ArticlePDF Available

MHC Class I Peptides as Chemosensory Signals in the Vomeronasal Organ

Authors:

Abstract and Figures

The mammalian vomeronasal organ detects social information about gender, status, and individuality. The molecular cues carrying this information remain largely unknown. Here, we show that small peptides that serve as ligands for major histocompatibility complex (MHC) class I molecules function also as sensory stimuli for a subset of vomeronasal sensory neurons located in the basal Gao- and V2R receptor-expressing zone of the vomeronasal epithelium. In behaving mice, the same peptides function as individuality signals underlying mate recognition in the context of pregnancy block. MHC peptides constitute a previously unknown family of chemosensory stimuli by which MHC genotypic diversity can influence social behavior.
Content may be subject to copyright.
MHC Class I Peptides as
Chemosensory Signals in the
Vomeronasal Organ
Trese Leinders-Zufall,
1
Peter Brennan,
2
Patricia Widmayer,
3
Prashanth Chandramani S.,
3
Andrea Maul-Pavicic,
4
Martina Ja
¨
ger,
4
Xiao-Hong Li,
1
Heinz Breer,
3
Frank Zufall,
1
*
Thomas Boehm
4
*
The mammalian vomeronasal organ detects social information about gender,
status, and individuality. The molecular cues carrying this information remain
largely unknown. Here, we show that small peptides that serve as ligands for
major histocompatibility complex (MHC) class I molecules function also as
sensory stimuli for a subset of vomeronasal sensory neurons located in the
basal G
a
o- and V2R receptor–expressing zone of the vomeronasal epithelium.
In behaving mice, the same peptides function as individuality signals under-
lying mate recognition in the context of pregnancy block. MHC peptides
constitute a previously unknown family of chemosensory stimuli by which
MHC genotypic diversity can influence social behavior.
The mammalian vomeronasal organ (VNO) is
essential for social recognition. Vomeronasal
sensory neurons (VSNs) detect pheromones
and other chemosignals that carry informa-
tion about gender, sexual and social status,
dominance hierarchies, and individuality, but
it has been very difficult to define the molec-
ular nature of these chemosignals (1–13). The
VNO epithelium is segregated into two dis-
tinct zones, both of which express a unique
set of transduction-related molecules (2, 3,
10–12): (i) an apical (superficial) zone that
expresses the G protein Gai2 as well as mem-
bers of the V1R family of vomeronasal re-
ceptors (È150 genes) and (ii) a basal (deep)
zone that characteristically contains VSNs
that express Gao and members of the V2R
receptor family (9150 genes). The few mol-
ecules that have been identified as sensory
stimuli thus far are all small, urine-derived
volatiles that activate VSN subpopulations in
the apical zone (4, 8, 12, 13). Stimuli for
VSNs in the basal zone have not yet been
found, nor has it been possible to identify any
nonvolatile molecules that are widely assumed
to be detected by the VSNs (1–3, 9, 12).
We hypothesized that peptide ligands of
the major histocompatibility complex (MHC)
class I molecules, in addition to their well-
established role in the immune system (14),
may function as sensory stimuli for VSNs.
MHC peptides are excellent candidates for
social recognition signals that convey infor-
mation about genetic individuality. The poly-
morphisms of MHC molecules directly
translate into structurally diverse peptide-
binding grooves, such that different MHC
molecules bind different peptides (14). Hence,
the structures of peptide ligands mirror the
structures of MHC molecules and thus pro-
vide a unique molecular signature for each
individual. When peptide/MHC complexes
are not retained at the cell surface but are
instead released into the extracellular space
and appear in the urine and other bodily se-
cretions (15), any information contained in
their chemical complexity becomes a property
of the entire individual and potentially can be
used for interindividual communication (16).
By using an intact VNO preparation to
record extracellular field potentials from the
microvillous surface of the sensory epitheli-
um (4), we tested whether known ligands of
MHC class I molecules (14) Etable S1 and
Supporting Online Material (SOM) Text^ elic-
it electrical responses in VSNs of C57BL/6
mice (17). These ligands were chosen to cor-
respond to prototypical representatives for
two disparate H-2 haplotypes, namely
AAPDNRETF (for the H-2
b
haplotype of
C57BL/6 mice) and SYFPEITHI (for the
unrelated H-2
d
haplotype of BALB/c mice)
(14, 18). Both peptides evoked negative field
potentials in a dose-dependent manner in the
VNOs of female C57BL/6 mice carrying the
H-2
b
haplotype, indicating that cognate MHC
class I molecules are not required for this
response. Threshold responses were observed
with concentrations below 10
j11
Mand
10
j12
M, respectively (Fig. 1, A to D).
Control peptides in which the characteristic
anchor residues of the two MHC class I
ligands were replaced by alanines (i.e.,
AAPDARETA and SAFPEITHA, respective-
ly) failed to activate VSNs at all concen-
trations tested (up to 10
j7
M) (Fig. 1, E and
F), indicating that certain structural features
of peptides may be required for VSN ac-
tivation and ruling out an involvement of
trace by-products from peptide synthesis
1
Department of Anatomy and Neurobiology, Univer-
sity of Maryland School of Medicine, Baltimore, MD
21201, USA.
2
Sub-Department of Animal Behaviour,
University of Cambridge, Cambridge CB3 8AA, UK.
3
Institut fu¨r Physiologie, Universita
¨
t Hohenheim, D-
70593 Stuttgart, Germany.
4
Department of Devel-
opmental Immunology, Max-Planck Institute of
Immunobiology, D-79108 Freiburg, Germany.
*To whom correspondence should be addressed.
E-mail: boehm@immunbio.mpg.de (T.B.); fzufa001@
umaryland.edu (F.Z.)
Fig. 1. Class I MHC
ligands induce excitato-
ry electrical responses in
mouse VNO. (A and B)
Examples of negatively
directed field potentials
and their dose depen-
dency registered in in-
tact VNO of female
C57BL/6 mice. Re-
sponses are produced
by 500-ms pulses of
peptides that were fo-
cally ejected from a
multibarrelled stimula-
tion pipette. Responses
are representative of a
total of 17 recordings in
six mice. (C and D)
Dose-response plots of
peak responses from
the two experiments
shownin(A)and(B),
respectively. Smooth
curves are fitted by the
Hill equation, with K
1/2
value and Hill coeffi-
cient of 13.3 pM and 0.9 (AAPDNRETF) and 1.2 pM and 1.0 (SYFPEITHI), respectively. (E and F)Two
control peptides (each at 10
j7
M) failed to elicit an electrical response (n 0 13). (G) The response to
SYFPEITHI and AAPDNRETF (each at 10
j11
M) is reversibly suppressed by 2-APB (50 6M, n 0 3). (H)
Onset kinetics and initial slope of the rising phase of the field potential depend on stimulus
concentration. Responses were scaled to yield the same peak amplitudes.
R EPORTS
www.sciencemag.org SCIENCE VOL 306 5 NOVEMBER 2004
1033
and purification in the registered responses.
