Download Molecular and functional anatomy of the mouse olfactory epithelium

Document related concepts

Axon guidance wikipedia , lookup

Development of the nervous system wikipedia , lookup

Molecular neuroscience wikipedia , lookup

Subventricular zone wikipedia , lookup

Clinical neurochemistry wikipedia , lookup

Synaptogenesis wikipedia , lookup

Endocannabinoid system wikipedia , lookup

Feature detection (nervous system) wikipedia , lookup

Sensory cue wikipedia , lookup

Signal transduction wikipedia , lookup

Channelrhodopsin wikipedia , lookup

Neuropsychopharmacology wikipedia , lookup

Optogenetics wikipedia , lookup

Stimulus (physiology) wikipedia , lookup

Olfactory bulb wikipedia , lookup

Transcript
UMEÅ UNIVERSITY MEDICAL DISSERTATIONS
New series No. 1047
ISSN 0346-6612
Molecular and functional anatomy
of the mouse olfactory epithelium
Viktoria Vedin
Department of Molecular Biology
Umeå University
Umeå 2006
Department of Molecular Biology
Umeå University
S-901 87 Umeå, Sweden
Copyright © 2006 by Viktoria Vedin
ISBN 91-7264-138-X
Printed by Solfjädern Offset AB, Umeå, Sweden, 2006
2
"Det är en god vana att då och då ifrågasätta saker och ting som man i
åratal har tagit för givna."
Tyskt ordspråk
3
4
TABLE OF CONTENTS
ABSTRACT ........................................................................................................... 6
PAPERS IN THIS THESIS ................................................................................... 7
ABBREVIATIONS................................................................................................ 8
INTRODUCTION ................................................................................................ 9
Cell types of the olfactory epithelium ........................................................... 10
Proliferative capacity of the olfactory epithelium........................................ 12
Odorant receptors ........................................................................................... 14
Signal transduction in olfactory sensory neurons ........................................ 15
Odorant receptor expression patterns ........................................................... 18
Projections to the brain .................................................................................. 21
Initial steps in constructing a smell in the brain .......................................... 24
AIM OF THE THESIS ........................................................................................ 28
METHODS.......................................................................................................... 29
Studying odorant detection threshold........................................................... 29
RESULTS AND DISCUSSION ........................................................................... 33
Gene expression correlated to the zonal topography of ORs (paper I) ....... 33
Regionally restricted detection of odorants (papers II and III) ................... 37
Odorants induce c-Fos in OSNs in an OR-like pattern ............................ 37
Zonal ablation and effects on odorant detection....................................... 39
Regional differences in maintenance of the olfactory
epithelium (papers III and IV) ....................................................................... 43
The olfactory epithelium can be divided into two major domains ............. 48
CONCLUSIONS.................................................................................................. 50
ACKNOWLEDGEMENTS ................................................................................. 51
REFERENSES...................................................................................................... 52
5
ABSTRACT
The olfactory system is important for social behaviors, feeding and avoiding
predators. Detection of odorous molecules is made by odorant receptors on
specialized sensory neurons in the olfactory epithelial sheet. The olfactory
sensory neurons are organized into a few regions or “zones” based on the
spatially limited expression of odorant receptors. In this thesis the zonal division
and functional specificity of olfactory sensory neurons have been studied in the
mouse. We find that zones 2-4 show overlapping expression of odorant
receptors while the border between the regions that express a zone 1 and a zone
2 odorant receptor, respectively, is sharp. This result indicates that zone 1 and
zones 2-4 are inherently different from each other. In cDNA screens, aimed at
finding genes whose expression correlate to the zonal expression pattern of
odorant receptors, we have identified a number of signaling proteins implicated
in neural-tissue organogenesis in other systems. The differential expression
pattern of identified genes suggests that regional organization is maintained
during the continuous neurogenesis in the olfactory epithelium as a result of
counter gradients of positional information. We show that the gene c-fos is
induced in olfactory sensory neurons as a result of cell activation by odorant
exposure. A zonal and scattered distribution of c-Fos-positive neurons
resembled the pattern of odorant receptor expression and a change of odorant
results in a switch in which zone that is activated. Whereas earlier studies
suggest that the odorant receptors are relatively broadly tuned with regard to
ligand specificity, the restricted patterns of c-Fos induction suggests that low
concentrations of odorous molecules activate only one or a few ORs. Studies on
olfactory detection abilities of mice with zonal-restricted lesions in the olfactory
epithelium show that loss of a zone has severe effects on the detection of some
odorants but not others. These findings lend support to a hypothesis that
odorant receptors are tuned to more limited numbers of odorants. Regional
differences in gene expression and differences in response to toxic compounds
between the zones indicate that there may be differences in tissue homeostasis
within the epithelium. We have found that there are differences in proliferation
and survival of olfactory sensory neurons in regions correlating to receptor
expression zones. Identified differences with regard to gene expression, tissue
homeostasis and odorant detection show that the olfactory epithelium is divided
into regions that transduce different stimulus features.
6
PAPERS IN THIS THESIS
The thesis is based on following articles and manuscripts, which will be
referred to in the text by their Roman numerals (I-IV).
I. Norlin EM, Alenius M, Gussing F, Hägglund M, Vedin V, Bohm S,
(2001). Evidence for gradients of gene expression correlating with
zonal topography of the olfactory sensory map. Mol Cell Neurosci.
18:283-295.
II. Norlin EM, Vedin V, Bohm S, Berghard A, (2005). Odorantdependent, spatially restricted induction of c-fos in the olfactory
epithelium of the mouse. J Neurochem. 93:1594-1602.
III. Vedin V, Slotnick B, Berghard A, (2004). Zonal ablation of the
olfactory sensory neuroepithelium of the mouse: effects on odorant
detection. Eur J Neurosci. 20:1858-1864
IV. Vedin V, Berghard A, (2006). Regional differences in turnover of
olfactory sensory neurons in the adult. Manuscript.
Articles are reprinted with permission from the publishers.
7
ABBREVIATIONS
ACIII
AMP
ATP
CaM
CAMKII
cAMP
cGMP
CNG channel
Dichlobenil
GBC
HBC
Mash1
MRR
NQO1
O/E
OMP
OR
OSN
PDE
NCAM2
Npn2
RA
RALDH2
Z1-4
8
adenylyl cyclase III
adenosine monophosphate
adenosine triphosphate
calmodulin
calmodulin kinase II
cyclic adenosine monophosphate
cyclic guanine monophosphate
cyclic nucleotide-gated channel
2,6-dichlorobenzonitrile
globose basal cell
horizontal basal cell
mammalian achaete-scute homologue 1
molecular receptive range
NADPH quinone oxidoreductase 1
olfactory/early B-cell factor
olfactory marker protein
odorant receptor
olfactory sensory neuron
phosphodiesterase
neural cell adhesion molecule 2
neuropilin 2
retinoic acid
retinaldehyde dehydrogenase 2
odorant receptor expression zones 1-4
INTRODUCTION
“You cannot suppose that atoms of the same shape are entering
your nostrils when stinking corpses are roasting as when the
stage is freshly sprinkled with saffron of Cicilia and a nearby
alter exhales the perfumes of the Orient”
Lucretius
The human nose can detect an almost unbelievable large number of
smells or odors (400 000 distinct odors, (Mori et al., 2006) and
references therein) even though our sense of smell is not our most
important sense under normal circumstances. The sense of smell is very
important to most animals in order to survive, find food, avoid predators
and to find a mating partner. I have in this thesis studied different
aspects of the functional anatomy of the olfactory epithelium, focusing
on one model organism, the mouse.
Figure 1. The olfactory epithelium covers the turbinates protruding into the
nasal cavity. OSNs in the olfactory epithelium send their axons to target cells
in the olfactory bulb which is located in the anterior part of the brain.
9
A perceived odor is made up of one or several distinct odorous
molecules, or odorants. Odorants are typically small organic molecules
that are detected by the olfactory system by specialized receptor cells in
the olfactory epithelium located in the nasal cavity. The main primary
olfactory system consists of the olfactory epithelium located in the roof
of the posterior nasal cavity and the olfactory bulb located at the most
anterior part of the telencephalon. The nasal cavity is in rodents highly
convoluted with turbinates protruding into the nasal lumen (fig 1). The
olfactory epithelium covers the turbinates, thus resulting in a larger
surface area of epithelium specialized to detect odorants. The intricate
structure of the turbinates also result in a decrease in air velocity as
indicated by computer simulations of the airflow in the rat nasal cavity
(Kimbell et al., 1997). It is possible that the decrease in air velocity
facilitates sorbtivness of certain odorants. The odorant-containing air
that enters the nasal cavity during inhalation is flushed over the
epithelium and the odorant molecules stimulate specialized sensory
cells, olfactory sensory neurons (OSNs). The OSNs project their axons
through the cribiform plate up to the olfactory bulb where the axons
make contact with second order neurons. The signal is here refined and
transferred to higher areas of the brain where it is interpreted as a scent
and the animal becomes aware of the sensation and can respond
accordingly.
Cell types of the olfactory epithelium
The olfactory epithelium is a pseudostratified epithelium consisting of
several cell types in which the nuclei form distinct layers; sustentacular
cells apically, neuronal cells in the middle compartment and basal cells
located closest to the basal lamina (fig 2) (Graziadei & Graziadei, 1979;
Nomura et al., 2004). The basal cell layer population contains horizontal
basal cells (HBC) and globose basal cells (GBC). The HBCs form a single
layer of flat cells situated directly on the basal lamina with round GBCs
located above the HBCs. The OSN cell layer consists of both immature
10
and mature OSNs where the mature OSNs constitute the majority of
cells in the adult olfactory epithelium. The mature OSNs are located
more closely to the nasal lumen compared to the immature OSNs
(Graziadei & Graziadei, 1979). The OSN is a bipolar neuron with a
single dendrite that terminates in a knob from which 10-20 fine cilia
extend into the mucus that covers the epithelium (Menco, 1980;
Mendoza, 1993). The cilia are the structures that detect odorant and it is
also here the binding of an odorant to the odorant receptor (OR) is
converted into electric activity by the intracellular signal transduction
cascade. The mature OSN sends a single unmyelinated axon through the
cribiform plate to the target in the brain, the olfactory bulb. The
sustentacular cells are non-neuronal supporting cells forming a single
row of cells occupying the full height of the epithelium although the
cells are wider closer to the nasal lumen. The sustentacular cells have
Figure 2. The olfactory epithelium consists of several cell types. Sustentacular
cells are located closest to the nasal lumen while the middle layers contain
mature and immature OSNs. The layers closest to the basal membrane
contain basal cells (GBCs and HBCs) of which some cells are the progenitors
to OSNs. The epithelium is covered by mucus produced by Bowman’s glands
which ducts extend through the epithelium.
11
branched foot processes spreading on the basal lamina. These cells
provide support to the OSNs and metabolize xenobiotic compounds and
odorants (Ding & Coon, 1988; Chen et al., 1992; Yu et al., 2005).
Sustentacular cells are also known to phagocytous dead OSNs (Suzuki et
al., 1995). The sustentacular cells are connected with tight junctions
resulting in separation of the receptor cell body from the nasal lumen
(Miragall et al., 1994; Rash et al., 2005). Bowman’s glands are located in
the mesenchyme underneath the olfactory epithelium and their ducts
protrude the epithelium. Bowman’s glands produce mucus covering the
olfactory epithelium and are like sustentacular cells rich in detoxifying
enzymes (Yu et al., 2005).
