electrophysiological, and optical recordings. Neuroscience 79: 871– 885, 1997.
372. Sugai T, Yoshimura H, Kato N, Onoda N. Component-dependent
urine responses in the rat accessory olfactory bulb. Neuroreport 17:
1663–1667, 2006.
373. Tanaka M, Treloar H, Kalb RG, Greer CA, Strittmatter SM.
G(o) protein-dependent survival of primary accessory olfactory
neurons. Proc Natl Acad Sci USA 96: 14106 –14111, 1999.
374. Taniguchi M, Wang D, Halpern M. Chemosensitive conductance
and inositol 1,4,5-trisphosphate-induced conductance in snake
vomeronasal receptor neurons. Chem Senses 25: 67–76, 2000.
375. Thom MD, Stockley P, Jury F, Ollier WE, Beynon RJ, Hurst
JL. The direct assessment of genetic heterozygosity through scent
in the mouse. Curr Biol 18: 619 – 623, 2008.
376. Thompson ML, Edwards DA. Olfactory bulb ablation and hormonally induced mating in spayed female mice. Physiol Behav 8:
1141–1146, 1972.
377. Thompson RN, Robertson BK, Napier A, Wekesa KS. Sexspecific responses to urinary chemicals by the mouse vomeronasal
organ. Chem Senses 29: 749 –754, 2004.
378. Tian H, Ma M. Molecular organization of the olfactory septal
organ. J Neurosci 24: 8383– 8390, 2004.
379. Tillman JA, Seybold SJ, Jurenka RA, Blomquist GJ. Insect
pheromones–an overview of biosynthesis and endocrine regulation. Insect Biochem Mol Biol 29: 481–514, 1999.
380. Tirindelli R, Ryba NJ. The G-protein gamma-subunit G gamma 8
is expressed in the developing axons of olfactory and vomeronasal
neurons. Eur J Neurosci 8: 2388 –2398, 1996.
381. Treloar HB, Feinstein P, Mombaerts P, Greer CA. Specificity
of glomerular targeting by olfactory sensory axons. J Neurosci 22:
2469 –2477, 2002.
382. Trinh K, Storm DR. Vomeronasal organ detects odorants in absence of signaling through main olfactory epithelium. Nat Neurosci
6: 519 –525, 2003.
382a.Trotier D, Doving KB, Ore K, Shalchian-Tabrizi C. Scanning
electron microscopy and gramicidin patch calmp recordings of
microvillous receptor neurons dissociated from the rat vomeronasal organ. Chem Senses 23: 49 –57, 1998.
383. Trotier D, Eloit C, Wassef M, Talmain G, Bensimon JL, Doving KB, Ferrand J. The vomeronasal cavity in adult humans.
Chem Senses 25: 369 –380, 2000.
384. Ukhanov K, Leinders-Zufall T, Zufall F. Patch-clamp analysis of
gene-targeted vomeronasal neurons expressing a defined V1r or
V2r receptor: ionic mechanisms underlying persistent firing. J Neurophysiol 98: 2357–2369, 2007.
385. Urban NN, Sakmann B. Reciprocal intraglomerular excitation
and intra- and interglomerular lateral inhibition between mouse
olfactory bulb mitral cells. J Physiol 542: 355–367, 2002.
386. Valenta JG, Rigby MK. Discrimination of the odor of stressed
rats. Science 161: 599 – 601, 1968.
387. Van Der LEE, Boot LM. Spontaneous pseudopregnancy in mice.
Acta Physiol Pharmacol Neerl 4: 442– 444, 1955.
388. Vandenbergh JG. Male odor accelerates female sexual maturation
in mice. Endocrinology 84: 658 – 660, 1969.
389. Vassar R, Ngai J, Axel R. Spatial segregation of odorant receptor
expression in the mammalian olfactory epithelium. Cell 74: 309 –
318, 1993.
389a.Venderbergh JG. Pheromones and mammalian reproduction. In:
The Physiology and Reproduction, edited by E. Knobil and J. D.
Neil. New York: Raven, 1994, p. 343–356.
390. Venkatachalam K, Montell C. TRP channels. Annu Rev Biochem
76: 387– 417, 2007.