Peptide-induced potentials were abolished
by 2-aminoethoxydiphenyl borate (2-APB)
(50 6M) (Fig. 1G), a blocker of Ca
2þ
-
permeable, diacylycerol-gated cation channels
essential for VNO transduction, in keeping
with results obtained for urine stimuli (19).
Next, we investigated the cellular logic
underlying peptide recognition and discrim-
ination in the VNO. We first used freshly dis-
sociated VSNs from a transgenic C57BL/6
mouse strain in which all mature VSNs can
be visualized on the basis of their expres-
sion of green fluorescent protein (GFP)
under the control of the regulatory sequences
of the OMP gene (20). VSNs generate an
increase in intracellular Ca
2þ
concentration
in response to chemostimulation (4, 13,
21). Cells were loaded with a Ca
2þ
in-
dicator dye, fura-2, and cellular responses
were examined optically. Transient somatic
Ca
2þ
elevations to peptide stimulations
were reproducibly detected in a subset of
GFP-positive cells (Fig. 2, A and B). A
total of 27 cells responded to stimulation
with AAPDNRETF or SYFPEITHI pep-
tides. Of these cells, 10 responded only to
the D
b
ligand, 15 only to the K
d
ligand, and
2cellsrespondedtobothpeptides.The
response to structurally different peptides is
thus specific for individual subsets of
VSNs, with only minimal overlap. We then
systematically analyzed the spatial repre-
sentation of peptide responses in large
populations of VSNs in the sensory epithe-
lium by using in situ mapping of neuronal
activity (4). Each peptide produced robust
and reproducible increases in intracellular
Ca
2þ
in specific subsets of VSNs when
tested in such slices (Fig. 2, C to K). In total,
we imaged 5183 VSNs (26 slices from 22
mice) of which 85 cells (1.6%) responded to
peptide ligands. These signals were re-
versibly abolished by 2-APB (50 6M) (Fig.
2E), in accord with the field potential
recordings of Fig. 1G. Analysis of stimulus-
response curves of single VSNs responding
to either AAPDNRETF or SYFPEITHI es-
tablished that these cells are exceptionally
sensitive detectors of MHC peptides, with
activation thresholds near or below 10
j12
M.
VSNs responding to the same peptide ex-
hibited almost identical dose-response curves
(Fig. 2, I and J). The sets of neurons acti-
Fig. 2. MHC peptides are detected by distinct populations of VSNs. (A)
Fluorescence image of freshly dissociated VSNs obtained from an OMP-GFP
mouse. White arrow indicates the cell that was analyzed in (B). Scale bar,
10 6m. (B) Waveform of somatic Ca
2þ
transients evoked by the application
of SYFPEITHI (5 10
j10
M) or KCl (100 mM). This cell did not respond to
AAPDNRETF (5 10
j10
M). The latencies between stimulus onset and
response decrease from left to right because of the design of the perfusion
apparatus. (C and D) Spatial representation of peptide-induced activity in
VNO sensory epithelium. Shown are reconstructed VSN response maps
(%F/F confocal Ca
2þ
images digitally superimposed onto a transmitted light
image of the same slice. F, fluorescence units) for AAPDNRETF (10
j12
M,
green) and SYFPEITHI (10
j12
M, red). Cells responding to both peptides are
color-coded yellow. Black arrows indicate peptide-sensitive VSNs that are
localized at the very base of the epithelium. Black boxes, regions that were
imaged in these experiments. The white box in (C) is shown at higher
magnifications in (F) to (H). Scale bar, 100 6m. (E)Ca
2þ
response to
SYFPEITHI (10
j12
M) is reversibly abolished by 2-APB (50 6M, n 0 7). (F to
H) High-resolution pseudocolor images of the relative increase in peptide-
induced Ca
2þ
fluorescence (ratio between the peak fluorescence before and
after stimulation, %F/F). In this example, AAPDNRETF (10
j12
M, green)
activated three VSNs (cell 2, 3, and 4) and SYFPEITHI (10
j12
M, red)
activated two VSNs (cell 1 and 4). Cell 4 responded to both ligands. Scale
bar, 10 6m. Dose-dependency of stimulus-induced Ca
2þ
peak responses of
12 VSNs that recognized either AAPDNRETF (I) or SYFPEITHI (J). (K)Time
course of peptide-induced Ca
2þ
responsesfromthesamecellsthatare
depicted in (F) to (H).
R EPORTS
5 NOVEMBER 2004 VOL 306 SCIENCE www.sciencemag.org
1034
vated by structurally different peptides were
largely distinct, with only a few cells re-
sponding to both peptides (Fig. 2, C, D, F
to H, and K). Of 2067 imaged VSNs that
were tested with both peptides, 25 (1.2%)
responded only to the D
b
ligand, 20 (1.0%)
only to the K
d
ligand, and 8 cells (0.4%) re-
sponded to both peptides. To support the
physiological relevance of nonvolatile pep-
tides as stimuli of VSNs, we confirmed that
nonvolatile stimuli in urine gain access to the
vomeronasal epithelium in behaving mice
(fig. S1 and SOM Text) and that VSNs ac-
tivated by synthetic peptides respond also to
urine obtained from mice of the relevant
haplotype (fig. S2).
Peptides were recognized by sparse pop-
ulations of VSNs that were widely distrib-
uted in the sensory epithelium (Fig. 2, C and
D). We noted that activated VSNs were
mostly localized to the basal half of the
epithelium. Almost one-third of these cells
(16/53 or 30%) were found at the very base
of the epithelium, close to the basal lamina
(Fig. 2, C and D, black arrows). Do peptide-
detecting neurons thus belong to those of the
basal zone? To address this question, we first
identified peptide-sensitive VSNs by in situ
Ca
2þ
mapping and then immunostained the
tissue with antibodies against Gao Especific
for the basal zone (2, 3, 10–12)^ and phospho-
diesterase PDE4A Especific for the apical
zone (22)^ (fig. S3A). All 18 peptide-sensitive
VSNs were identified as Gao-positive and
PDE4A-negative, irrespective of whether they
were located in deep or more superficial
regions of the epithelium (fig. S3, B to E).
This result was confirmed with dissociated
VSNs by using Gao antibody (23). To
demonstrate that peptide-sensitive VSNs ex-
press V2R receptors, we used an antibody that
recognizes the V2R2 receptor, which is
broadly expressed in the basal VNO layer
(24). Double-label immunohistochemistry
showed that V2R2 is coexpressed in all
Gao-positive VSNs (fig. S3, F to H). Com-
bining in situ Ca
2þ
mapping and V2R2 im-
munolabeling, we showed directly that
peptide-sensitive VSNs express V2Rs (fig.
S3, I to K) (n 0 7).
To determine whether peptide stimula-
tion leads to action potential generation in
single VSNs, we used the loose-patch tech-
nique to register extracellular spike activity
from visually identified VSNs in VNO slices
(4). At 10
j11
M, the MHC peptides elicited
excitatory, sequence-specific responses in a
subset of basal VSNs (Fig. 3A) (n 0 6),
consistent with the Ca
2þ
imaging data. We
also used whole-cell current clamp record-
ings from VSNs in slices (19) to demonstrate
directly that peptide stimulation produced a
membrane depolarization that, in turn,
evoked action potential discharges (Fig. 3B)
(n 0 5).