Proliferative capacity of the olfactory epithelium
The mammalian olfactory system maintain neurogenesis throughout life
(Graziadei & Graziadei, 1979) which makes it an effective model for
studying principles of tissue renewal and neuronal cell differentiation in
the adult as well as during development. The average life-span of an
OSN has been reported to be about 90 days (Mackay-Sim & Kittel,
1991), even though OSNs have been reported to survive for more than
12 months (Hinds et al., 1984). The olfactory epithelium is, due to its
unprotected position, easily damaged as a result of exposure to toxic
compounds, viral infections or mechanical trauma. The olfactory
epithelium is often functionally repopulated with OSNs after damage
which is achieved by progenitor cell proliferation and differentiation.
The nature of the stem cell of the olfactory epithelium is still unclear;
possible candidates have been suggested to reside in the HBC
compartment, in the GBC compartment as well as amongst the
Bowman’s duct cells (Matulionis, 1975; Graziadei & Graziadei, 1979;
Calof & Chikaraishi, 1989; Caggiano et al., 1994; Huard et al., 1998).
HBCs proliferate at a low rate where about 2% of the actively dividing
cells in the basal compartment of the normal epithelium are HBCs
(Holbrook et al., 1995; Huard & Schwob, 1995). HBCs have been
12
indicated by in vitro experiments to give rise to both OSNs, glia cells
and HBCs (Carter et al., 2004). Nevertheless, the general view is that the
stem cell is a member of the GBC population (for review on
regeneration of olfactory epithelium see (Schwob, 2002)). GBCs have
been suggested to give rise to OSNs in vivo (Mackay-Sim & Kittel, 1991;
Schwartz Levey et al., 1991; Caggiano et al., 1994; Gordon et al., 1995)
and proliferation in the GBC layer is increased after damage to the
primary olfactory system, for example by bulbectomy (Schwartz Levey
et al., 1991; Suzuki & Takeda, 1991), axotomy (Suzuki & Takeda, 1991)
or damage by compounds that are toxic to the olfactory epithelium.
After injury to the olfactory epithelium by the gas methyl bromide are
areas with GBCs repopulated with all cell types of the olfactory
epithelium whereas areas where GBCs are absent will become an
epithelium with respiratory epithelial character (Jang et al, 2003). Areas
that are repopulated sometimes lack HBCs during an early phase of
recovery (Schwob et al., 1995). Moreover, some GBCs exhibit
multipotency since highly purified GBCs transplanted into methyl
bromide lesioned epithelium will give rise to clonal clusters containing
both neural (OSNs and GBCs) or non-neural cells (Bowman’s glands,
sustentacular cells and cells of the respiratory epithelium) or clusters
containing both non-neural cells and neural cells, indicating the
existence of different GBC populations (Chen et al., 2004). A subset of
GBCs remains quiescent during long time-periods as observed by
retention of intensity of labeled genomic DNA and this subset may
resume active cell cycling after damage by methyl bromide (MackaySim & Kittel, 1991).
During regeneration as well as normal turnover of the olfactory
epithelium, GBCs function as the immediate progenitors to the OSNs.
The GBCs are the cells that have the highest proliferative rate of all cells
in the olfactory epithelium. About 97,2% of the actively dividing cells
in the basal compartment of the normal epithelium are GBCs (Huard &
Schwob, 1995). When a progenitor in the GBC population divides it
gives rise to immature OSNs expressing immature markers (e.g. SCG10
(Camoletto et al., 2001)), the immature OSNs will, in time, mature to
13
functional OSNs expressing mature markers (e.g. olfactory marker
protein (OMP) (Farbman & Margolis, 1980)) that establish functional
contact with the olfactory bulb.
Odorant receptors
Odorant molecules are bound to and detected by odorant receptors
(ORs) located on the cilia of the OSNs. The ORs were first identified in
the rat in 1991 (Buck & Axel, 1991) and were predicted to be G proteincoupled receptors that span the cell membrane seven times. Mice have
roughly 1,300 OR genes of which about 20% are predicted to be
pseudogenes (Zhang & Firestein, 2002; Godfrey et al., 2004). The
importance of this sense for many animals may be mirrored by the
abundance of ORs. The OR family is the largest gene superfamily in the
mammalian genomes (Zhang & Firestein, 2002). Humans have, in
comparison to the mouse, few OR coding genes and many genes are
predicted to be pseudogenes (roughly 640 genes of which 46% are
pseudogenes) (Malnic et al., 2004). The mouse OR genes are organized
in 27 clusters containing multiple OR genes and clusters are located on
all chromosomes except 12 and Y (Sullivan et al., 1996; Zhang &
Firestein, 2002).
The OR sequences separate in two broad classes, class I and class II
receptors (Zhang & Firestein, 2002). Class I ORs was first identified in
fish (Ngai et al., 1993b) and was later identified in humans (Glusman et
al., 2001) and mouse (Zhang & Firestein, 2002). All class I ORs are
located in a cluster on chromosome 7 (Tsuboi et al., 2006). In the
mouse, there are 147 class I OR genes of which 120 are potentially
functional (Zhang & Firestein, 2002).
The OR protein is encoded by a gene with a single coding intronless
exon of about 1 kilobase pairs (e.g. (Sosinsky et al., 2000)). The OR
protein has seven transmembrane regions where transmembrane
domains 4 and 5, and to a lesser extent transmembrane domain 3, are
14
the most variable regions (Pilpel & Lancet, 1999). Computer models of
two ORs indicate that the recognition of an odorant by its OR is
mediated by amino acids in the transmembrane domains 3 to 6 (Singer
& Shepherd, 1994; Katada et al., 2005), where point mutations can
change ligand specificity (Katada et al., 2005).
Signal transduction in olfactory sensory neurons
The detection of an odorant starts with odorant binding to the G
protein-coupled OR. The activation of ORs by odorants starts an
intracellular signal transduction cascade in the tiny volume of the
cilium, which in the end results in an action potential in the OSN (fig
3). Odorant binding to the OR results in the activation of a
heterotrimeric G protein. The Gα-subunit Gαolf is present at high levels
in the olfactory cilia (Jones & Reed, 1989) and stimulates the
transmembrane adenylyl cyclase III (ACIII) (Pace et al., 1985; Bakalyar
& Reed, 1990) to convert adenosine triphosphate (ATP) to cyclic
adenosine monophosphate (cAMP). Elevation of intraciliary cAMP
opens cyclic nucleotide-gated (CNG) channels (Nakamura & Gold,
1987) that are non-selective cation channels (Balasubramanian et al.,
1995; Dzeja et al., 1999). The opening of the CNG channels results in an
increase in intraciliary Ca2+ and Na+ ions (Leinders-Zufall et al., 1997)
resulting in a slight depolarization of the ciliary membrane. The
increase in Ca2+ in turn opens Ca2+-activated Cl--channels (Kleene, 1993;
Reisert et al., 2003) driving Cl- to exit the cell due to its high
intracellular concentration and its release hence further depolarize the
cell (Reuter et al., 1998). The resulting receptor potential is then
converted to trains of action potentials propagating to the axon
terminals.
There is a general consensus that the vast majority of OSNs react to
odorant exposure by activating ORs that are linked to stimulation of
adenylyl cyclase by Gαolf, since single knock-outs of either of the
transduction components Gαolf, ACIII or the CNG channel subunit
15
Figure 3. Schematic picture of the signal transductions cascade in OSNs in
response to odorant binding to the OR. Adaptation is mediated by feed-back
mechanisms on several levels of the transduction cascade provided by
increased levels of intracellular calcium.
CNGA2 result in animals that are anosmic for a wide range of odorants
(Brunet et al., 1996; Belluscio et al., 1998; Wong et al., 2000). However,
there might be exceptions since mice genetically null for CNGA2 still
detect some odorants (Lin et al., 2004). There have been reports of OSNs
not expressing either Gαolf, ACIII or CNGA2 but that instead express
guanylyl cyclase-D, cyclic guanine monophosphate (cGMP)-stimulated
phosphodiesterase and a cGMP-selective cyclic nucleotide gated
channel (Juilfs et al., 1997; Meyer et al., 2000). Inositol-1,4,5trisphosphate and diacylglycerol produced by phospholipase C have also
been implicated in odorant signal transduction (reviewed in (Elsaesser
& Paysan, 2005)). The roles of these alternative signaling pathways are
currently unresolved.
Six CNG channel genes have been identified in the mammalian genome,
CNGA1-4, CNGB1 and CNGB3 (for a review of CNG channels see
(Kaupp & Seifert, 2002)). CNG channels in OSNs are most likely
16
heterotetramers (Liu et al., 1996) and have a fixed stoichiometry of two
CNGA2 subunits, one CNGA4 and one CNGB1b (Zheng & Zagotta,
2004). CNGA2 forms functional homomeric channels when expressed
in a heterologous system while CNGA4 and CNGB1b only form
functional channels as heteromers with CNGA2 suggesting that CNGA2
is the functional channel and CNGA4 and CNGB1 have roles in
modulating the channel properties (Kaupp et al., 1989; Bradley et al.,
1994; Liman & Buck, 1994; Zheng & Zagotta, 2004).
Feed-back mechanisms likely ensure that the transduction cascade is
terminated and may result in OSNs adjusting the sensitivity, adapting,
to different stimulus intensities to prevent saturation of the
transduction machinery. The process of signal termination and
adaptation occurs already at the level of signal transduction by feedback
of Ca2+ that has modulatory roles at various steps in the transduction
cascade (adaptation reviewed in (Zufall & Leinders-Zufall, 2000)). Ca2+
is important both in signal termination and adaptation where the
adaptive actions of Ca2+ can be divided into two categories; those on the
CNG channel itself and those on other components in the transduction
cascade. The principal mechanism underlying signal termination and
adaptation is modulation of the CNG channel by Ca2+ feedback
(Kurahashi & Menini, 1997). This inhibition is a result of a complex
between Ca2+ and calmodulin (CaM) where, at high Ca2+ concentrations,
Ca2+-CaM reduces the sensitivity of the CNG-channel to cAMP 10-fold
(Chen & Yau, 1994; Trudeau & Zagotta, 2003). Negative modulation by
Ca2+ through CaM on CNGs requires the subunits CNGB1b and CNGA4
(Bradley et al., 2001; Munger et al., 2001; Bradley et al., 2004). It has
been suggested that CNGA4 null mice exhibit slower Ca2+-CaMmediated channel desensitization (Munger et al., 2001) and have
problems with detecting odorants in a background of another odorant,
suggesting that CNGA4 is important for adaptation (Kelliher et al.,
2003). The inhibitory effect of Ca2+ on the CNG channel is fast and this
can be achieved by constitutive binding of CaM to the CNG channel
(Bradley et al., 2004). In addition of regulating the sensitivity of the
CNG channel to cAMP, Ca2+-CaM also activates a phosphodiesterase
17
(PDE) resulting in breakdown of cAMP to adenosine monophosphate
(AMP) and thus reduces the level of cAMP (Borisy et al., 1992; Yan et
al., 1995). Ca2+-CaM also influences the level of cAMP by inhibiting
adenylyl cyclase by the action of the activated kinase calmodulin kinase
II (CamKII) that phosphorylates ACIII (Wei et al., 1996; Wei et al.,
1998; Leinders-Zufall et al., 1999). The effect of Ca2+-CaM on the cAMP
level in the cilia is somewhat slower than the effect on the CNG
channel.
The response of an odorant in the OSN can be influenced by other
adaptive processes like phosphorylation of the receptor via G-proteincoupled receptor kinase 3 which is mediated by the βγ subunits
(Boekhoff et al., 1994; Peppel et al., 1997). The amount of cGMP is, as
well as the level of cAMP, increased after odorant stimulation. The rise
in cGMP concentration after odorant stimulation is both slower and
more sustained than the increase in cAMP (Zufall & Leinders-Zufall,
1997; Moon et al., 2005).
Odorant receptor expression patterns
Shortly after the cloning of the first ORs it was evident that each OR is
expressed in a small fraction of OSNs, starting at embryonic day 11.5
(Strotmann et al., 1992; Ressler et al., 1993; Vassar et al., 1993; Sullivan
et al., 1995). Each OSN expresses only one receptor (Chess et al., 1994;
Malnic et al., 1999) in a monoallelic and mutually exclusive manner
(Chess et al., 1994; Serizawa et al., 2000). Exception to the singular
expression of ORs exist as an OSN have been reported to express two
different ORs (Rawson et al., 2000).
Several mechanisms behind the mutual exclusive expression of the ORs
have been suggested, including irreversible DNA changes like DNA
recombination or gene conversion as well as reversible mechanisms. By
cloning mice from post mitotic OSNs two studies have shown that
irreversible DNA changes in the nuclei of OSNs is not the mechanism
18
behind OR gene selection (Eggan et al., 2004; Li et al., 2004). The
resulting cloned mice do not display monoclonal expression of ORs but
rather have an olfactory epithelium expressing several types of ORs, in a
similar manner as control mice. A cis-acting DNA region, called the Hregion, was suggested to activate expression of OR genes in a specific
OR gene cluster and to restrict the expression to only one out of the