391. Verberne G. Chemocommunication among domestic cats, mediated by the olfactory and vomeronasal senses. II. The relation
between the function of Jacobson’s organ (vomeronasal organ) and
Flehmen behaviour. Z Tierpsychol 42: 113–128, 1976.
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
346. Shi P, Zhang J. Comparative genomic analysis identifies an evolutionary shift of vomeronasal receptor gene repertoires in the
vertebrate transition from water to land. Genome Res 17: 166 –174,
2007.
347. Shimazaki R, Boccaccio A, Mazzatenta A, Pinato G, Migliore
M, Menini A. Electrophysiological properties and modeling of
murine vomeronasal sensory neurons in acute slice preparations.
Chem Senses 31: 425– 435, 2006.
348. Shinohara H, Kato K, Asano T. Differential localization of G
proteins, Gi and Go, in the olfactory epithelium and the main
olfactory bulb of the rat. Acta Anat 144: 167–171, 1992.
349. Shipley MT, Ennis M. Functional organization of olfactory system. J Neurobiol 30: 123–176, 1996.
350. Shirokova E, Raguse JD, Meyerhof W, Krautwurst D. The
human vomeronasal type-1 receptor family: detection of volatiles
and cAMP signaling in HeLa/Olf cells. FASEB J 22: 1416 –1425,
2008.
351. Sicard G, Holley A. Receptor cell responses to odorants: similarities and differences among odorants. Brain Res 292: 283–296,
1984.
352. Signoret JP. Reproductive behaviour of pigs. J Reprod Fertil
Suppl 11: Suppl, 1970.
353. Silverman AJ, Livne I, Witkin JW. The gonadotropin-releasing
hormone (GnRH), neuronal systems: immunocytochemistry and in
situ hybridization. In: The Physiology of Reproduction, edited by
Neill JD. New York: Raven, 1994, p. 1683–1709.
354. Silvotti L, Giannini G, Tirindelli R. The vomeronasal receptor
V2R2 does not require escort molecules for expression in heterologous systems. Chem Senses 30: 1– 8, 2005.
355. Silvotti L, Moiani A, Gatti R, Tirindelli R. Combinatorial coexpression of pheromone receptors, V2Rs. J Neurochem 103: 1753–
1763, 2007.
356. Singer AG, Macrides F, Clancy AN, Agosta WC. Purification
and analysis of a proteinaceous aphrodisiac pheromone from hamster vaginal discharge. J Biol Chem 261: 13323–13326, 1986.
357. Singh PB. Chemosensation and genetic individuality. Reproduction 121: 529 –539, 2001.
358. Singh PB, Brown RE, Roser B. MHC antigens in urine as olfactory recognition cues. Nature 327: 161–164, 1987.
359. Smith TD, Bhatnagar KP, Dennis JC, Morrison EE, Park TJ.
Growth-deficient vomeronasal organs in the naked mole-rat (Heterocephalus glaber). Brain Res 1132: 78 – 83, 2007.
360. Smith TE, Faulkes CG, Abbott DH. Combined olfactory contact
with the parent colony and direct contact with nonbreeding animals does not maintain suppression of ovulation in female naked
mole-rats (Heterocephalus glaber). Horm Behav 31: 277–288, 1997.
361. Snoddy AM, Heckathorn D, Tessel RE. Cold-restraint stress and
urinary endogenous beta-phenylethylamine excretion in rats. Pharmacol Biochem Behav 22: 497–500, 1985.
362. Spehr J, Hagendorf S, Weiss J, Spehr M, Leinders-Zufall T,
Zufall F. Ca2⫹-calmodulin feedback mediates sensory adaptation
and inhibits pheromone-sensitive ion channels in the vomeronasal
organ. J Neurosci 29: 2125–2135, 2009.
363. Spehr M, Hatt H, Wetzel CH. Arachidonic acid plays a role in rat
vomeronasal signal transduction. J Neurosci 22: 8429 – 8437, 2002.
364. Spehr M, Kelliher KR, Li XH, Boehm T, Leinders-Zufall T,
Zufall F. Essential role of the main olfactory system in social
recognition of major histocompatibility complex peptide ligands.