What are the structural constraints un-
derlying peptide discrimination? We hy-
pothesized that the peptide anchor residues
may substantially contribute to the specific
recognition by VSNs. Indeed, two different
ligands of the K
d
MHC molecule, SYFPEITHI
and SYIPSAEKI (14, 18), which share the
same anchor residues (Y and I) at positions 2
and 9, respectively, but differ substantially in
the other positions, activated the same 6
neurons (out of 1109 cells imaged), and their
stimulus-response curves were nearly identical
(Fig. 4, A to E). VSNs that recognized only
one, but not the other, peptide were not
observed. The recognition mode of such
peptides may thus at least partially resemble
that of MHC molecules.
Given that the anchor residues of peptides
appear to be essential for VSN activation and
that a goldfish V2R-like receptor is known to
recognize free amino acids (25), it was nec-
essary to rule out that the VSN responses
were caused by free amino acids. Isoleucine
(10
j12
M), the C-terminal anchor residue in
the K
d
peptides, failed to generate a Ca
2þ
signal in five of five cells that detected
SYFPEITHI and SYIPSAEKI (Fig. 4D). It
also failed to produce a response in 881 other
VSNs with unknown tuning properties.
Furthermore, in field potential recordings, a
mixture containing all amino acids (in free
form, each at 10
j11
M) that constitute the
SYIPSAEKI peptide failed to induce any re-
sponse (Fig. 4F) (n 0 11). Likewise, a scram-
Fig. 4. Structural features of VSN peptide discrimination. (A to C)Ca
2þ
responses in individual VSNs to two MHC peptides, SYFPEITHI (green) and
SYIPSAEKI (red) (each at 10
j12
M). Both cells responded to both peptides. (D)
Time courses of peptide-evoked Ca
2þ
responses from the two cells shown in
(A) to (C). The free amino acid isoleucine (10
j12
M) failed to elicit a response.
(E) Comparison of the dose-dependency of Ca
2þ
peak responses induced by
SYFPEITHI (black curve, open circles) or SYIPSAEKI (red curve, solid circles).
Each data point represents the mean T SD of at least five independent
measurements. (F) A mixture containing all the amino acids that constitute the SYIPSAEKI peptide (in free form, each at 10
j11
M) fails to elicit a VNO
field potential (11 recordings from three mice). (G) A scrambled version of the AAPDNRETF peptide, ANPRAFDTE (10
j11
M), fails to evoke a field
potential response (14 recordings from five mice).
Fig. 3. MHC peptides induce action potential
generation in individual VSNs. (A) Spontaneous
and stimulus-evoked impulse discharges in a
VSN after successive application of three dif-
ferent peptides (all at 10
j11
M). AAPDNRETF,
but not SYFPEITHI or AAPDARETA, elicited a
transient excitation in this neuron. (B) Whole-
cell current clamp recording of a VSN that
responded to AAPDNRETF (10
j11
M) with a
transient increase in the rate of action potential
firing. Resting potential was –62 mV. Arrows
indicate the time point at which peptide applica-
tion was turned on.
R EPORTS
www.sciencemag.org SCIENCE VOL 306 5 NOVEMBER 2004
1035
bled version of the D
b
ligand AAPDNRETF,
ANPRAFDTE, failed to evoke any response
(Fig. 4G) (n 0 14). Thus, peptides must meet
precise structural specifications for VSN
activation, and peptides of random sequence
are unlikely to function as ligands for the
receptors on VSNs.
Given that MHC peptides activate VSNs
in a sequence-specific manner, they could
potentially function as individuality signals
during social recognition. In mice, selective
pregnancy failure Ethe Bruce effect (26)^ rep-
resents an excellent paradigm to assess this
hypothesis in vivo, because it depends criti-
cally on signaling via the accessory olfactory
system (SOM Text) and requires the capacity
to differentiate between individuality cues
(27). Female mice of the BALB/c inbred
strain (H-2
d
haplotype) were mated with
BALB/c males and then exposed to urine
taken from either a BALB/c male (mating
male urine) or C57BL/6 male (unfamiliar
male urine; H-2
b
haplotype). Application of
the unfamiliar urine, coincident with the
postmating peaks in prolactin levels, reliably
resulted in a high level of pregnancy failure,
whereas the familiar urine did not (Fig. 5,
experiments 1 and 2). When BALB/c fe-
males were mated with C57BL/6 males,
pregnancy block occurred after the applica-
tion of BALB/c urine, but not of C57BL/6
urine (Fig. 5, experiments 3 and 4), estab-
lishing the strain specificity of pregnancy
block (28). To test whether familiar male
urine could be converted to unfamiliar urine,
we added peptides of disparate H-2 haplo-
type specificity. Exposure of BALB/c-mated
BALB/c females to a mixture of H-2
b
class I
peptides in BALB/c mating male urine was
equally effective as C57BL/6 urine (Fig. 5,
experiment 6). The addition of a mixture of
H-2
d
peptides had no effect (Fig. 5, experi-
ment 5). Conversely, exposure of C57BL/6-
mated BALB/c females to BALB/c peptides
in C57BL/6 mating male urine was effective
at blocking pregnancy; here, C57BL/6 pep-
tides were ineffective. Experiments 5 to
8 show that peptides per se do not cause
pregnancy failure and that this function
depends on the previous mating combination.
The overall occurrence of pregnancy block
upon exposure to the strange male H-2
peptides of 64% (n 0 47; combined results
of experiments 6 and 7) was significantly
higher than the 25% (n 0 24; combined re-
sults of experiments 5 and 8) elicited by ex-
posure to mating male H-2 peptides (P 0
0.002, Fisher exact probability test).
Our experiments identified an unexpected
role for MHC class I peptides as chemosen-
sory stimuli. MHC class I ligands are recog-
nized by VSNs in the basal layer of the
VNO. Recognition of peptides by VSNs is
independent of MHC haplotype, and pep-
tides specific for different MHC molecules
(i.e., carrying different anchor residues)
generate unique VSN activation patterns,
providing the basis for the neural represen-
tation of the structural diversity of this new
family of chemosignals.
Our data demonstrate that the VNO can
detect both nonvolatile and volatile stimuli, a
result that is fully compatible with early pre-
dictions (29). Peptide responses were found
exclusively in Gao- and V2R-positive neu-
rons, whereas responses to volatile stimuli
have been mapped to the apical V1R-
expressing zone (4). Whether this functional
segregation is true for all vomeronasal stim-
uli remains to be seen. V2R receptors con-
stitute a large family of orphan receptors
(2, 3, 10–12) that differ from V1Rs and
olfactory receptors by the presence of a large
N-terminal domain. They are coexpressed
with MHC class Ib molecules at the cell sur-
face of VSNs (30, 31). MHC class Ib mol-
ecules can bind peptides, but they lack the
typical peptide-binding groove and specificity
of classical MHC class I molecules (32).