three OR genes present in a transgenic mouse line (Serizawa et al.,
2003). The H-region was later implied to mediate monoallelic
expression of all ORs (Lomvardas et al., 2006). The H-region exists in
two copies in each cell and one allele is methylated in OSNs. The
methylation is implied in inactivation of one allele in order to achieve
one functional H-region and expression of only one OR gene copy. A
functional OR protein might stabilize expression of the chosen OR and
Figure 4. The expression areas of four ORs in one nostril are shown by
radioactive in situ hybridizations, positive signal in white. ORs are expressed
in restricted areas of the olfactory epithelium in scattered cells where the
most dorsomedial expression area is called zone 1 (Z1) and the most
ventrolateral is called zone 4 (Z4).
19
exclude expression of additional ORs as the expression of a pseudogene
fail to inhibit expression of a functional OR gene (Serizawa et al., 2003)
and the presence of an untranslatable OR coding sequence in the mRNA
is not sufficient to exclude expression of a second OR (Lewcock & Reed,
2004).
OSNs expressing a given OR are not spread over the entire epithelial
sheet but are rather restricted to broad areas of the epithelium. A zonal
topography is evident at embryonic day 13.5 when enough cells are
expressing ORs to be able to discern a pattern (Sullivan et al., 1995). The
olfactory epithelium is classically divided into four zones based on the
expression pattern of a few ORs (Ressler et al., 1993; Vassar et al., 1993).
The OSNs expressing a particular OR are intermingled with other OSNs
expressing other ORs within such an expression zone (Ressler et al.,
1993). The zones are distributed in a dorsomedial to ventrolateral
orientation where zone 1 (Z1) is the most dorsomedial zone and zone 4
(Z4) is the most ventrolateral zone (fig 4). Other patterns of expression
have also been identified (Kubick et al., 1997; Strotmann et al., 1999).
Expression differences can be observed between class I and class II ORs
where class I ORs are expressed predominantly in Z1 (Tsuboi et al.,
2006), class II ORs, on the other hand, are expressed in any of the zones,
but are in Z1 intermingled with OSNs expressing class I ORs (Zhang et
al., 2004; Tsuboi et al., 2006).
Small regions upstream of the transcription start site of OR genes have
been implicated in the regulation of the zonal specific expression (Qasba
& Reed, 1998; Serizawa et al., 2000; Rothman et al., 2005). These
regions contain homeobox and olfactory/early B-cell factor (O/E)
binding sites where a minimal promoter containing one homeobox and
one O/E site faithfully mimic the spatial distribution of one OR studied
(Rothman et al., 2005). The number and order of motifs (e.g. homeobox
and O/E sites) may regulate the pattern of OR expression by modulating
the probability of expression of different ORs differentially across the
olfactory epithelium (Rothman et al., 2005; Hoppe et al., 2006).
20
Projections to the brain
In the olfactory epithelium axons of OSNs form loose bundles between
the "feet" of HBCs, traveling in a rostro-caudal direction. As they cross
the basal lamina and exit the epithelium, processes of olfactory
ensheating cells tightly enwrap the axon bundles (Holbrook et al.,
1995). Ensheating cells can be observed forming tunnels containing
hundreds of OSN axons traveling to make synaptic contacts with target
neurons in the olfactory bulb (Li et al., 2005). The axons pass the
cribiform plate and form, together with glial cells, the nerve layer
surrounding the olfactory bulb. The axons defasciculate upon reaching
the olfactory bulb and target different regions of the bulb (Whitesides &
LaMantia, 1996). The OSN axons terminate in anatomically distinct
structures called glomeruli (Graziadei & Graziadei, 1979) where
Figure 5. OSNs expressing the same OR send their axons to the same
glomeruli and make synaptic contacts with target cells e.g. mitral cells.
Mitral cells send the olfactory signal further up to the olfactory cortex.
Granule cells make dendrodendritic contacts between different mitral cells
and regulate mitral cell output to the olfactory cortex. The olfactory bulb
has distinct layers formed by different cell types.
21
synaptic connections are made with e.g. principal projection neurons
(fig 5). The mouse olfactory bulb contains about 1800 glomeruli (Royet
et al., 1989). Neurons expressing the same OR converge upon the same
glomeruli, usually one medial and one lateral (first noticed by (Ressler
et al., 1994; Vassar et al., 1994)). The zonal organization is maintained
in the olfactory bulb (Alenius & Bohm, 1997; Yoshihara et al., 1997;
Gussing & Bohm, 2004; Miyamichi et al., 2005) where the position of
OR expression in the olfactory epithelium determines the dorsal/ventral
positioning of target glomeruli in the olfactory bulb (Miyamichi et al.,
2005). The glomerulus targeted by a particular OR is located
approximately in the same position between individuals (fig 6)
(Strotmann et al., 2000), forming a stereotype map of projections from
the epithelium to the olfactory bulb. OSNs expressing highly
homologous receptors from the mOR37 subfamily project to glomeruli
that are situated close to each other in the olfactory bulb (Strotmann et
al., 2000). In the fruit fly Drosophila, OSNs expressing similar ORs are
not clustered in the periphery but OSN expressing the same OR project
to the same glomeruli and OSNs expressing similar ORs project to
closely positioned glomeruli (Couto et al., 2005).
Each glomerulus receives input from thousands of OSNs which in the
glomerulus make synaptic connections with second order neurons, i.e.
mitral and tufted cells that in turn project to higher olfactory centras.
Each glomerulus contains the dendrites of several mitral and tufted cells
and each mitral cell innervates one single glomerulus ((Schoppa &
Urban, 2003) and references therein). There are two classes of
interneurons in the olfactory bulb; the periglomerular cells and the
granule cells. Cells of both types are added throughout life of the animal
(Kaplan et al., 1985) and modulate the activity of OSNs as well as the
mitral and tufted neurons. Periglomerular cell synapses are mostly
restricted to one glomerulus where they mediate intraglomerular
interactions. Granule cells make synaptic connections between mitral
cells associated with different glomeruli. The synaptic organization of
the glomerulus is reviewed in (Schoppa & Urban, 2003; Kosaka &
Kosaka, 2005).
22
The axons from the olfactory bulb constitute the lateral olfactory tract
projecting to the olfactory cortex. Olfactory cortex is defined as the
cortical areas receiving direct input from the olfactory bulb and
includes, among several areas, the anterior olfactory nucleus, the
entorhinal cortex, the piriform cortex and the olfactory tubercle. A
transgenic mouse line expressing Barley lectin under the control of
specific promoters made it possible to chart neural circuits due to the
ability of the lectin to transfer transneuronally (Horowitz et al., 1999).
It was indicated, by expressing Barley lectin under the control of two
different OR promoters, that the projections from the olfactory bulb to
cortical areas display some similarity between individuals. The inputs
originating from ORs, diverge in the olfactory cortex to clusters of
thousands of neurons overlapping with clustered inputs from other ORs
(fig 6) (Zou et al., 2001). These findings are supported by a study using
c-Fos as a neuronal activity marker where single odorants activate a
small subset of cells in the anterior piriform cortex. The different
odorants had different but overlapping representations of activated cells
(Zou et al., 2005).
Figure 6. OSNs expressing the same OR project to the same glomerulus and
the zonal topography of OR expression in the olfactory epithelium is
maintained in the projections to the olfactory bulb. In contrast, the
projections to the olfactory cortex are overlapping between axons coming
from different areas of the olfactory bulb. The projections from the olfactory
epithelium to the bulb and cortex are similar between individuals.
23
Initial steps in constructing a smell in the brain
In the environment, thousands of chemical molecules exists that need to
be detected and identified in order for the animal to be able to interpret
the world. Large numbers of structurally diverse odorant molecules
interact with the ORs. It is likely that more chemicals can be smelled
than there are ORs for them, making a combinatorial representation
necessary.
Computer models indicate that air entering the rat nasal cavity first
flows through dorsomedial areas and then courses more peripherally
(Kimbell et al., 1997; Zhao et al., 2006). The uptake of odorants in the
mucus may be changed by changing airflow velocity over the
epithelium so that efficient adsorption of odorants is achieved in areas
expressing ORs most sensitive to the odorant ((Schoenfeld & Cleland,
2005) and references therein). The odorant is, after being adsorbed in
the mucus covering the olfactory epithelium, identified by ORs located
on the surface of the OSNs. An OSN expresses only one out of the
thousand ORs in the mouse genome and it is hypothesized that an OR
recognizes a particular chemical structure or molecular feature, for
example a specific polar group or molecules with a certain carbon chain
length (Malnic et al., 1999). Hence an OSN can be activated by one
particular chemical structure; the structure identified by the OR. An
odorant can, because it often contains several chemical structures, be
detected by several ORs (Araneda et al., 2000) and thus several different
types of OSNs. The range of odorants that activates an OR can be called
the ORs molecular receptive range (MRR). In vitro studies have
indicated that ORs that are >60% identical has a high probability of
being stimulated by odorants with related structures (Malnic et al.,
1999; Kajiya et al., 2001). The structure of an odorant might thus be
coded by activation of a specific subset of OSNs (Sato et al., 1994;
Malnic et al., 1999; Ma & Shepherd, 2000; Kajiya et al., 2001). The
perceived quality of an odorant can change when the odorant
concentration is changed. The explanation for this may be that the ORs
24
have different affinities to odorants and at higher odorant concentration
a recruitment of additional receptors occurs (Amoore, 1982; Ma &
Shepherd, 2000; Kajiya et al., 2001). The olfactory information that is
extracted from the environment by the OSN is influenced by the OR
repertoire of the animal. Comparison of human and mouse OR
subfamilies shows that the two species have many, but not all,
subfamilies in common. However, mouse subfamilies are usually larger
than the human counterparts. This finding suggests that humans and
mice recognize many of the same odorant structural motifs, but mice
may be superior in odor discrimination and may be superior in
identifying different odorants (Godfrey et al., 2004). The inability to
detect specific odorants, so called specific anosmia, has been described
both in humans and rodents. One example is the mouse strain C57Bl/6
that is anosmic to the odorant isovaleric acid (Wysocki et al., 1977). The
anosmia was later found to be associated with a locus containing an OR
cluster, indicating that loss of ORs may be the cause of the specific
anosmia (Griff & Reed, 1995; Zhang & Firestein, 2002).
The identification of ligands for ORs has been hampered by the
difficulty to express ORs at the surface in heterolougous cell lines and
up to date a few proteins have been identified that facilitate cell surface
localization of some of the ORs (Saito et al., 2004; Neuhaus et al., 2006).
The first in vivo identification of a ligand to an OR was made in the rat
when the I7 receptor was expressed in OSNs normally not expressing
that OR by employing adenovirus mediated delivery. The OSNs
displayed physiological responses to the odorant octanal (Zhao et al.,
1998). Since then several identifications of ligands have been made in
cell culture. Some of the mouse ORs, identified as having ligands such as
aliphatic acids, alcohols and aldehydes, are known to belong to class I
ORs (Malnic et al., 1999; Saito et al., 2004).
All OSNs expressing a certain OR converge to the same glomeruli. The
response profile of a glomerulus is determined by the OR expressed by
the OSNs projecting to it (Bozza et al., 2002) and the MRR of the
glomerulus is thus the same as the MRR of the OSNs that innervate the
25
glomerulus. The odorant stimuli received by the olfactory epithelium
thus appear to be converted to a topographic map of activated glomeruli
in the olfactory bulb (Ngai et al., 1993a; Ressler et al., 1994; Vassar et
al., 1994; Malnic et al., 1999; Rubin & Katz, 1999; Uchida et al., 2000;
Fried et al., 2002). Imaging studies demonstrate that different odorants
elicit defined patterns of glomerular activation, revealing a functional
representation of the anatomical map (Rubin & Katz, 1999; Uchida et
al., 2000; Wachowiak & Cohen, 2001). OSNs expressing a limited
number of highly homologous receptors project to glomeruli that are
closely juxtaposed in the olfactory bulb (Strotmann et al., 2000) which