J Neurosci 26: 1961–1970, 2006.
365. Steel E, Keverne EB. Effect of female odour on male hamsters
mediated by the vomeronasal organ. Physiol Behav 35: 195–200,
1985.
366. Stensaas LJ, Lavker RM, Monti-Bloch L, Grosser BI, Berliner
DL. Ultrastructure of the human vomeronasal organ. J Steroid
Biochem Mol Biol 39: 553–560, 1991.
367. Stern K, McClintock MK. Regulation of ovulation by human
pheromones. Nature 392: 177–179, 1998.
368. Stowers L, Holy TE, Meister M, Dulac C, Koentges G. Loss of
sex discrimination and male-male aggression in mice deficient for
TRP2. Science 295: 1493–1500, 2002.
369. Strotmann J, Wanner I, Helfrich T, Beck A, Meinken C, Kubick S, Breer H. Olfactory neurones expressing distinct odorant
955
956
TIRINDELLI ET AL.
Physiol Rev • VOL
409. Xu F, Schaefer M, Kida I, Schafer J, Liu N, Rothman DL,
Hyder F, Restrepo D, Shepherd GM. Simultaneous activation of
mouse main and accessory olfactory bulbs by odors or pheromones. J Comp Neurol 489: 491–500, 2005.
410. Yamaguchi T, Inamura K, Kashiwayanagi M. Increases in Fosimmunoreactivity after exposure to a combination of two male
urinary components in the accessory olfactory bulb of the female
rat. Brain Res 876: 211–214, 2000.
411. Yamazaki K, Beauchamp GK. Genetic basis for MHC-dependent
mate choice. Adv Genet 59: 129 –145, 2007.
412. Yildirim E, Birnbaumer L. TRPC2: molecular biology and functional importance. Handb Exp Pharmacol 53–75, 2007.
413. Yoon H, Enquist LW, Dulac C. Olfactory inputs to hypothalamic
neurons controlling reproduction and fertility. Cell 123: 669 – 682, 2005.
414. Yoshihara Y, Kawasaki M, Tamada A, Fujita H, Hayashi H,
Kagamiyama H, Mori K. OCAM: a new member of the neural cell
adhesion molecule family related to zone-to-zone projection of
olfactory and vomeronasal axons. J Neurosci 17: 5830 –5842, 1997.
415. Young JM, Kambere M, Trask BJ, Lane RP. Divergent V1R
repertoires in five species: amplification in rodents, decimation in
primates, and a surprisingly small repertoire in dogs. Genome Res
15: 231–240, 2005.
416. Young JM, Waters H, Dong C, Fulle HJ, Liman ER. Degeneration of the olfactory guanylyl cyclase D gene during primate evolution. PLoS ONE 2: e884, 2007.
417. Yu WP, Grunwald ME, Yau KW. Molecular cloning, functional
expression and chromosomal localization of a human homolog of
the cyclic nucleotide-gated ion channel of retinal cone photoreceptors. FEBS Lett 393: 211–215, 1996.
418. Zalaquett C, Thiessen D. The effects of odors from stressed mice
on conspecific behavior. Physiol Behav 50: 221–227, 1991.
419. Zancanaro C, Caretta CM, Bolner A, Sbarbati A, Nordera GP,
Osculati F. Biogenic amines in the vomeronasal organ. Chem
Senses 22: 439 – 445, 1997.
420. Zhang J, Webb DM. Evolutionary deterioration of the vomeronasal pheromone transduction pathway in catarrhine primates. Proc
Natl Acad Sci USA 100: 8337– 8341, 2003.
421. Zhang P, Yang C, Delay RJ. Urine stimulation activates BK channels in mouse vomeronasal neurons. J Neurophysiol 100: 1824 –
1834, 2008.
422. Zhang X, Rodriguez I, Mombaerts P, Firestein S. Odorant and
vomeronasal receptor genes in two mouse genome assemblies.
Genomics 83: 802– 811, 2004.