Sequence-specific recognition of peptides may
thus be achieved by the N-terminal domain of
certain V2R receptors (or receptor combina-
tions), whereas the MHC class Ib molecules
may serve as a general presentation device.
Given the limited diversity of amino acid res-
idues occupying the two anchor positions of
mouse MHC class I peptides (14), we estimate
that about 50 different receptors should be suf-
ficient to discriminate ligands from all known
mouse MHC class I molecules.
Considerable work has focused on the
main olfactory system in the detection of
MHC-related odor signals (33, 34). Our
results highlight the role of the VNO in this
process but do not preclude a role of the
main olfactory system in individual recogni-
tion, nor do they preclude a role for other
molecules such as volatile urinary constitu-
ents (34) or polymorphic major urinary pro-
teins (35). MHC peptides may thus form one
class of different signals that may be used in
different behavioral contexts. For a meaning-
ful biological response to occur in MHC-
related behaviors, signals about gender,
reproductive status, and species identity must
be evaluated alongside signals of genetic
individuality, and this may involve remotely
sensed signals as well as signals that are de-
tected during direct contact. A chemosensory
function of MHC peptides provides a direct
link between MHC diversity and MHC-
related behavior, converting a MHC geno-
type into an olfactorily detectable quality.
References and Notes
1. C. J. Wysocki, M. Meredith, in Neurobiology of Taste
and Smell, T. E. Finger, W. L. Silver, Eds. (Wiley, New
York, 1987), pp. 125–150.
2. R. Tirindelli, C. Mucignat-Caretta, N. J. Ryba, Trends
Neurosci. 21, 482 (1998).
3. P. A. Brennan, E. B. Keverne, Curr. Biol. 14, R81
(2004).
4. T. Leinders-Zufall et al., Nature 405, 792 (2000).
5. T. E. Holy, C. Dulac, C. M. Meister, Science 289, 1569
(2000).
6. L.Stowers,T.E.Holy,M.Meister,C.Dulac,G.Koentges,
Science 295, 1493 (2002); published online 31 January
2002 (10.1126/science.1069259).
7. B. G. Leypold et al., Proc. Natl. Acad. Sci. U.S.A. 99,
6376 (2002).
8. K. Del Punta et al., Nature 419, 70 (2002).
9. M. Luo, M. S. Fee, L. C. Katz, Science 299, 1196 (2003).
10. C. I. Bargmann, Cell 90, 585 (1997).
11. C. Dulac, A. T. Torello, Nature Rev. Neurosci. 4,551
(2003).
12. P. Mombaerts, Nature Rev. Neurosci. 5, 263 (2004).
13. C. Boschat et al., Nature Neurosci. 5, 1261 (2002).
14. H. G. Rammensee, J. Bachmann, S. Stefanovic, MHC
Fig. 5. Peptides func-
tion as individuality
signals in the context
of pregnancy block.
Percent pregnancy fail-
ure in female BALB/c
(H-2
d
haplotype) mice
mated with either
BALB/c (H-2
d
haplo-
type) or C57BL/6 (H-
2
b
haplotype) males as
indicated (þ). Exposure
to different male urine
types, which, in some
cases, were supple-
mented with peptides
specific for BALB/c or
C57BL/6 haplotypes, is
indicated. *P G 0.05;
***P G 0.001 (Fisher
exact probability test
with Bonferroni cor-
rection for multiple
comparisons).
R EPORTS
5 NOVEMBER 2004 VOL 306 SCIENCE www.sciencemag.org
1036
Ligands and Peptide Motifs (Landes Bioscience,
Georgetown, TX, 1997).
15. P. B. Singh, R. E. Brown, B. Roser, Nature 327, 161 (1987).
16. T. Boehm, C. C. Bleul, M. Schorpp, Immunol. Rev.
195, 15 (2003).
17. Materials and methods are available as supporting
material on Science Online.
18. Single-letter abbreviations for the amino acid resi-
dues are as follows: A, Ala; C, Cys; D, Asp; E, Glu; F,
Phe; G, Gly; H, His; I, Ile; K, Lys; L, Leu; M, Met; N, Asn;
P, Pro; Q, Gln; R, Arg; S, Ser; T, Thr; V, Val; W, Trp;
and Y, Tyr.
19. P. Lucas, K. Ukhanov, T. Leinders-Zufall, F. Zufall,
Neuron 40, 551 (2003).
20. S. M. Potter et al., J. Neurosci. 21, 9713 (2001).
21. M. Spehr, H. Hatt, C. H. Wetzel, J. Neurosci. 22, 8429
(2002).
22. Y. E. Lau, J. A. Cherry, Neuroreport 11, 27 (2000).
23. T. Leinders-Zufall et al., data not shown.
24. S. Martini, L. Silvotti, A. Shirazi, N. J. Ryba, R. Tirindelli,
J. Neurosci. 21, 843 (2001).
25. D. J. Speca et al., Neuron 23, 487 (1999).
26. H. M. Bruce, Nature 184, 105 (1959).
27. A. Lloyd-Thomas, E. B. Keverne, Neuroscience 7, 907
(1982).
28. The efficiency of pregnancy block by direct applica-
tion of unfamiliar urine is similar to that caused by
the presence of unfamiliar mice (for instance, when
BALB/c females are mated with C57BL/6 males,
presence of BALB/c males causes pregnancy failure
in 9 of 12 mice).
29. R. J. O’Connell, M. Meredith, Behav. Neurosci. 98,
1083 (1984).
30. T. Ishii, J. Hirota, P. Mombaerts, Curr. Biol. 13, 394
(2003).
31. J. Loconto et al., Cell 112, 607 (2003).
32. X. L. He, P. Tabaczewski, J. Ho, I. Stroynowski, K. C.
Garcia, Structure 9, 1213 (2001).
33. D. Penn, W. Potts, Adv. Immunol. 69, 411 (1998).
34. G. K. Beauchamp, K. Yamazaki, Biochem. Soc.
Trans. 31, 147 (2003).
35. J. L. Hurst et al., Nature 414, 631 (2001).
36. We thank R. Escher for peptide synthesis, J. Cherry for
PDE4A antibodies, R. Tirindelli for V2R2 antibodies,
P. Mombaerts for OMP-GFP mice, and K. Kelliher for
the demonstration that nonvolatile chemicals can
reach the VNO in vivo. This work was supported by
the Leibniz program of the Deutsche Forschungsge-
meinschaft (T.B. and H.B.), by NIH/National Institute
on Deafness and other Communication Disorders
(T.L.-Z. and F.Z.), and by the Hochshul- und
Wissenschafts-Programm (P.W.).