is consistent with clustered glomerular responses to odorants with
similar structure (reviewed in (Mori et al., 2006)) and the fact that
highly homologous receptors detect molecules with related structures
(Malnic et al., 1999; Kajiya et al., 2001). Imaging studies of a limited
area of the olfactory bulb have suggested that the response patterns after
activation with natural complex odors are highly specific and that the
number of glomeruli activated by a complex odor are well predicted by
the sum of activated glomeruli after stimulation with the individual
components of the complex odor (Lin et al., 2006).
The OSN axons make synaptic contacts with mitral and tufted cells in
the glomeruli, which then project to the olfactory cortex. The signals in
the glomeruli are processed by local interneurons like periglomerular
cells and granule cells. Mitral cells connected to glomeruli that respond
to a wide range of related odorants receive inhibitory inputs from
neighboring glomerular units via lateral inhibition. Lateral inhibition is
mediated by granule cells via reciprocal dendrodendritic connections
between mitral cells from different glomeruli, thereby regulating mitral
cell axonal output (Rall et al., 1966; Isaacson & Strowbridge, 1998;
Margrie et al., 2001). Moreover, strongly activated glomeruli suppress
transmitter release in surrounding glomeruli via excitatory actions on
periglomerular cells (Aungst et al., 2003; Vucinic et al., 2006). The
glomerulus that is exhibiting the strongest activation thereby inhibits
the other activated glomeruli. The MRR of a mitral cell is thus not the
same as the MRR of the OSN targeting the glomerulus, but is modified
26
by interactions with other glomeruli. The olfactory bulb thus functions
to sharpen the signal coming from the olfactory epithelium before it is
sent to the olfactory cortex (Mori et al., 1999). After processing in the
olfactory bulb, the signal is transferred to olfactory cortex where mitral
cells axons show overlapping patterns of innervations. Two mitral cell
responding to different odorants may make contact with the same
olfactory cortex neuron and individual neurons in the olfactory cortex
can respond to multiple odorants (Haberly, 1969; Tanabe et al., 1975).
Increasing the odorant concentration results in higher number of
activated cells. Neurons activated by a certain odorant are scattered in
cortex (Cattarelli et al., 1988) and neurons projecting from two different
glomeruli have overlapping termination sites (Zou et al., 2001). There
have also been reports on cells that act as coincidence detectors, cells
that are activated by binary mixtures of odorants but not activated by
the individual odorant components (Zou et al., 2005; Zou & Buck,
2006). These studies suggest integration of signals from different ORs at
the level of the olfactory cortex which may be part of the processing
that eventually leads to an odor percept (Zou et al., 2005). Olfactory
cortex in turn projects to subcortical, limbic and neocortical regions
involved in motivated behavior, autonomic reflexes and hormonal
modulation.
27
AIM OF THE THESIS
The reason for the spatially restricted expression of ORs and the
division of the epithelium into several zones that express unique ORs
are up to date not known. I have in this thesis studied the olfactory
epithelium of the mouse focusing on molecular and anatomical
differences between OR expression zones of the olfactory epithelial
sheet.
Specific aims were to:
- Identify genes whose expression pattern corresponds to the spatial
expression of ORs in order to find a mechanism used in the
specification of the OSNs.
- Study zonal activation of OSNs after odorant exposure in order to
examine the response properties of the ORs.
- Explore differences in tissue homeostasis over the olfactory
epithelium in intact as well as in damaged epithelium.
28
METHODS
Studying odorant detection threshold
A number of behavior assays have been used to measure olfactory
detection in mice and rats. One of the most simple is the hidden cookie
test were a piece of food is hidden in the bedding of the cage or another
arena. The animal is then introduced into the arena and allowed to
explore the area and to find the hidden snack. A course estimate of the
animals olfactory abilities can then be based on the time it takes for the
animal to achieve the task. This test contains a lot of parameters that is
not related to olfaction like activity and attention of the animal. A test
with a little better resolution is the two choice paradigm where the
animal is conditioned to an odorant by simultaneous presentation of a
reward during the training phase. During the test period the animal is
presented with the test odorant and an unrelated odorant, both without
the reward. A positive answer would here be if the animal chooses to go
to the conditioned odorant. The concentration of the odorants can be
altered to get a course estimate of the detection threshold. Although
these methods are simple and do not require any special equipment they
are either too unspecific or require much work in order to obtain higher
resolution of detection threshold values.
Automated operant conditioning enables the analysis of behavior in a
standardized way with relatively little time required for handling the
animals compared to the quality of the result obtained. Slotnick and
colleagues have worked on automated operant conditioning for odorant
detection and then mainly in the rat (e.g. (Slotnick & Bodyak, 2002;
Slotnick et al., 2004)). Almost all of the mammalian research of the
molecular biology of olfaction has been done in the mouse which is an
animal that has not been widely used in operant conditioning studies of
sensory functions. The increasing use of gene targeted mice in
29
Figure 7. The olfactometer is used to analyze detection or discrimination of
odorants. This setting was used in paper III to analyze for effects on odorant
detection thresholds after selective ablation of OSNs. In the lower picture a
mouse is inserting the nose in the odor sampling port, trying to detect an
odorant in order to get a reward in form of a drop of water.
molecular biology and the need to measure olfactory capabilities in such
strains motivated Slotnick and co-worker to test mice in a modified
olfactometer originally designed for rats (Bodyak & Slotnick, 1999). The
test showed that mice do perform well on a go-no go task that can be
used to determine the detection thresholds of some odorants in wild
type mice.
We have analyzed the detection threshold for a few odorants in adult
mice subjected to OSN ablation as described in paper III. The
olfactometer used in our study and by (Bodyak & Slotnick, 1999) is an
30
eight-odor olfactometer (see fig 7). Air is passed through a carbon filter
and over the surface of diluted odorants placed in odor saturator bottles.
The air is then presented to the mouse at a sampling port. The odorant
that reaches the mouse is 2.5% of the concentration in the headspace
above the odorant in liquid.
The initial training is aimed at teaching the mouse to lick at a
reinforcement tube in the sampling port when an odorant is presented.
The lick is what is measured and automatically recorded. The first stage
trains the mouse to lick at the tube for a drop of water. The mouse is
kept at a restricted water schedule in order to ensure that the mouse is
motivated during the experiment. The second stage aims at training the
mouse to withdraw the snout before a new trial can be initiated. In the
third stage odorant is introduced and the animal receives the reward
after sampling the odorant. When a mouse is properly trained, it
responds to odorant presentation by licking at the reinforcement tube
only when odorant is presented. Odorant and clean air is presented in
random order in blocks of 20 trials during testing. The mouse is run in a
total of five consecutive blocks (a session). Licking when odorant is
presented and not licking when there is no odorant are considered as
correct responses. Not licking when odorant is presented or licking
when there is no odorant are considered as incorrect responses. If the
mouse achieves 90% correct responses in a single block it is considered
to have reached criterion performance and is tested in a new session
with a lower concentration of the odorant investigated. The mice were
in paper III tested for three sessions per day where a maximum of two
sessions were at the same concentration to ensure that the mouse does
not fail to show that it detects a certain odorant concentration due to
inter daily fluctuations in factors beyond our control. When the mouse
has failed to detect the odorant of a certain concentration in three
sessions the test is terminated and the mouse is assigned the preceding
concentration as the detection threshold.
One has to make a number of precautions in order to be sure that the
mouse actually responds as a consequence of sensing the odorant and
31
not due to e.g. sound cues from the valves or from odor cues from
tubing or the odor saturator bottles. For example, we use “distraction
valves” that open randomly when the odor valves are operated. The
positions of the saturator bottles for one particular concentration are
changed, both within and between sessions, so that the mouse does not
easily learn to distinguish between the sounds of opening or closing of
odor valves. Moreover, we frequently change the tubing and odor
saturator bottles. The odor saturator bottles are changed daily and
washed in ethanol and stored at 70°C. The tubing can easily be pinched
above the odor saturator tubes in order to close the odor stream which
is yet another way to test if the mouse responds to the odorant and not
to other cues. The most reliable way to test for cues is to run a trained
mouse. Cues exist if the mouse is able to get high number of correct
responses when no odorant or a concentration below detection
threshold is used.
Since the publication of our study, Clevenger & Restrepo have reported
on an alternative protocol for determining a odorant detection
threshold using the olfactometer (Clevenger & Restrepo, 2006). Instead
of using descending concentrations of odorants they used random
presentation of odorant concentrations to “home in” on the detection
threshold. Alternating presentation of different odorant concentrations
also decrease the number of trials needed.
32
RESULTS AND DISCUSSION
Gene expression correlated to the zonal topography of ORs
(paper I)
ORs are expressed in limited regions of the olfactory epithelium. The
epithelium has classically been divided into four zones based on OR
expression (Ressler et al., 1993; Vassar et al., 1993). We could in paper I,
however, show that Z2-4 ORs have overlapping areas of expression
while Z1 ORs appear not to overlap thus forming a distinct border
between Z1 and Z2. Later studies confirm and extend our findings on
the overlapping expression of ORs in Z2-4 (Iwema et al., 2004;
Miyamichi et al., 2005). It was suggested, by studying a relatively large
number of ORs, that each OR located in Z2-4 has its own unique
expression area which is overlapping between different ORs, forming
broad but overlapping bands of OR expression (Miyamichi et al., 2005).
The possibility exists that Z1 is inherently different from the other
zones as this zone is the only zone that expresses both class I and class II
ORs (Tsuboi et al., 2006). Z1 and Z2-4 are different from each other
with regard to expression of other genes as well, e.g. expression of
NADPH quinone oxidoreductase 1 (NQO1) in Z1 (Gussing & Bohm,
2004) and neural cell adhesion molecule 2 (NCAM2) in Z2-4 (Alenius &
Bohm, 1997; Yoshihara et al., 1997).
The function of the zonal topography and the mechanisms that control
OSN zonal specification and specific zone projection are up to date not
completely understood. OR zones may be established during early
development when OSNs are generated as the zonal topography is
evident already at embryonic day 13.5 (Sullivan et al., 1995). New OSN
are, however, continuously generated throughout life and the zonal
arrangement maintained. Specification of OSNs, i.e. OR choice, is
independent of the target neurons (Sullivan et al., 1995; Fan & Ngai,
2001). Conceivably, information within and/or in proximity of the
33
olfactory epithelium may continuously contribute to restrict OSNs to
express ORs and axon guidance proteins according to the cell's zonal
localization (e.g. (Alenius & Bohm, 1997), paper I). The fact that Z2-4
ORs display overlapping expression areas indicate that the regulation of
OR choice may not be regulated by zonally distinct transcription factors
but may rather be regulated by gradients of instructive factors. In a
screen aimed at identifying genes displaying expression patterns
correlated to the expression of ORs we have identified several genes.
Msx1, Alk6, retinaldehyde dehydrogenase 2 (RALDH2) and
Neuropilin2 (Npn2) are expressed in a graded manner in or in close
proximity to the olfactory epithelium (fig 8).
Zone-specific expression of ORs is in part regulated by DNA regions
upstream of the OR genes. The correct expression pattern of an OR
(M71) has been implied to depend on only a small DNA region close to