423. Zhang Y, Hoon MA, Chandrashekar J, Mueller KL, Cook B,
Wu D, Zuker CS, Ryba NJ. Coding of sweet, bitter, and umami
tastes: different receptor cells sharing similar signaling pathways.
Cell 112: 293–301, 2003.
424. Zhou A, Moss RL. Effect of urine-derived compounds on cAMP
accumulation in mouse vomeronasal cells. Neuroreport 8: 2173–
2177, 1997.
425. Zucchi R, Chiellini G, Scanlan TS, Grandy DK. Trace amineassociated receptors and their ligands. Br J Pharmacol 149: 967–
978, 2006.
89 • JULY 2009 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.5 on October 13, 2016
392. Von Campenhausen H, Mori K. Convergence of segregated
pheromonal pathways from the accessory olfactory bulb to the
cortex in the mouse. Eur J Neurosci 12: 33– 46, 2000.
393. Wagner S, Gresser AL, Torello AT, Dulac C. A multireceptor
genetic approach uncovers an ordered integration of VNO sensory
inputs in the accessory olfactory bulb. Neuron 50: 697–709, 2006.
394. Wakabayashi Y, Mori Y, Ichikawa M, Yazaki K, Hagino-Yamagishi K. A putative pheromone receptor gene is expressed in two
distinct olfactory organs in goats. Chem Senses 27: 207–213, 2002.
395. Wang Z, Balet SC, Li V, Nudelman A, Chan GC, Storm DR.
Pheromone detection in male mice depends on signaling through
the type 3 adenylyl cyclase in the main olfactory epithelium. J Neurosci 26: 7375–7379, 2006.
396. Weiler E, Farbman AI. The septal organ of the rat during postnatal development. Chem Senses 28: 581–593, 2003.
397. Wekesa KS, Anholt RR. Pheromone regulated production of inositol-(1,4,5)-trisphosphate in the mammalian vomeronasal organ.
Endocrinology 138: 3497–3504, 1997.
398. Wekesa KS, Anholt RR. Differential expression of G proteins in
the mouse olfactory system. Brain Res 837: 117–126, 1999.
399. Wekesa KS, Lepri JJ. Removal of the vomeronasal organ reduces
reproductive performance and aggression in male prairie voles.
Chem Senses 19: 35– 45, 1994.
400. Weldon PJ. The evolution of alarm pheromones. In: Chemical
Signals in Vertebrates, edited by Muller-Schwarze D and Silverstein RM. New York: Plenum, 1983, p. 309 –312.
401. Whitten WK. Modification of the oestrous cycle of the mouse by
external stimuli associated with the male. J Endocrinol 13: 399 –
404, 1956.
402. Wilson KC, Raisman G. Age-related changes in the neurosensory
epithelium of the mouse vomeronasal organ: extended period of
postnatal growth in size and evidence for rapid cell turnover in the
adult. Brain Res 185: 103–113, 1980.
403. Witt M, Georgiewa B, Knecht M, Hummel T. On the chemosensory nature of the vomeronasal epithelium in adult humans. Histochem Cell Biol 117: 493–509, 2002.
404. Woodley SK, Cloe AL, Waters P, Baum MJ. Effects of vomeronasal organ removal on olfactory sex discrimination and odor
preferences of female ferrets. Chem Senses 29: 659 – 669, 2004.
405. Wray S, Fueshko SM, Kusano K, Gainer H. GABAergic neurons
in the embryonic olfactory pit/vomeronasal organ: maintenance of
functional GABAergic synapses in olfactory explants. Dev Biol 180:
631– 645, 1996.
406. Wyart C, Webster WW, Chen JH, Wilson SR, McClary A, Khan
RM, Sobel N. Smelling a single component of male sweat alters
levels of cortisol in women. J Neurosci 27: 1261–1265, 2007.
407. Wysocki CJ, Lepri JJ. Consequences of removing the vomeronasal organ. J Steroid Biochem Mol Biol 39: 661– 669, 1991.
408. Xia J, Sellers LA, Oxley D, Smith T, Emson P, Keverne EB.
Urinary pheromones promote ERK/Akt phosphorylation, regeneration and survival of vomeronasal (V2R) neurons. Eur J Neurosci
24: 3333–3342, 2006.