Supporting Online Material
www.sciencemag.org/cgi/content/full/306/5698/1033/
DC1
Materials and Methods
SOM Text
Figs. S1 to S4
Table S1
References
15 July 2004; accepted 9 September 2004
Autophagy Defends Cells Against
Invading Group A Streptococcus
Ichiro Nakagawa,
1,3
*
Atsuo Amano,
2,4
Noboru Mizushima,
3,5
Akitsugu Yamamoto,
6
Hitomi Yamaguchi,
7
Takahiro Kamimoto,
7
Atsuki Nara,
6,7
Junko Funao,
1
Masanobu Nakata,
1
Kayoko Tsuda,
7
Shigeyuki Hamada,
1
Tamotsu Yoshimori
4,7
*
We found that the autophagic machinery could effectively eliminate
pathogenic group A Streptococcus (GAS) within nonphagocytic cells. After
escaping from endosomes into the cytoplasm, GAS became enveloped by
autophagosome-like compartments and were killed upon fusion of these
compartments with lysosomes. In autophagy-deficient Atg5
j/j
cells, GAS
survived, multiplied, and were released from the cells. Thus, the autophagic
machinery can act as an innate defense system against invading pathogens.
Autophagy mediates the bulk degradation of
cytoplasmic components in eukaryotic cells
in which a portion of the cytoplasm is
sequestered in an autophagosome and even-
tually degraded upon fusion with lysosomes
(1–3). Streptococcus pyogenes (also known
as group A Streptococcus, GAS) is the etio-
logical agent for a diverse collection of
human diseases (4). GAS invades nonphago-
cytic cells (5, 6), but the destination of GAS
after internalization is not well understood.
To clarify the intracellular fate of GAS,
especially any possible involvement of auto-
phagy, we first investigated whether intra-
cellular GAS colocalizes with LC3, an
autophagosome-specific membrane marker,
following invasion of HeLa cells (7–9).
After infection, GAS strain JRS4 cells colo-
calized with LC3-positive vacuole-like struc-
tures in HeLa cells (Fig. 1A). The size (5 to
10 6m) and morphology of the structures were
distinct from standard starvation-induced
autophagosomes with a diameter of about
1 6m (fig. S1A), so we designated these
structures GAS-containing LC3-positive
autophagosome-like vacuoles (GcAVs). The
number of cells bearing GcAVs, the area of
GcAVs, and ratio of GAS trapped in GcAVs
to total intracellular GAS increased in a time-
dependent manner, reaching a maximum at
3 hours after infection (Fig. 1, B and C;
figs. S1B and S2A). A similar result was ob-
tained in mouse embryonic stem (ES) cells
(figs. S2B and S3A). About 80% of intracel-
lular GAS were eventually trapped by the
compartments (Fig. 1C; fig. S1B). LC3 fre-
quently surrounded GAS, fitting closely around
a GAS chain (Fig. 1, D and E; movie S1).
LC3 exists in two molecular forms. LC3-I
(18 kD) is cytosolic, whereas LC3-II (16 kD)
binds to autophagosomes (7, 8). The amount
of LC3-II, which directly correlates with the
number of autophagosomes (8), increased
after infection (Fig. 1F). Thus, GAS invasion
appears to induce autophagy, specifically
trapping intracellular GAS.
To substantiate this idea, we examined
GcAV formation in Atg5-deficient (Atg5
j/j
)
cells lacking autophagosome formation (7).
In contrast to the wild-type cells (fig. S2, B
and C), no GcAVs were observed in Atg5
j/j
ES cells (J1-2) (Fig. 2A) or in Atg5
j/j
mouse embryonic fibroblasts (MEFs) (fig.
S2C). Thus, GcAV formation requires an
Atg5-mediated mechanism. We also exam-
ined LC3-II formation. During infection with
GAS, Atg5
j/j
cells showed no induction of
LC3-II (Fig. 2B). By electron microscopy, in
wild-type MEF cells infected with GAS, we
observed characteristic cisternae surrounding
GAS in the cytoplasm (Fig. 2C). No GAS
were found surrounded by the membranes in
Atg5
j/j
cells (Fig. 2C). The autophagosome-
like multiple membrane–bound compartment
containing GAS was also found in HeLa cells
(Fig. 2D).
Next, we asked whether the bacteria are
killed or survive after entering the compart-
ments. To address this question, we directly
scored bacterial viability by counting colony-
forming units (CFU viability assay) in wild-type
and Atg5
j/j
MEFs (Fig. 2E). In wild-type
MEFs at 4 hours after infection, intracellular
GAS had been killed (Fig. 2E), whereas the
decrease of GAS viability was suppressed
in the Atg5
j/j
MEFs. Tannic acid is a cell-
impermeable fixative that prevents fusion
between secretory vesicles and the plasma
membrane but does not affect intracellular
membrane trafficking (10). In Atg5
j/j
cells treated with tannic acid to prevent
external escape of GAS, the viable bacteria
increased by 2 hours after infection and
maintained this level at 4 hours after in-
fection (Fig. 2E). In contrast, the numbers
of intracellular GAS decreased rapidly in
tannic acid–treated wild-type cells as well
1
Department of Oral and Molecular Microbiology,
2
Department of Oral Frontier Biology, Osaka Univer-
sity Graduate School of Dentistry, 1-8 Yamadaoka,
Suita-Osaka 565-0871, Japan.
3
PRESTO,
4
CREST,
Japan Science and Technology Agency, Kawaguchi-
Saitama 332-0012, Japan.
5
Department of Bioregu-
lation and Metabolism, The Tokyo Metropolitan
Institute of Medical Science, 3-18-22 Honkomagome,
Bunkyo-ku, Tokyo 113-8613, Japan.
6
Department of
Cell Biology, Faculty of Bio-Science, Nagahama
Institute of Bio-Science and Technology, 1266
Tamura-cho, Nagahama-Shiga 526-0829, Japan.
7
De-
partment of Cell Genetics, National Institute of
Genetics/SOKENDAI, Yata 1111, Mishima-Shizuoka
411-8540, Japan.
*To whom correspondence should be addressed.
E-mail: ichiro@dent.osaka-u.ac.jp and tamyoshi@lab.
nig.ac.jp
R EPORTS
www.sciencemag.org SCIENCE VOL 306 5 NOVEMBER 2004
1037
... By far, the most widely studied secretion in animal chemosensory research is urine (see Mohrhardt et al., 2018, and references therein), which is a rich source of semiochemicals that serves a well-established function in social communication. While we still lack a comprehensive molecular description of this broadband vomeronasal stimulus, previous work has identified several putative semiochemicals in mouse urine and other bodily secretions, which activate VSNs and cover many structural groups and feature dimensions (Chamero et al., 2007;Doyle et al., 2016;Fu et al., 2015;Hurst et al., 2001;Kimoto et al., 2005;Leinders-Zufall et al., 2004;Leinders-Zufall et al., 2000;Nodari et al., 2008;Novotny, 2003;Overath et al., 2014;Rivière et al., 2009;Röck et al., 2006;Sturm et al., 2013;Wyatt, 2017). Prominent molecularly identified VSN stimuli include various sulfated steroids (Celsi et al., 2012;Fu et al., 2015;Haga-Yamanaka et al., 2015;Haga-Yamanaka et al., 2014;Isogai et al., 2011;Nodari et al., 2008;Turaga and Holy, 2012), which could reflect the dynamic endocrine state of an individual. ...