Figure 8. OR expression areas are overlapping in Z2-4 but not in Z1. Alk6
are expressed in sustentacular cells (sus) in a graded manner with high
expression in Z1 and low expression in Z4. Reverse gradients of Npn2, Msx1
and RALDH2 was observed in OSNs, basal cells (bc) and in the lamina
propria, respectively.
34
the transcription start site. This minimal promoter contains an O/E
binding site and a homeobox binding site (Rothman et al., 2005). The
identification of the same common regulatory regions in several OR
genes expressed in different zones indicates that common regulatory
proteins may be involved in the choice of which OR an OSN expresses.
If so, it is unlikely that expression of regulatory factors would follow
zonal borders but may instead by graded over the olfactory epithelium.
A candidate is the homeobox protein Msx1, identified in paper I to be
expressed in a graded manner in the basal cell layer of the epithelium as
this cell layer contains the progenitors of the OSNs.
Retinoic acid (RA) is a molecule crucial to olfactory development (Mic
et al., 2002) that has been implicated in Msx1 regulation (Shen et al.,
1994; Wang & Sassoon, 1995; Song et al., 2004). RA is a possible
candidate for regulating the low dorsomedial to high ventrolateral
expression of Msx1 as the rate-limiting enzyme in RA synthesis,
RALDH2, is expressed in mesenchymal cells as a gradient that follows
the Msx1 gradient. Our data correlates with previous work as cells
responding to RA are located in lateral areas of the olfactory epithelium
(Whitesides et al., 1998). Moreover, RA is suggested to be important for
the formation of lateral structures of the olfactory epithelium during
early embryogenesis (LaMantia et al., 2000). Mature OSNs are
dependent on RA for their survival (Hägglund et al., 2006) and RA
deficiency has been shown to increase cell proliferation in the olfactory
epithelium in rat (Asson-Batres et al., 2003).
The identification of genes with graded expression in the sustentacular
cells, basal cells and in cells in the mesenchyme makes it likely that
graded expression can be observed in OSNs as well. The fact that the
localization of OSNs expressing a given OR strongly influences the
projection sites in the olfactory bulb so that the OR location in the
epithelium determines the dorsal-ventral position of the glomeruli
(Miyamichi et al., 2005) suggests that there may be gradients of
molecules regulating projections to the target glomeruli. We found that
Npn2, involved in axon guidance, is expressed in a gradient in OSNs. It
35
can be hypothesized that different levels of Npn2 may help in guiding
the axons towards their correct glomeruli. Npn2 is however not a major
factor in the zonal topography of projections to the olfactory bulb as
mice without a functional Npn2 gene do not display gross defects in the
guidance of the axons to their correct glomeruli (Walz et al., 2002). The
Npn2 mutant mice however have axons that overshoot their target
glomeruli and instead extend into deeper layers of the olfactory bulb.
Target cells in the olfactory bulb express the ligands Sema3B and
Sema3F (paper I) which mediate repulsive actions when interacting
with Npn2 and may function to stop axons in the right compartment of
the olfactory bulb. Nevertheless, reasons for the graded expression of
Npn2 are currently unknown.
36
Regionally restricted detection of odorants (papers II and
III)
A number of ligand-receptor matches have been made in heterologous
as well as in homologous in vitro systems, e.g. (Zhao et al., 1998; Malnic
et al., 1999). These results suggest that the ORs are relatively broadly
tuned, i.e. a receptor can be activated by quite a range of odorant
molecules and one odorant can activate several different receptor types.
These studies are however made for a limited number of odorants and
ORs and do not take the spatially restricted expression of OR genes in
the olfactory epithelium or the number of responding cells into account.
Electro-olfactograms have shown that the response to odorant exposure
may vary over the olfactory epithelial sheet but if such variation is
correlated to zonal topography has not been addressed (Thommesen &
Doving, 1977; Scott et al., 2000). In papers II and III we have aimed to
analyze for topography of odorant-activated OSNs in their proper
context, i.e. the olfactory epithelium of behaving mice. We show that
this is possible by using c-Fos as an activity marker after odorant
exposure (paper II) and by measuring odorant detection thresholds of
mice subjected to selective ablation of OR zones (paper III).
Odorants induce c-Fos in OSNs in an OR-like pattern
We have exposed live mice to a number of odorants in order to activate
ORs and hence OSNs. The epithelium was analyzed for changes in
expression of genes known to be regulated by neural activity in order to
be able to identify odorant-activated OSNs. c-fos is an immediate-early
gene that has been used as an activity marker in other systems in
response to stimuli (reviewed in (Herdegen & Leah, 1998)). Normally
the epithelium is negative for c-Fos but in paper II we have found that
c-Fos is detectable in the olfactory epithelium after exposure to odorant.
c-Fos is readily distinguished 20 minutes after odorant exposure and
both mRNA and protein can be detected several hours after odorant
37
exposure, indicating that long-lasting changes in gene-regulation occur
in the OSNs as a result of odorant exposure. Preferentially mature OSNs
respond since 90% of cells positive for c-Fos are also OMP positive.
Several odorants with varying chemical structure, such as pyridine,
pyrazine, trimethylphosphite, cyclohexanone and octanal, α-ionone,
benzene and benzaldehyde could induce c-Fos.
It was soon evident that c-Fos positive OSNs are stochastically
distributed in the olfactory epithelium. The expression frequency of a
particular OR is often between 0.1-1% (Ressler et al., 1993). The
number of cells showing induced c-Fos after odorant is similar to the
number of OSNs positive for one OR. OSNs responding with c-Fos
induction are restricted to limited areas of the epithelial sheet, e.g.
benzaldehyde induces c-Fos in areas also expressing Z1 ORs and octanal
induce c-Fos in areas with more laterally expressed ORs (Z4). The zonal
response patterns of benzaldehyde (Z1) and octanal (Z4) are in
accordance with data that M71, expressed in Z1 is responding to
benzaldehyde, and that I7, expressed in Z4 is activated by octanal
(Ressler et al., 1993; Zhao et al., 1998; Bozza et al., 2002; Araneda et al.,
2004). Stimulations of rat olfactory epithelium with benzaldehyde or
benzene result in electro-olfactogram responses in dorsal and medial
areas, respectively (Scott-Johnson et al., 2000). We show that small
changes in chemical structure can result in that the OSNs responding to
odorants by inducing c-Fos are located in different OR expression
zones. This shift in zonal activation can be expected if the odorants
activate unique ORs, as ORs are expressed in a zonally limited manner.
Figure 9A illustrates zones activated in response to odorants used in
paper II.
The similarities between the activation of OSNs after odorant exposure
and the pattern of OR expression indicate that the induction of c-Fos is
a result of activation of one or a few ORs by the odorant. A change in
odorant concentration influences the level of c-Fos in the OSN rather
than the number of responding cells, even if concentrations several
orders of magnitude above the detection threshold are used in our
38
exposure paradigm. Neither is the spatial distribution of c-Fos induction
altered with increasing concentrations of odorant which indicates that
additional ORs are not recruited when odorant concentration is
changed.
The fact that induction of c-Fos after odorant exposure exhibits the
same pattern as the expression of an OR suggests that the odorant used
interacts with one or a few ORs. This is not in line with the view of ORs
as broadly tuned but rather points to ORs as more narrow in their
response to single odorant; at least with regard to long-term cellular
changes such as expression of odorant induced genes.
Zonal ablation and effects on odorant detection
The relatively restricted activation of OSNs after odorant observed in
paper II raises the question if removal of specific OR expression zones in
mice will alter the ability to detect low levels of odorants. If ORs are
relatively broadly tuned then removal of one or a few of the ORs
interacting with the odorant would not result in a large difference for
the animal when attempting to detect an odorant. If, on the other hand,
the ORs are more narrow in their response properties as our c-Fos study
indicates, then removal of specific OR expression zones would affect
odorant detection.
Acute effects of 2,6-dichlorobenzonitrile
In paper III, we have used the herbicide 2,6-dichlorobenzonitrile
(dichlobenil) to selectively destroy specific areas of the olfactory
epithelium. Dichlobenil has previously been shown to damage the
dorsomedial part of the olfactory epithelium in mice by irreversible
binding to Bowman’s glands. Metabolism of dichlobenil by a P450dependent mechanism indirectly results in degeneration of the olfactory
epithelium (Brandt et al., 1990; Brittebo, 1997; Mancuso et al., 1997).
Our study confirms that dichlobenil induces OSN death primarily in the
39
dorsomedial part of the epithelium. In addition we find that a low dose
of dichlobenil results in ablation of areas corresponding to OR
expression Z1. A higher dose results in ablation of larger areas that also
involves Z2 and parts of Z3. The distribution of a Z4 OR, however, was
normal even shortly after the higher dose of dichlobenil.
Effects on odorant detection by removal of OR zones
The spatial restriction of OR expression and the limited MRR of each
OR may have consequences for the animal, e.g. in the case of localized
injury as a result of physical injury or encounters to toxic substances. In
order to analyze for the effects of ablation of OR expression zones on
odorant detection we have analyzed mice treated with dichlobenil and
compared their performance to that of untreated mice. Odorant
detection thresholds of several odorants were measured using an
olfactometer (see Methods section). Both treated and untreated mice
readily learn to perform the task to respond correctly to the stimulus
presented. Dichlobenil treated mice make as few errors as controls
when analyzed for their ability to detect benzene and ethyl acetate.
Dichlobenil treated mice have the same detection threshold as controls
which suggests that they do not have difficulties in learning and
memory, a fact previously shown in rat subjected to dichlobenil
treatment as assessed in a Morris water maze spatial task (Genter et al.,
1996). We show that when testing mice for detection of n-decyl alcohol
and pyridazine, then dichlobenil treated mice are less sensitive than
control mice. ORs activated by n-decyl alcohol are thus likely located in
Z1 and Z2 as mice treated with a high dose are almost anosmic to the
odorant while mice treated with a low dose still can detect n-decyl
alcohol although a higher concentration is needed than for control
mice. Pyridazine on the other hand is detected by ORs located in Z1 as
well as ORs in Z3 and/or Z4 since dichlobenil treatment results in an
increase in detection threshold but does not result in a severe inability
to detect pyridazine. The detection threshold for benzene and ethyl
acetate were not affected by removal of Z1 and Z2, suggesting that
detection at threshold concentrations of these odorants occur in Z3
40
and/or Z4. Figure 9B shows probable location of high affinity ORs for
the odorants tested in the olfactometer.
Zonally restricted activation of OSNs
We have, in two independent studies, shown that activation of OSNs by
odorants is mediated by a limited number of ORs in vivo. OSNs
responding to odorants are located in all four zones where no zone
appears to display higher sensitivity to odorants in general. Our studies
show that there may be substantial differences between zones in their
ability to mediate detection of the same odorant. Although ORs are
Figure 9. (A) Summary of zones containing OSNs activated by different
odorants as observed by induction of c-Fos. The location of identified ORs for
benzaldehyde (M71) and octanal (I7) is located in Z1 and Z4 respectively.
(B) Location of high affinity ORs as determined by studies on detection
threshold after zone-specific removal of OSNs.
41
broadly tuned in heterologous systems it appears that ORs with lower
affinity for a given odorant may not be able to compensate for the loss
of high affinity ORs when attempting to detect odorants at low
concentrations.
42
Regional differences in maintenance of the olfactory
epithelium (papers III and IV)
The life-long function of the olfactory system is dependent on
continuous production of new OSNs to replace damaged or worn-out
neurons. The OSN progenitors, the GBCs, reside in the basal
compartment of the olfactory epithelium. This is most readily observed
when the olfactory epithelium is physically damaged as an increase in
proliferation of the GBCs (e.g. (Schwartz Levey et al., 1991; Suzuki &
Takeda, 1991)). The immature cells produced by the GBCs mature into
OSNs to replace dying neurons with functional neurons able to respond
to olfactory cues. The studies made on proliferation in the rodent
olfactory epithelia have not taken the different zonal identities or
differences in milieu between the zones into account. One may
hypothesize that the intricate structure of the turbinates protruding into
the nasal lumen result in differences in the milieu for the neurons and
hence differences in proliferation and survival. We have examined
proliferation in the mouse olfactory epithelium by utilizing
bromodeoxyuridine incorporation in paper IV.
We demonstrate that the basal cell proliferation is uneven over the
epithelial sheet in the adult and that Z1 has a lower rate of proliferation
compared to Z2-4. The ratio of mature and immature OSNs is not the
same in Z1 as in Z4. Z1 has fewer immature neurons compared to Z4 in
adult mice. The number of immature neurons is in agreement with the
lower proliferation in Z1 compared to Z4 at this age. The epithelium is
however thicker in Z1 and the major part of the Z1 neuronal layer is
composed of mature, OMP positive OSNs. The more mature nature of
Z1 indicates that OSNs are retained longer in the epithelium and pulsechase studies show that the newly born OSNs survive longer in Z1
compared to more lateral areas. RA has been suggested to promote
survival of mature OSNs and to decrease proliferation of progenitor
cells. The rate-limiting enzyme in RA production, RALDH2, is
expressed in a low level in Z1 and a high level in Z4 (paper I) which
43
does not correlate with the high level of proliferation in lateral zones or
the shorter life-span of Z4 OSNs. Intriguingly, the adult pattern of
proliferation in the olfactory epithelium appears at a time period where
olfactory ensheating cells acquire their adult proliferation rate and the
start of elimination of mistargeted axons occurs (e.g. (Zou et al., 2004;
Watanabe et al., 2006)).
It can be speculated that ongoing neurogenesis is an advantage for an
animal in a changing environment where birth of new sensory cells
may help the olfactory system to regulate the sensitivity to new
odorants. It has previously been shown that mice exposed to certain
odorants can become more sensitive to these odorants as determined by
a behavioral assay (Wang et al., 1993). OSNs that are activated have
been suggested to have a strong survival advantage over non-activated
OSNs (e.g. (Zhao & Reed, 2001)). It is thus plausible that activated OSN
are selected for in the olfactory epithelium. By exposing young mice to
odorants we find that newly produced OSNs responding to the specific
odorants are not enriched in the epithelium, suggesting that the
sensitization to odorant occurs at the level of the individual cell e.g. by
up-regulation of components of the signaling transduction cascade. The
period of odorant exposure was approximately three weeks which
corresponds to a time period during which OSNs in Z4 are turned over
but Z1 neurons have slower turn over times. It is thus still possible that
a longer exposure period will result in selection of responding OSNs in
Z1.
Regeneration after dichlobenil treatment
In paper III we show that parts of the olfactory epithelium are able to
regenerate after dichlobenil treatment during a time period of a few
weeks. Some areas do however not recover even after 8 months. The
non-repopulated areas correspond to the areas expressing ORs specific
for the most dorsomedial zone, Z1, as shown using in situ hybridization.
The damage to Z2 and the following regeneration as well as the lack of
regeneration in Z1 suggest that there may be differences between the
44
most dorsomedial zone and the other zones with regard to ability of
cells to respond to the damage and to repopulate the epithelium. The
inability of Z1 to regenerate after dichlobenil is not an effect of lack of
proliferation in Z1 in the adult. Z1 has a small number of proliferating
cells that in time should be able to repopulate the epithelium. It is
instead possible that cells responsive for repopulation are missing in Z1
after dichlobenil treatment.
Mammalian achaete-scute homologue 1 (Mash1) is expressed in a
subpopulation of proliferative cells of the olfactory epithelium (Cau et
al., 1997). Mice lacking a functional Mash1 gene almost completely lack
OSNs which suggests that Mash1 is necessary for OSN formation
(Guillemot et al., 1993). When analyzing for Mash1 expression in
dichlobenil treated mice by in situ hybridization, we have observed that
areas not recovering from dichlobenil treatment are devoid of Mash1
positive cells. Cells expressing Mash1 are however present in Z2 (fig
10). It is not known whether the cells that repopulate the epithelium in
e.g. Z2 are spared from the toxic effects of dichlobenil. It is not unlikely
that cells from more lateral, undamaged areas migrate to repopulate
Figure 10. A population of cells located in the basal part of the epithelium is
positive for Mash1 as seen by in situ hybridization. The area marked by the
dotted line is devoid of Mash1 positive cells and is unable to regenerate
after dichlobenil treatment. The arrow heads point to Mash1 expressing
cells.
45
areas that regenerate but that these fail to migrate over the Z1/2 border.
Cells from the Z2-4 areas may not be able to survive in the region of Z1
OR expression as the sharp boundary between regenerated and nonregenerated areas indicates
that there are differences in
local signals/milieu between
Z1 and the other zones.
During regeneration of the
epithelium some ORs appear
to be expressed in spatially
expanded regions compared
to control mice. This can be
seen in fig 11 as an
expansion of Z3 and Z4 ORs
up to the Z1/2 border as
assessed
by
in
situ
hybridization using OR
probes. This is evident early
during the regeneration
process after a high dose of
dichlobenil.
Figure 11. The distribution of
OR expression in Z3 and Z4 is
expanded in epithelium of
mice treated with dichlobenil
as seen with radioactive in
situ hybridization using OR
probes, signal in white. Lines
demarcate the Z1/Z2 border
and the dorsal extent of Z4,
respectively.
46
The graded expression of genes with possible instructive roles for OSNs
(paper I) may be perturbed early during the regeneration process
leading to expanded regions of OR expression. Later, however, the
normal pattern of OR expression in Z2-4 is restored. OSNs expressing
ORs outside their normal expression area likely project to ectopic
glomeruli that in time are eliminated, aided by sensory experience
(Nakatani et al., 2003; St John & Key, 2003; Kerr & Belluscio, 2006).
47
The olfactory epithelium can be divided into two major
domains
Throughout our studies we noticed differences between epithelial Z1
and the other zones with respect to gene expression (paper I), effect of a
toxic compound (paper III) and tissue homeostasis (paper IV). In
addition we describe an overlap of OR expression between Z2-4
suggesting that these zones are not strictly separated from each other
(paper I). I have in fig 12 illustrated this by listing some of our findings
together with findings made by others. Genes expressed in all cell layers
of the epithelium, as well in the mesechyme, can display expression
areas limited by the Z1/Z2 border. Such genes include e.g. metabolizing
enzymes, axon guidance molecules and transcription factors (Miyawaki
et al., 1996; Alenius & Bohm, 1997; Yoshihara et al., 1997; Norlin &
Berghard, 2001; Oka et al., 2003; Gussing & Bohm, 2004). It is likely
that life-span and maturity of the OSNs in the olfactory epithelium
follow the same zonal pattern. About 50% of all ORs are expressed in
Z1 (Zhang et al., 2004). Combined with the fact that Z1 has a higher
number of mature OSNs compared to more lateral zones suggests
greater diversity of OSNs located in Z1. The large number of differences
between Z1 and the other zones make it likely that there is a division of
the olfactory epithelium into two major domains and this may have a
functional role.
48
Figure 12. Z1 is different from the other zones in regard to expression of ORs
and other genes, response to toxic compound and tissue homeostasis
indicating that Z1 might be inherently different from the other zones. The
location of e.g. gene expression is color coded as insert.
49
CONCLUSIONS
ORs, as well as other genes, display regional limitations of expression in
the olfactory epithelium. The zonally restricted pattern of expression
exists in all cell types of the olfactory epithelium of the mouse. Reasons
for the zonal anatomy are still elusive although the answer is hopefully
not far ahead.
The adult pattern of OR expression is apparent from the time point
where spatial OR expression are first evident in the embryo. The
expression of Z2-4 ORs is overlapping while Z1 ORs appear not to
overlap with ORs from other zones. Zonal arrangement must be
maintained throughout life as there is continuous neurogenesis and
gradients of expression of genes with instructive roles may be involved
in specifying OSNs and thus directly or indirectly be important for
correct zonal OR expression.
An odorant is detected by limited numbers of ORs as suggested by the
induction of the activity marker c-Fos after odorant exposure. In the
adult mice localized injury may result in specific inability to detect
odorants at their normal detection threshold concentration and thus
ORs with lower affinity are not able to compensate for the lack of high
affinity ORs. The damaged epithelium is, after regeneration, displaying
the same zonally limited expression of Z2, Z3 and Z4 ORs as before
damage. This may be a result of gradients of instructive gene expression
that is observed in the olfactory epithelium.
It can be concluded that Z1 OSNs are different from Z2-4 with regard to
toxic responses, gene expression, proliferation and OSN survival which
may open for different functions of Z1 and Z2-4.
50
ACKNOWLEDGEMENTS
51
REFERENSES
Alenius, M. and Bohm, S. (1997). "Identification of a novel neural cell
adhesion molecule-related gene with a potential role in selective axonal
projection." J Biol Chem 272: 26083-26086.
Amoore, J. E. (1982). "Odor theory and odor classification." Fragance
Chemistry Academic Press, Inc.: pp 1-76.
Araneda, R., Peterlin, Z., Zhang, X., Chesler, A. and Firestein, S. (2004). "A
pharmacological profile of the aldehyde receptor repertoire in rat olfactory
epithelium." J Physiol 555: 743-56.
Araneda, R. C., Kini, A. D. and Firestein, S. (2000). "The molecular receptive
range of an odorant receptor." Nature Neurosci 3: 1248-1255.
Asson-Batres, M., Zeng, M., Savchenko, V., Aderoju, A. and McKanna, J.
(2003). "Vitamin A deficiency leads to increased cell proliferation in
olfactory epithelium of mature rats." J Neurobiol 54: 539-54.
Aungst, J. L., Heyward, P. M., Puche, A. C., Karnup, S. V., Hayar, A., Szabo,
G. and Shipley, M. T. (2003). "Centre-surround inhibition among olfactory
bulb glomeruli." Nature 426: 623-629.
Bakalyar, H. A. and Reed, R. R. (1990). "Identification of a specialized
adenylate cyclase that may mediate odorant detection." Science 250:
14031406.
Balasubramanian, S., Lynch, J. W. and Barry, P. H. (1995). "The permeation
of organic cations through cAMP-gated channels in mammalian olfactory
receptor neurons." J Membr Biol 146: 177-91.
Belluscio, L., Gold, G. H., Nemes, A. and Axel, R. (1998). "Mice deficient in
Golf are anosmic." Neuron 20: 69-81.
52
Bodyak, N. and Slotnick, B. (1999). "Performance of mice in an automated
olfactometer: odor detection, discrimination and odor memory." Chem
Senses 24: 637-645.
Boekhoff, I., Inglese, J., Schleicher, S., Koch, W. J., Lefkowitz, R. J. and Breer,
H. (1994). "Olfactory desensitization requires membrane targeting of receptor
kinase mediated by βγ-subunits of heterotrimeric G proteins." J Biol Chem
269: 37-40.
Borisy, F. F., Ronnett, G. V., Cunningham, A. M., Juilfs, D., Beavo, J. and
Snyder, S. H. (1992). "Calcium/calmodulin-activated phosphodiesterase
expressed in olfactory receptor neurons." J Neuroscience 12: 915-923.
Bozza, T., Feinstein, P., Zheng, C. and Kauer, J. S. (2002). "Odorant receptor
expression defines functional units in the mouse olfactory system." J Neurosci
22: 3033-3043.
Bradley, J., Bonigk, W., Yau, K.-W. and Frings, S. (2004). "Calmodulin
permanently associates with rat olfactory CNG channels under native
conditions." Nat Neurosci 7: 705-710.
Bradley, J., Li, J., Davidson, N., Lester, H. A. and Zinn, K. (1994).
"Heteromeric olfactory cyclic nucleotide-gated channels: a subunit that
confers increased sensitivity to cAMP." Proc Natl Acad Sci 91: 8890-8894.
Bradley, J., Reuter, D. and Frings, S. (2001). "Facilitation of CalmodulinMediated Odor Adaptation by cAMP-Gated Channel Subunits." Science 294:
2176-2178.
Brandt, I., Brittebo, E. B., Feil, V. J. and Bakke, J. E. (1990). "Irreversible
binding and toxicity of the herbicide dichlobenil (2,6-dichlorobenzonitrile)
in the olfactory mucosa of mice." Toxicol Appl Pharmacol 103: 491-501.
Brittebo, E. B. (1997). "Metabolism-dependent activation and toxicity of
chemicals in nasal glands." Mutat Res 380: 61-75.
53
Brunet, L. J., Gold, G. H. and Ngai, J. (1996). "General anosmia caused by a