Article
Full-text available
In most mammals, conspecific chemosensory communication relies on semiochemical release within complex bodily secretions and subsequent stimulus detection by the vomeronasal organ (VNO). Urine, a rich source of ethologically relevant chemosignals, conveys detailed information about sex, social hierarchy, health, and reproductive state, which becomes accessible to a conspecific via vomeronasal sampling. So far, however, numerous aspects of social chemosignaling along the vomeronasal pathway remain unclear. Moreover, since virtually all research on vomeronasal physiology is based on secretions derived from inbred laboratory mice, it remains uncertain whether such stimuli provide a true representation of potentially more relevant cues found in the wild. Here, we combine a robust low-noise VNO activity assay with comparative molecular profiling of sex- and strain-specific mouse urine samples from two inbred laboratory strains as well as from wild mice. With comprehensive molecular portraits of these secretions, VNO activity analysis now enables us to (i) assess whether and, if so, how much sex/strain-selective ‘raw’ chemical information in urine is accessible via vomeronasal sampling; (ii) identify which chemicals exhibit sufficient discriminatory power to signal an animal’s sex, strain, or both; (iii) determine the extent to which wild mouse secretions are unique; and (iv) analyze whether vomeronasal response profiles differ between strains. We report both sex- and, in particular, strain-selective VNO representations of chemical information. Within the urinary ‘secretome’, both volatile compounds and proteins exhibit sufficient discriminative power to provide sex- and strain-specific molecular fingerprints. While total protein amount is substantially enriched in male urine, females secrete a larger variety at overall comparatively low concentrations. Surprisingly, the molecular spectrum of wild mouse urine does not dramatically exceed that of inbred strains. Finally, vomeronasal response profiles differ between C57BL/6 and BALB/c animals, with particularly disparate representations of female semiochemicals.
... GABA B receptors have been implicated in aversive learning in the MOB (Okutani et al., 2003;Bhattarai et al., 2020). The vomeronasal system is known to initiate a variety of innate, pheromone-evoked behavioral responses (Mohrhardt et al., 2018) as well as specific forms of learned, experience-dependent behaviors (Brennan et al., 1990;Leinders-Zufall et al., 2004;Brennan and Zufall, 2006;Roberts et al., 2012;Hattori et al., 2017;Kaba et al., 2020;Trouillet et al., 2021). Modulation of synaptic transmission by GABA B receptors at VSN terminals could potentially contribute to memory formation in the AOB. ...
Article
Full-text available
Vomeronasal sensory neurons (VSNs) recognize pheromonal and kairomonal semiochemicals in the lumen of the vomeronasal organ. VSNs send their axons along the vomeronasal nerve (VN) into multiple glomeruli of the accessory olfactory bulb (AOB) and form glutamatergic synapses with apical dendrites of mitral cells, the projection neurons of the AOB. Juxtaglomerular interneurons release the inhibitory neurotransmitter γ-aminobutyric acid (GABA). Besides ionotropic GABA receptors, the metabotropic GABAB receptor has been shown to modulate synaptic transmission in the main olfactory system. Here we show that GABAB receptors are expressed in the AOB and are primarily located at VN terminals. Electrical stimulation of the VN provokes calcium elevations in VSN nerve terminals, and activation of GABAB receptors by the agonist baclofen abolishes calcium influx in AOB slice preparations. Patch clamp recordings reveal that synaptic transmission from the VN to mitral cells can be completely suppressed by activation of GABAB receptors. A potent GABAB receptor antagonist, CGP 52432, reversed the baclofen-induced effects. These results indicate that modulation of VSNs via activation of GABAB receptors affects calcium influx and glutamate release at presynaptic terminals and likely balances synaptic transmission at the first synapse of the accessory olfactory system.
... In rodents, olfaction is largely involved in social communication between conspecifics (Wyatt 2014). This olfactory communication occurs via chemosignals of diverse chemical nature, including small volatile organic compounds (VOCs), steroids, peptides and proteins such as lipocalins (Novotny et al. 1985;Leinders-Zufall et al. 2004;Kimoto et al. 2005;Chamero et al. 2007;Liberles 2014). These chemosignals are usually secreted in urine, saliva, tears or fluids from exocrine glands (Johnston 2003). ...
Article
Full-text available
In mammals, especially rodents, social behaviours, such as parenting, territoriality or mate attraction, are largely based on olfactory communication through chemosignals. These behaviours are mediated by species-specific chemosignals, including small organic molecules and proteins that are secreted in the urine or in various fluids from exocrine glands. Chemosignal detection is mainly ensured by olfactory neurons in two specific sensory organs, the vomeronasal organ (VNO) and the main olfactory epithelium (MOE). This study aimed to characterise the olfactory communication in the fossorial ecotype of the water voles, Arvicola terrestris. We first measured the olfactory investigation of urine and lateral scent gland secretions from conspecifics. Our results showed that water voles can discriminate the sex of conspecifics based on the smell of urine, and that urinary male odour is attractive for female voles. Then, we demonstrated the ability of the VNO and MOE to detect volatile organic compounds (VOCs) found in water vole secretions using live-cell calcium imaging in dissociated cells. Finally, we evaluated the attractiveness of two mixtures of VOCs from urine or lateral scent glands in the field during a cyclical outbreak of vole populations.
Preprint
Full-text available
In birds and insects, females uptake sperm for a specific duration post-copulation known as the ejaculate holding period (EHP) before expelling unused sperm and the mating plug through sperm ejection. Our study uncovered that encountering males or mated females after mating substantially shortens EHP, a phenomenon we term ‘ m ale-induced E HP s hortening (MIES)’. MIES requires Or47b+ olfactory and ppk23+ gustatory neurons, activated by 2-methyltetracosane and 7-Tricosene, respectively. These odorants raise cAMP levels in pC1 neurons, responsible for processing male courtship and regulating female mating receptivity. Elevated cAMP levels in pC1 neurons reduce EHP and reinstate their responsiveness to male courtship cues, promoting re-mating with faster sperm ejection. This study establishes MIES as a genetically tractable model of sexual plasticity with a conserved neural mechanism.Sexual plasticity, adapting reproductive behaviors to social changes, was explored in the fruit fly, a genetically tractable model insect. Findings revealed that inseminated females, encountering another courting male post-mating, shorten the ejaculate holding period (EHP). Specific olfactory and gustatory pathways regulating this phenomenon were identified, converging on the pC1 neurons in the brain-a conserved neural circuit regulating female mating activity. Odors associated with EHP shortening increased the second messenger cAMP. The elevated cAMP transiently heightened the excitability of pC1 neurons, enabling inseminated females to promptly remove the male ejaculate and engage in the subsequent mating more readily. This study establishes a behavioral model for sexual plasticity and provide a framework for understanding the involved neural processes.