targeted disruption of the mouse olfactory cyclic nucleotide-gated cation
channel." Neuron 17: 681-693.
Buck, L. and Axel, R. (1991). "A novel multigene family may encode odorant
receptors: a molecular basis for odor recognition." Cell 65: 175-187.
Caggiano, M., Kauer, J. S. and Hunter, D. D. (1994). "Globose basal cells are
neuronal progenitors in the olfactory epithelium: a linage analysis using a
replication-incompetent retrovirus." Neuron 13: 339-352.
Calof, A. L. and Chikaraishi, D. M. (1989). "Analysis of neurogenesis in a
mammalian neuro-epithelium: proliferation and differentiation of an
olfactory neuron precursor in vitro." Neuron 3: 115-127.
Camoletto, P., Colesanti, A., Ozon, S., Sobel, A. and Fasolo, A. (2001).
"Expression of stathmin and SCG10 proteins in the olfactory neurogenesis
during development and after lesion in the adulthood." Brain Res Bull 54: 1928.
Carter, L. A., MacDonald, J. L. and Roskams, A. J. (2004). "Olfactory
Horizontal Basal Cells Demonstrate a Conserved Multipotent Progenitor
Phenotype." J Neurosci 24: 5670-5683.
Cattarelli, M., Astic, L. and Kauer, J. S. (1988). "Metabolic mapping of 2deoxyglucose uptake in the rat piriform cortex using computerized image
processing." Brain Research 442: 180-184.
Cau, E., Gradwohl, G., Fode, C. and Guillemot, F. (1997). "Mash1 activates a
cascade of bHLH regulators in olfactory neuron progenitors." Development
124: 1611-21.
Chen, T.-Y. and Yau, K.-W. (1994). "Direct modulation by Ca2+-calmodulin
of cyclic nucleotide-activated channel of rat olfactory receptor neurons."
Nature 368: 545-548.
54
Chen, X., Fang, H. and Schwob, J. E. (2004). "Multipotency of purified,
transplanted globose basal cells in olfactory epithelium." J Comp Neurol 469:
457-74.
Chen, Y., Getchell, M. L., Ding, X. and Getchell, T. V. (1992).
"Immunolocalization of two cytochrome P450 isozymes in rat nasal
chemosensory tissue." Neuroreport 3: 749-52.
Chess, A., Simon, I., Cedar, H. and Axel, R. (1994). "Allelic inactivation
regulates olfactory receptor gene expression." Cell 78: 823-834.
Clevenger, A. C. and Restrepo, D. (2006). "Evaluation of the Validity of a
Maximum Likelihood Adaptive Staircase Procedure for Measurement of
Olfactory Detection Threshold in Mice." Chem Senses 31: 9-26.
Couto, A., Alenius, M. and Dickson, B. J. (2005). "Molecular, Anatomical, and
Functional Organization of the Drosophila Olfactory System." Current
Biology 15: 1535-1547.
Ding, X. X. and Coon, M. J. (1988). "Purification and characterization of two
unique forms of cytochrome P-450 from rabbit nasal microsomes."
Biochemistry 27: 8330-7.
Dzeja, C., Hagen, V., Kaupp, U. B. and Frings, S. (1999). "Ca2+ permeation in
cyclic nucleotide-gated channels." EMBO J 18: 131-44.
Eggan, K., Baldwin, K., Tackett, M., Osborne, J., Gogos, J., Chess, A., Axel, R.
and Jaenisch, R. (2004). "Mice cloned from olfactory sensory neurons."
Nature 428: 44-9.
Elsaesser, R. and Paysan, J. (2005). "Morituri te salutant? Olfactory signal
transduction and the role of phosphoinositides." J Neurocytol 34: 97-116.
Fan, J. and Ngai, J. (2001). "Onset of odorant receptor gene expression during
olfactory sensory neuron regeneration." Dev Biol 229: 119-127.
55
Farbman, A. I. and Margolis, F. L. (1980). "Olfactory marker protein during
ontogeny: immunohistochemical localization." Dev Biol 74: 205-215.
Fried, H. U., Fuss, S. H. and Korsching, S. I. (2002). "Selective imaging of
presynaptic activity in the mouse olfactory bulb shows concentration and
structure dependence of odor responses in identified glomeruli." Proc Natl
Acad Sci U S A 99: 3222-3227.
Genter, M. B., Owens, D. M., Carlone, H. B. and Crofton, K. M. (1996).
"Characterization of olfactory deficits in the rat following administration of
2,6-dichlorobenzonitrile
(dichlobenil),
3,3'-iminodipropionitrile,
or
methimazole." Fund Appl Toxicol 29: 71-77.
Glusman, G., Yanai, I., Rubin, I. and Lancet, D. (2001). "The complete human
olfactory subgenome." Genome Res 11: 685-702.
Godfrey, P. A., Malnic, B. and Buck, L. B. (2004). "The mouse olfactory
receptor gene family." Proc Natl Acad Sci USA 101: 2156-2161.
Gordon, M. K., Mumm, J. S., Davis, R. A., Holcomb, J. D. and Calof, A. L.
(1995). "Dynamics of MASH1 expression in vitro and in vivo suggest a nonstem cell site of MASH1 action in the olfactory receptor neuron lineage." Mol
Cell Neurosci 6: 363-79.
Graziadei, P. P. and Graziadei, G. A. (1979). "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.
Griff, I. C. and Reed, R. R. (1995). "The genetic basis for specific anosmia to
isovaleric acid in the mouse." Cell 83: 407-14.
Guillemot, F., Lo, L.-C., Johnson, J. E., Auerbach, A., Anderson, D. J. and
Joyner, A. L. (1993). "Mammalian achaete-scute homolog 1 is required for the
early development of olfactory and autonomic neurons." Cell 75: 463-476.
56
Gussing, F. and Bohm, S. (2004). "NQO1 activity in the main and the
accessory olfactory systems correlates with the zonal topography of
projection maps." Eur J Neurosci 19: 2511-8.
Haberly, L. B. (1969). "Single unit responses to odor in the prepyriform
cortex of the rat." Brain Res 12: 481-4.
Herdegen, T. and Leah, J. D. (1998). "Inducible and constitutive transcription
factors in the mammalian nervous system: control of gene expression by jun,
fos and krox, and CREB/ATF proteins." Brain Res Rev 28: 370-490.
Hinds, J. W., Hinds, P. L. and McNelly, N. A. (1984). "An autoradiographic
study of the mouse olfactory epithelium: evidence for long-lived receptors."
Anat Rec 210: 375-83.
Holbrook, E. H., Szumowski, K. E. and Schwob, J. E. (1995). "An
immunochemical, ultrastructural, and developmental characterization of the
horizontal basal cells of rat olfactory epithelium." J Comp Neurol 363: 12946.
Hoppe, R., Breer, H. and Strotmann, J. (2006). "Promoter motifs of olfactory
receptor genes expressed in distinct topographic patterns." Genomics 87: 71123.
Horowitz, L. F., Montmayeur, J. P., Echelard, Y. and Buck, L. B. (1999). "A
genetic approach to trace neural circuits." Proc Natl Acad Sci U S A 96: 1949.
Huard, J. M. and Schwob, J. E. (1995). "Cell cycle of globose basal cells in rat
olfactory epithelium." Dev Dyn 203: 17-26.
Huard, J. M. T., Youngentob, S. L., Goldstein, B. J., Luskin, M. B. and
Schwob, J. E. (1998). "Adult olfactory epithelium contains multipotent
progenitors that give rise to neurons and non-neuronal cells." J Comp Neurol
400: 469-486.
57
Hägglund, M., Berghard, A., Strotmann, J. and Bohm, S. (2006). "Retinoic
acid receptor-dependent survival of olfactory sensory neurons in postnatal
and adult mice." J Neurosci 26: 3281-91.
Isaacson, J. S. and Strowbridge, B. W. (1998). "Olfactory reciprocal synapses:
dendritic signaling in the CNS." Neuron 20: 749-761.
Iwema, C. L., Fang, H., Kurtz, D. B., Youngentob, S. L. and Schwob, J. E.
(2004). "Odorant Receptor Expression Patterns Are Restored in LesionRecovered Rat Olfactory Epithelium." J Neurosci 24: 356-369.
Jones, D. T. and Reed, R. R. (1989). "Golf: an olfactory neuron specific Gprotein involved in odorant signal transduction." Science 244: 790-795.
Juilfs, D. M., Fülle, H.-J., Zhao, A. Z., Houslay, M. D., Garbers, D. L. and
Beavo, J. A. (1997). "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 94: 33883395.
Kajiya, K., Inaki, K., Tanaka, M., Haga, T., Kataoka, H. and Touhara, K.
(2001). "Molecular bases of odor discrimination: reconstitution of olfactory
receptors that recognize overlapping sets of odorants." J Neurosci 21: 60186025.
Kaplan, M. S., McNelly, N. A. and Hinds, J. W. (1985). "Population dynamics
of adult-formed granule neurons of the rat olfactory bulb." J Comp Neurol
239: 117-25.
Katada, S., Hirokawa, T., Oka, Y., Suwa, M. and Touhara, K. (2005).
"Structural basis for a broad but selective ligand spectrum of a mouse
olfactory receptor: mapping the odorant-binding site." J Neurosci 25: 180615.
Kaupp, U. B., Niidome, T., Tanabe, T., Terada, S., Bonigk, W., Stuhmer, W.,
Cook, N. J., Kangawa, K., Matsuo, H., Hirose, T., Miyata, T. and Numa, S.
(1989). "Primary structure and functional expression from complementary
58
DNA of the rod photoreceptor cyclic GMP-gated channel." Nature 342: 762766.
Kaupp, U. B. and Seifert, R. (2002). "Cyclic Nucleotide-Gated Ion Channels."
Physiol Rev 82: 769-824.
Kelliher, K. R., Ziesmann, J., Munger, S. D., Reed, R. R. and Zufall, F. (2003).
"Importance of the CNGA4 channel gene for odor discrimination and
adaptation in behaving mice." Proc Natl Acad Sci 100: 4299-4304.
Kerr, M. A. and Belluscio, L. (2006). "Olfactory experience accelerates
glomerular refinement in the mammalian olfactory bulb." Nat Neurosci 9:
484-486.
Kimbell, J. S., Godo, M. N., Gross, E. A., Joyner, D. R., Richardson, R. B. and
Morgan, K. T. (1997). "Computer Simulation of Inspiratory Airflow in All
Regions of the F344 Rat Nasal Passages." Toxicology and Applied
Pharmacology 145: 388-398.
Kleene, S. J. (1993). "Origin of the chloride current in olfactory
transduction." Neuron 11: 123-132.
Kosaka, K. and Kosaka, T. (2005). "synaptic organization of the glomerulus in
the main olfactory bulb: compartments of the glomerulus and heterogeneity
of the periglomerular cells." Anat Sci Int 80: 80-90.
Kubick, S., Strotmann, J., Andreini, I. and Breer, H. (1997). "Subfamily of
Olfactory Receptors Characterized by Unique Structural Features and
Expression Patterns." J Neurochem 69: 465-475.
Kurahashi, T. and Menini, A. (1997). "Mechanism of odorant adaptation in
the olfactory receptor cell." Nature 385: 725-9.
LaMantia, A.-S., Bhasin, N., Rhodes, K. and Heemskerk, J. (2000).
"Mesenchymal/epithelial induction mediates olfactory pathway formation."
Neuron 28: 411-425.
59
Leinders-Zufall, T., Ma, M. and Zufall, F. (1999). "Impaired odor adaptation
in olfactory receptor neurons after inhibition of Ca2+/calmodulin kinase II." J
Neurosci 19: RC19 (1-6).
Leinders-Zufall, T., Rand, M. N., Shepherd, G. M., Greer, C. A. and Zufall, F.
(1997). "Calcium entry through cyclic nucleotide-gated channels in
individual cilia of olfactory receptor cells: spatiotemporal dynamics." J
Neurosci 17: 4136-48.
Lewcock, J. W. and Reed, R. R. (2004). "A feedback mechanism regulates
monoallelic odorant receptor expression." Proc Natl Acad Sci U S A 101:
1069-1074.
Li, J., Ishii, T., Feinstein, P. and Mombaerts, P. (2004). "Odorant receptor
gene choice is reset by nuclear transfer from mouse olfactory sensory
neurons." Nature 428: 393-399.
Li, Y., Field, P. M. and Raisman, G. (2005). "Olfactory ensheathing cells and
olfactory nerve fibroblasts maintain continuous open channels for regrowth
of olfactory nerve fibres." Glia 52: 245-251.
Liman, E. R. and Buck, L. B. (1994). "A second subunit of the olfactory cyclic
nucleotide-gated channel confers high sensitivity to cAMP." Neuron 13: 611621.
Lin, D. Y., Shea, S. D. and Katz, L. C. (2006). "Representation of Natural
Stimuli in the Rodent Main Olfactory Bulb." Neuron 50: 937-949.
Lin, W., Arellano, J., Slotnick, B. and Restrepo, D. (2004). "Odors detected by
mice deficient in cyclic nucleotide-gated channel subunit A2 stimulate the
main olfactory system." J Neurosci 24: 3703-3710.
Liu, D. T., Tibbs, G. R. and Siegelbaum, S. A. (1996). "Subunit Stoichiometry
of Cyclic Nucleotide-Gated Channels and Effects of Subunit Order on
Channel Function." Neuron 16: 983-990.
60
Lomvardas, S., Barnea, G., Pisapia, D. J., Mendelsohn, M., Kirkland, J. and
Axel, R. (2006). "Interchromosomal Interactions and Olfactory Receptor
Choice." Cell 126: 403-413.
Ma, M. and Shepherd, G. M. (2000). "Functional mosaic organization of
mouse olfactory receptor neurons." Proc Natl Acad Sci 97: 12869-12874.
Mackay-Sim, A. and Kittel, P. (1991). "Cell dynamics in the adult mouse
olfactory epithelium: a quantitative autoradiographic study." J Neurosci 11:
979-984.
Malnic, B., Godfrey, P. A. and Buck, L. B. (2004). "The human olfactory
receptor gene family." Proc Natl Acad Sci U S A 101: 2584-9.
Malnic, B., Hirono, J., Sato, T. and Buck, L. B. (1999). "Combinatorial
receptor codes for odors." Cell 96: 713-723.
Mancuso, M., Giovanetti, A. and Brittebo, E. B. (1997). "Effects of dichlobenil
on ultrastructural morphology and cell replication in the mouse olfactory
mucosa." Tox Pathol 25: 186-194.
Margrie, T. W., Sakmann, B. and Urban, N. N. (2001). "Action potential
propagation in mitral cell lateral dendrites is decremental and controls
recurrent and lateral inhibition in the mammalian olfactory bulb." Proc Natl
Acad Sci U S A 98: 319-324.
Matulionis, D. H. (1975). "Ultrastructural study of mouse olfactory
epithelium following destruction by ZnSO4 and its subsequent regeneration."
Am J Anat 142: 67-89.
Menco, B. P. M. (1980). "Qualitative and quantitative freeze-fracture studies
on olfactory and nasal respiratory epithelial surfaces of frog, ox, rat, and dog
III. Tight-junctions." Cell Tissue Res 211: 361-373.
Mendoza, A. S. (1993). "Morphological studies on the rodent main and
accessory olfactory systems: the regio olfactoria and vomeronasal organ."
Ann Anat 175: 425-446.
61
Meyer, M. R., Angele, A., Kremmer, E., Kaupp, B. and Muller, F. (2000). "A
cGMP-signaling pathway in a subset of olfactory sensory neurons." Proc Natl
Acad Sci 97: 10595-10600.
Mic, F., Haselbeck, R., Cuenca, A. and Duester, G. (2002). "Novel retinoic