Article
Sensory signals detected by olfactory sensory organs are critical regulators of animal behavior. An accessory olfactory organ, the vomeronasal organ, detects cues from other animals and plays a pivotal role in intra‐ and inter‐species interactions in mice. However, how ethologically relevant cues control mouse behavior through approximately 350 vomeronasal sensory receptor proteins largely remains elusive. The type 2 vomeronasal receptor‐A4 (V2R‐A4) subfamily members have been repeatedly detected from vomeronasal sensory neurons responsive to predator cues, suggesting a potential role of this receptor subfamily as a sensor for predators. This review focuses on this intriguing subfamily, delving into its receptor functions and genetic characteristics.
Article
The olfactory mucosa is a component of the nasal airway that mediates the sense of smell. Recent studies point to an important role for the olfactory mucosa as a barrier to both respiratory pathogens and to neuroinvasive pathogens that hijack the olfactory nerve and invade the CNS. In particular, the COVID-19 pandemic has demonstrated that the olfactory mucosa is an integral part of a heterogeneous nasal mucosal barrier critical to upper airway immunity. However, our insufficient knowledge of olfactory mucosal immunity hinders attempts to protect this tissue from infection and other diseases. This Review summarizes the state of olfactory immunology by highlighting the unique immunologically relevant anatomy of the olfactory mucosa, describing what is known of olfactory immune cells, and considering the impact of common infectious diseases and inflammatory disorders at this site. We will offer our perspective on the future of the field and the many unresolved questions pertaining to olfactory immunity.
Preprint
Full-text available
In birds and insects, females uptake sperm for a specific duration post-copulation known as the ejaculate holding period (EHP) before expelling unused sperm and the mating plug through sperm ejection. Our study uncovered that encountering males or mated females after mating substantially shortens EHP, a phenomenon we term ‘ m ale-induced E HP s hortening (MIES)’. MIES requires Or47b+ olfactory and ppk23+ gustatory neurons, activated by 2-methyltetracosane and 7-Tricosene, respectively. These odorants raise cAMP levels in pC1 neurons, responsible for processing male courtship and regulating female mating receptivity. Elevated cAMP levels in pC1 neurons reduce EHP and reinstate their responsiveness to male courtship cues, promoting re-mating with faster sperm ejection. This study establishes MIES as a genetically tractable model of sexual plasticity with a conserved neural mechanism. Significance Statement Sexual plasticity, adapting reproductive behaviors to social changes, was explored in the fruit fly, a genetically tractable model insect. Findings revealed that inseminated females, encountering another courting male post-mating, shorten the ejaculate holding period (EHP). Specific olfactory and gustatory pathways regulating this phenomenon were identified, converging on the pC1 neurons in the brain-a conserved neural circuit regulating female mating activity. Odors associated with EHP shortening increased the second messenger cAMP. The elevated cAMP transiently heightened the excitability of pC1 neurons, enabling inseminated females to promptly remove the male ejaculate and engage in the subsequent mating more readily. This study establishes a behavioral model for sexual plasticity and provide a framework for understanding the involved neural processes.
Article
Full-text available
This review addresses the role of chemical communication in mammals, giving special attention to the vomeronasal system in pheromone-mediated interactions. The vomeronasal system influences many social and sexual behaviors, from reproduction to species recognition. Interestingly, this system shows greater evolutionary variability compared to the olfactory system, emphasizing its complex nature and the need for thorough research. The discussion starts with foundational concepts of chemocommunication, progressing to a detailed exploration of olfactory systems. The neuroanatomy of the vomeronasal system stands in contrast with that of the olfactory system. Further, the sensory part of the vomeronasal system, known as the vomeronasal organ, and the integration center of this information, called the accessory olfactory bulb, receive comprehensive coverage. Secondary projections of both the olfactory and vomeronasal systems receive attention, especially in relation to the dual olfactory hypothesis. The review concludes by examining the organization of the vomeronasal system in four distinct mammalian groups: rodents, marsupials, herpestids, and bovids. The aim is to highlight the unique morphofunctional differences resulting from the adaptive changes each group experienced.
Chapter
In mammals, reproductive function can be modulated by sociosexual factors. In small ruminants such as goats, phenomena such as puberty acceleration and the “male effect” have been hypothesized to depend on male olfactory information, but the source of these odorants and how they are detected is currently unclear. Calcium (Ca2+) imaging is a widely developed approach in rodents to assess the activity of olfactory sensory neurons (OSNs) and vomeronasal neurons (VSNs), originating from the main olfactory epithelium (MOE) and the vomeronasal organ (VNO), respectively. Our study aimed to adapt this approach to primary cultures of dissociated goat OSNs and VSNs. MOE and VNO cells were stimulated with urine and fur extracts from castrated or sexually active bucks. Ca2+ responses were observed in cells from the two olfactory systems for each stimuli. These results indicate that (1) both MOE and VNO are involved in male chemosignal detection by females during the breeding season, and (2) VSNs and OSNs from ovulatory goats are activated by male olfactory information, independently of male sexual activity. The development of this approach will be of high interest in the search for potential chemosignals involved in the regulation of sociosexual behavior in goats and associated physiological regulations.
Preprint
Full-text available
In most mammals, conspecific chemosensory communication relies on semiochemical release within complex bodily secretions and subsequent stimulus detection by the vomeronasal organ (VNO). Urine, a rich source of ethologically relevant chemosignals, conveys detailed information about sex, social hierarchy, health and reproductive state, which becomes accessible to a conspecific via vomeronasal sampling. So far, however, numerous aspects of social chemosignaling along the vomeronasal pathway remain unclear. Moreover, since virtually all research on vomeronasal physiology is based on secretions derived from inbred laboratory mice, it remains uncertain whether such stimuli provide a true representation of potentially more relevant cues found in the wild. Here, we combine a robust low-noise VNO activity assay with comparative molecular profiling of sex- and strain-specific mouse urine samples from two inbred laboratory strains as well as from wild mice. With comprehensive molecular portraits of these secretions, VNO activity analysis now enables us to (i) assess whether and, if so, how much sex-/ strain-selective “raw” chemical information in urine is accessible via vomeronasal sampling; (ii) identify which chemicals exhibit sufficient discriminatory power to signal an animal’s sex, strain, or both; (iii) determine the extent to which wild mouse secretions are unique; and (iv) analyze whether vomeronasal response profiles differ between strains. We report both sex- and, in particular, strain-selective VNO representations of chemical information. Within the urinary ‘secretome’, both volatile compounds and proteins exhibit sufficient discriminative power to provide sex- and strain-specific molecular fingerprints. While total protein amount is substantially enriched in male urine, females secrete a larger variety at overall comparatively low concentrations. Surprisingly, the molecular spectrum of wild mouse urine does not dramatically exceed that of inbred strains. Finally, vomeronasal response profiles differ between C57BL/6 and BALB/c animals, with particularly disparate representations of female semiochemicals.
Article
Full-text available
Estrous hamster vaginal discharge (HVD) contains both volatile and nonvolatile chemical signals which collectively elicit both male attraction to females and male mating behavior. These two aspects of normal sexual behavior are differentially affected by lesions involving afferents of the main and the accessory olfactory systems. The results of lesions that involve the main olfactory system suggest that it provides information primarily concerning those volatile components of HVD that normally signal the presence of a female. Lesions restricted to the accessory olfactory system do not impair a male's interest in or relative preference for HVD but do significantly interfere with subsequent stages of sexual behavior by reducing the amount of mating obtained upon exposure to HVD. Stimulation of the accessory olfactory system by nonvolatile components of HVD has thus been implicated in the production of mating behavior. In these experiments, male hamsters were attracted to female odor and engaged in significant amounts of mating behavior with surrogate females when only the volatile components of HVD were available to them. These behaviors were further enhanced when both volatile and nonvolatile components of HVD were provided. Lesion studies of the afferents involved reinforce the hypothesis that the main olfactory system is preferentially involved with processing those volatile chemical signals in HVD that denote female attractiveness whereas the accessory olfactory system is preferentially involved with processing those chemical signals, both volatile and nonvolatile, that evoke subsequent steps in male sexual behavior.
Article
Distribution of the cAMP-specific phosphodiesterase PDE4A was examined in the accessory olfactory system by immunohistochemistry. Adjacent sections through the vomeronasal organ (VNO) and accessory olfactory bulb (AOB) were alternately immunostained with antibodies against PDE4A or the G-protein a subunit G(o alpha), which labels basal VNO neurons, in order to determine whether PDE4A occurs preferentially in one of two segregated VNO pathways. We found that PDE4A strongly labeled apical VNO neurons and rostral AOB glomeruli. There was virtually no overlap in G(o alpha) and PDE4A staining, and there were no regions of the VNO neuroepithelium or AOB glomeruli not labeled by either antibody. These results identify a potential member of the pheromone transduction cascade in apical neurons, and provide further evidence that the VNO consists of functionally distinct pathways. NeuroReport 11:27-32 (C) 2000 Lippincott Williams & Wilkins.
Article
Vomeronasal sensory neurons play a crucial role in detecting pheromones, but the chemoelectrical transduction mechanism remains unclear and controversial. A major barrier to the resolution of this question has been the lack of an activation mechanism of a key transduction component, the TRPC2 channel. We have identified a Ca2+-permeable cation channel in vomeronasal neuron dendrites that is gated by the lipid messenger diacylglycerol (DAG), independently of Ca2+ or protein kinase C. We demonstrate that ablation of the TRPC2 gene causes a severe deficit in the DAG-gated channel, indicating that TRPC2 encodes a principal subunit of this channel and that the primary electrical response to pheromones depends on DAG but not Ins(1,4,5)P3, Ca2+ stores, or arachidonic acid. Thus, a previously unanticipated mechanism involving direct channel opening by DAG underlies the transduction of sensory cues in the accessory olfactory system.
Article
Lymphoid organs represent a specialized microenvironment for interaction of stromal and lymphoid cells. In primary lymphoid organs, these interactions are required to establish a self-tolerant repertoire of lymphocytes. While detailed information is available about the genes that control lymphocyte differentiation, little is known about the genes that direct the establishment and differentiation of principal components of such microenvironments. Here, we discuss genetic studies addressing the role of thymic epithelial cells (TECs) during thymopoiesis. We have identified an evolutionarily conserved key regulator of TEC differentiation, Foxn1, that is required for the immigration of prothymocytes into the thymic primordium. Because Foxn1 specifies the prospective endodermal domain that gives rise to thymic epithelial cells, it can be used to identify the evolutionary origins of this specialized cell type. In the course of these studies, we have found that early steps of thymus development in zebrafish are very similar to those in mice. Subsequently, we have used chemical mutagenesis to derive zebrafish lines with aberrant thymus development. Strengths and weaknesses of mouse and zebrafish models are largely complementary such that genetic analysis of mouse and zebrafish mutants may lead to a better understanding of thymus development.
Article
The classical class I antigens of the major histocompatibility complex (MHC) are cell-surface glycoproteins which were originally discovered because they cause rapid rejection of cells or tissues grafted between unrelated individuals. These molecules are encoded by the K, D and L loci of the mouse MHC (and analogous loci in other species) which show extreme species polymorphism and a large number of alleles. In an outbreeding population 3.6 X 10(9) unique MHC class I phenotypes can be encoded by the 100 alleles at each of the K and D loci and the 6 alleles at the L locus. This level of polymorphism ensures that the cells and tissues of each unrelated individual are uniquely identified by their class I membrane-bound antigens. Like other membrane bound proteins, these class I molecules are anchored in the lipid bilayer by a hydrophobic domain encoded by exon 5. However, there have been reports of the occurrence of classical class I molecules in true solution in the blood of humans, mice, and rats. We report here that classical polymorphic class I molecules in normal rats are constitutively excreted in the urine and that untrained rats can distinguish the smell of urine samples taken from normal donors that differ only at the class I MHC locus and therefore excrete different allelomorphs of class I molecules in their urine.
Article
Selective lesions have been made to the receptors in the main and accessory olfactory systems. In the absence of the accessory receptors, female mice are not able to show a neuroendocrine response to male pheromones but are able to detect male odours which induce the response. In the absence of main olfactory receptors, such discrimination of male odours is not possible, but the neuroendocrine response resulting in a block to pregnancy is maintained. These experiments suggest that cognitive aspects of olfaction are not essential for pregnancy block to occur and that in mice the dual olfactory systems are functional as well as anatomically distinct. Moreover, strain recognition can occur at the level of the accessory olfactory system without the female being able to display a behavioural awareness of this 'recognition'.
Article
Guidelines for submitting commentsPolicy: Comments that contribute to the discussion of the article will be posted within approximately three business days. We do not accept anonymous comments. Please include your email address; the address will not be displayed in the posted comment. Cell Press Editors will screen the comments to ensure that they are relevant and appropriate but comments will not be edited. The ultimate decision on publication of an online comment is at the Editors' discretion. Formatting: Please include a title for the comment and your affiliation. Note that symbols (e.g. Greek letters) may not transmit properly in this form due to potential software compatibility issues. Please spell out the words in place of the symbols (e.g. replace “α” with “alpha”). Comments should be no more than 8,000 characters (including spaces ) in length. References may be included when necessary but should be kept to a minimum. Be careful if copying and pasting from a Word document. Smart quotes can cause problems in the form. If you experience difficulties, please convert to a plain text file and then copy and paste into the form.
Article
Recently, two large multigene families of putative G-protein-linked receptors that are expressed in distinct subpopulations of neurones in the vomeronasal organ have been identified. These receptors probably mediate pheromone detection. The most surprising aspects of these findings are that there are so many receptors of two very different classes and that the receptors are unrelated to their counterparts in the main olfactory epithelium. This suggests that many active ligands are likely to exert effects through the vomeronasal organ. Parallel experiments addressing the nature of these ligands indicate a role for some proteins, as well as small molecules, as functional mammalian pheromones. In combination, these results begin to suggest a molecular basis for mammalian pheromone signalling.