acid generating activities in the neural tube and heart identified by
conditional rescue of Raldh2 null mutant mice." Development 129: 2271-82.
Miragall, F., Krause, D., de Vries, U. and Dermietzel, R. (1994). "Expression
of the tight junction protein ZO-1 in the olfactory system: presence of ZO-1
on olfactory sensory neurons and glial cells." J Comp Neurol 341: 433-48.
Miyamichi, K., Serizawa, S., Kimura, H. M. and Sakano, H. (2005).
"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.
Miyawaki, A., Homma, H., Tamura, H., Matsui, M. and Mikoshiba, K. (1996).
"Zonal distribution of sulfotransferase for phenol in olfactory sustentacular
cells." EMBO J 15: 2050-2055.
Moon, C., Simpson, P. J., Tu, Y., Cho, H. and Ronnett, G. V. (2005).
"Regulation of intracellular cyclic GMP levels in olfactory sensory neurons." J
Neurochem 95: 200-209.
Mori, K., Nagao, H. and Yoshihara, Y. (1999). "The olfactory bulb: coding
and processing of odor molecule information." Science 286: 711-715.
Mori, K., Takahashi, Y. K., Igarashi, K. M. and Yamaguchi, M. (2006). "Maps
of Odorant Molecular Features in the Mammalian Olfactory Bulb." 86: 409433.
Munger, S. D., Lane, A. P., Zhong, H., Leinders-Zufall, T., Yau, K.-W.,
Zufall, F. and Reed, R. R. (2001). "Central role of the CNGA4 channel
subunit in Ca2+-calmodulin-dependent odor adaptation." Science 294: 21722175.
62
Nakamura, T. and Gold, G. H. (1987). "A cyclic nucleotide-gated
conductance in olfactory receptor cilia." Nature 325: 442-444.
Nakatani, H., Serizawa, S., Nakajima, T., Imai, T. and Sakano, H. (2003).
"Developmental elimination of ectopic projection sites for the transgenic OR
gene that has lost zone specificity in the olfactory epithelium." Eur J Biochem
18: 2425-2432.
Neuhaus, E. M., Mashukova, A., Zhang, W., Barbour, J. and Hatt, H. (2006).
"A Specific Heat Shock Protein Enhances the Expression of Mammalian
Olfactory Receptor Proteins." Chem Senses 31: 445-452.
Ngai, J., Chess, A., Dowling, M. M., Necles, N., Macagno, E. R. and Axel, R.
(1993a). "Coding of olfactory information: topography of odorant receptor
expression in the catfish olfactory epithelium." Cell 72: 667-680.
Ngai, J., Dowling, M. M., Buck, L., Axel, R. and Chess, A. (1993b). "The
family of genes encoding odorant receptors in the channel catfish." Cell 72:
657-666.
Nomura, T., Takahashi, S. and Ushiki, T. (2004). "Cytoarchitecture of the
normal rat olfactory epithelium: Light and scanning electron microscopic
studies." Arch Histol Cytol 67: 159-170.
Norlin, E. M. and Berghard, A. (2001). "Spatially restricted expression of
regulators of G-protein signaling (RGS) in primary olfactory neurons." Mol
Cell Neurosci 17: 872-882.
Oka, Y., Kobayakawa, K., Nishizumi, H., Miyamichi, K., Hirose, S., Tsuboi,
A. and Sakano, H. (2003). "O-MACS, a novel member of the medium-chain
acyl-CoA synthetase family, specifically expressed in the olfactory
epithelium in a zone-specific manner." Eur J Biochem 270: 1995-2004.
Pace, U., Hanski, E., Salomon, Y. and Lancet, D. (1985). "Odorant-sensitive
adenylate cyclase may mediate olfactory reception." Nature 316: 255-258.
63
Peppel, K., Boekhoff, I., McDonald, P., Breer, H., Caron, M. G. and
Lefkowitz, R. J. (1997). "G protein-coupled receptor kinase 3 (GRK3) gene
disruption leads to loss of odorant receptor desensitization." J Biol Chem 272:
25425-25428.
Pilpel, Y. and Lancet, D. (1999). "The variable and conserved interfaces of
modeled olfactory receptor proteins." Protein Sci 8: 969-977.
Qasba, P. and Reed, R. R. (1998). "Tissue and zonal-specific expression of an
olfactory receptor transgene." J Neurosci 18: 227-236.
Rall, W., Shepherd, G. M., Reese, T. S. and Brightman, M. W. (1966).
"Dendrodendritic synaptic pathway for inhibition in the olfactory bulb."
Experimental Neurology 14: 44-56.
Rash, J., Davidson, K., Kamasawa, N., Yasumura, T., Kamasawa, M., Zhang,
C., Michaels, R., Restrepo, D., Ottersen, O., Olson, C. and Nagy, J. (2005).
"Ultrastructural localization of connexins (Cx36, Cx43, Cx45), glutamate
receptors and aquaporin-4 in rodent olfactory mucosa, olfactory nerve and
olfactory bulb." J Neurocytol 35: 307-41.
Rawson, N. E., Eberwine, J., Dotson, R., Jackson, J., Ulrich, P. and Restrepo,
D. (2000). "Expression of mRNAs encoding for two different olfactory
receptors in a subset of olfactory receptor neurons." J Neurochem 75: 185195.
Reisert, J., Bauer, P. J., Yau, K. W. and Frings, S. (2003). "The Ca-activated Cl
channel and its control in rat olfactory receptor neurons." J Gen Physiol 122:
349-63.
Ressler, K. J., Sullivan, S. L. and Buck, L. B. (1993). "A zonal organization of
odorant receptor gene expression in the olfactory epithelium." Cell 73: 597609.
Ressler, K. J., Sullivan, S. L. and Buck, L. B. (1994). "Information coding in
the olfactory system: evidence for a stereotyped and highly organized epitope
map in the olfactory bulb." Cell 79: 1245-1256.
64
Reuter, D., Zierold, K., Schröder, W. H. and Frings, S. (1998). "A depolarizing
chloride current contributes to chemoelectrical transduction in olfactory
sensory neurons in situ." J Neurosci 18: 6623-6630.
Rothman, A., Feinstein, P., Hirota, J. and Mombaerts, P. (2005). "The
promoter of the mouse odorant receptor gene M71." Mol Cell Neurosci 28:
535-46.
Royet, J. P., Jourdan, F., Ploye, H. and Souchier, C. (1989). "Morphometric
modifications associated with early sensory experience in the rat olfactory
bulb: II. Stereological study of the population of olfactory glomeruli." J Comp
Neurobiol 289: 594-609.
Rubin, B. D. and Katz, L. C. (1999). "Optical imaging of odorant
representations in the mammalian olfactory bulb." Neuron 23: 499-511.
Saito, H., Kubota, M., Roberts, R. W., Chi, Q. and Matsunami, H. (2004).
"RTP family members induce functional expression of mammalian odorant
receptors." Cell 119: 679-691.
Sato, T., Hirono, J., Tonoike, M. and Takebayashi, M. (1994). "Tuning
specificities to aliphatic odorants in mouse olfactory receptor neurons and
their local distribution." J Neurophysiol 72: 2980-2989.
Schoenfeld, T. and Cleland, T. (2005). "The anatomical logic of smell." Trends
Neurosci. 28: 620-7.
Schoppa, N. and Urban, N. (2003). "Dendritic processing within olfactory
bulb circuits." Trends Neurosci 26: 501-6.
Schwartz Levey, M., Chikaraishi, D. M. and Kauer, J. S. (1991).
"Characterization of potential precursor populations in the mouse olfactory
epithelium using immunohistochemistry and auroradiography." J Neurosci
11: 3556-3564.
65
Schwob, J. E. (2002). "Neural regeneration and the peripheral olfactory
system." Anat Rec 269: 33-49.
Schwob, J. E., Youngentob, S. L. and Mezza, R. C. (1995). "Reconstitution of
the rat olfactory epithelium after methyl bromide-induced lesion." J Comp
Neurol 359: 15-37.
Scott-Johnson, P., Blakley, D. and Scott, J. (2000). "Effects of air flow on rat
electroolfactogram." Chem Senses. 25: 761-8. .
Scott, J. W., Brierley, T. and Schmidt, F. H. (2000). "Chemical determinants
of the rat electro-olfactogram." J. Neurosci. 20: 4721-4731.
Serizawa, S., Ishii, T., Nakatani, H., Tsuboi, A., Nagawa, F., Asano, M., Sudo,
K., Sagakami, J., Sakano, H., Ijiri, T., Matsuda, Y., Suzuki, M., Yamamori, T.,
Iwakura, Y. and Sakano, H. (2000). "Mutually exclusive expression of odorant
receptor transgenes." Nature Neurosci 3: 687-693.
Serizawa, S., Miyamichi, K., Nakatani, H., Suzuki, M., Saito, M., Yoshihara,
Y. and Sakano, H. (2003). "Negative Feedback Regulation Ensures the One
Receptor-One Olfactory Neuron Rule in Mouse." Science 302: 2088-2094.
Shen, R., Chen, Y., Huang, L., Vitale, E. and Solursh, M. (1994).
"Characterization of the human MSX-1 promoter and an enhancer
responsible for retinoic acid induction." Cell Mol Biol Res 40: 297-312.
Singer, M. S. and Shepherd, G. M. (1994). "Molecular modeling of ligandreceptor interactions in the OR5 olfactory receptor." Neuroreport 5: 12971300.
Slotnick, B. and Bodyak, N. (2002). "Odor discrimination and odor quality
perception in rats with disruption of connections between the olfactory
epithelium and olfactory bulb." J Neurosci 22: 4205-4216.
Slotnick, B., Cockerham, R. and Pickett, E. (2004). "Olfaction in olfactory
bulbectomized rats." J Neurosci 24: 9195-200.
66
Song, Y., Hui, J., Fu, K. and Richman, J. (2004). "Control of retinoic acid
synthesis and FGF expression in the nasal pit is required to pattern the
craniofacial skeleton." Dev Biol 276: 313-29.
Sosinsky, A., Glusman, G. and Lancet, D. (2000). "The genomic structure of
human olfactory receptor genes." Genomics 70: 49-61.
St John, J. A. and Key, B. (2003). "Axon mis-targeting in the olfactory bulb
during regeneration of olfactory neuroepithelium." Chem Senses 28: 773-779.
Strotmann, J., Conzelmann, S., Beck, A., Feinstein, P., Breer, H. and
Mombaerts, P. (2000). "Local permutations in the glomerular array of the
mouse olfactory bulb." J Neurosci 15: 6927-6938.
Strotmann, J., Hoppe, R., Conzelmann, S., Feinstein, P., Mombaerts, P. and
Breer, H. (1999). "Small subfamily of olfactory receptor genes: structural
features, expression pattern and genomic organization." Gene 236: 281-291.
Strotmann, J., Wanner, I., Krieger, J., Raming, K. and Breer, H. (1992).
"Expression of odorant receptors in spatially restricted subsets of
chemosensory neurones." Neuroreport 3: 1053-1056.
Sullivan, S. L., Adamson, M. C., Ressler, K. J., Kozak, C. A. and Buck, L. B.
(1996). "The chromosomal distribution of mouse odorant receptor genes."
Proc Natl Acad Sci USA 93: 884-888.
Sullivan, S. L., Bohm, S., Ressler, K. J., Horowitz, L. F. and Buck, L. B. (1995).
"Target-independet pattern specification in the olfactory epithelium."
Neuron: 779-789.
Suzuki, Y., Schafer, J. and Farbman, A. I. (1995). "Phagocytic Cells in the Rat
Olfactory Epithelium after Bulbectomy." Experimental Neurology 136: 225233.
Suzuki, Y. and Takeda, M. (1991). "Basal cells in the mouse olfactory
epithelium after axotomy: immunohistochemical and electron-microscopic
studies." Cell Tissue Res 266: 239-45.
67
Tanabe, T., Iino, M. and Takagi, S. F. (1975). "Discrimination of odors in
olfactory bulb, pyriform-amygdaloid areas, and orbitofrontal cortex of the
monkey." J Neurophysiol 38: 1284-1296.
Thommesen, G. and Doving, K. B. (1977). "Spatial distribution of the EOG in
the rat; a variation with odour quality." Acta Physiol Scand 99: 270-280.
Trudeau, M. C. and Zagotta, W. N. (2003). "Calcium/Calmodulin Modulation
of Olfactory and Rod Cyclic Nucleotide-gated Ion Channels." J Biol Chem
278: 18705-18708.
Tsuboi, A., Miyazaki, T., Imai, T. and Sakano, H. (2006). "Olfactory sensory
neurons expressing class I odorant receptors converge their axons on an
antero-dorsal domain of the olfactory bulb in the mouse." Eur J Neurosci 23:
1436-44.
Uchida, N., Takahashi, Y. K., Tanifuji, M. and Mori, K. (2000). "Odor maps in
the mammalian olfactory bulb: domain organization and odorant structural
features." Nature Neurosci 3: 1035-1043.
Wachowiak, M. and Cohen, L. B. (2001). "Representation of odorants by
receptor neuron input to the mouse olfactory bulb." Neuron 32: 723-735.
Walz, A., Rodriguez, I. and Mombaerts, P. (2002). "Aberrant sensory
innervation of the olfactory bulb in neuropilin-2 mutant mice." J Neurosci
22: 4025-4035.
Wang, H.-W., Wysocki, C. J. and Gold, G. H. (1993). "Induction of olfactory
receptor sensitivity in mice." Science 260: 998-1000.
Wang, Y. and Sassoon, D. (1995). "Ectoderm-mesenchyme and mesenchymemesenchyme interactions regulate Msx-1 expression and cellular
differentiation in the murine limb bud." Dev Biol 168: 374-82.
68
Vassar, R., Chao, S. K., Sitcheran, R., Nunez, J. M., Vosshall, L. B. and Axel,
R. (1994). "Topographic organization of sensory projections to the olfactory
bulb." Cell 79: 981-992.
Vassar, R., Ngai, J. and Axel, R. (1993). "Spatial segregation of odorant
receptor expression in the mammalian olfactory epithelium." Cell 74: 309318.
Watanabe, K., Kondo, K., Takeuchi, N., Nibu, K. and Kaga, K. (2006). "Agerelated changes in cell density and the proliferation rate of olfactory
ensheathing cells in the lamina propria of postnatal mouse olfactory mucosa."
Brain Res doi:10.1016/j.brainres.2006.07.124
Wei, J., Wayman, G. and Storm, D. R. (1996). "Phosphorylation and
Inhibition of Type III Adenylyl Cyclase by Calmodulin-dependent Protein
Kinase II in Vivo." J Biol Chem 271: 24231-24235.
Wei, J., Zhao, A. Z., Chan, G. C. K., Baker, L. P., Impey, S., Beavo, J. A. and
Storm, D. R. (1998). "Phosphorylation and inhibition of olfactory adenylyl
cyclase by CaM kinase II in neurons: a mechanism for attenuation of
olfactory signals." Neuron 21: 495-504.
Whitesides, J., Hall, M., Anchan, R. and LaMantia, A.-S. (1998). "Retinoid
signaling distinguishes a subpopulation of olfactory receptor neurons in the
developing and adult mouse." J. Comp. Neurol. 394: 445-461.
Whitesides, J. G. r. and LaMantia, A. S. (1996). "Differential adhesion and the
initial assembly of the mammalian olfactory nerve." J Comp Neurol 373: 24054.
Wong, S. T., Trinh, K., Hacker, B., Chan, G. C. K., Xia, Z., Gold, G. H. and
Storm, D. R. (2000). "Disruption of the type III adenylyl cyclase gene leads to
peripheral and behavioral anosmia in transgenic mice." Neuron 27: 487-497.
Vucinic, D., Cohen, L. B. and Kosmidis, E. K. (2006). "Interglomerular
Center-Surround Inhibition Shapes Odorant-Evoked Input to the Mouse
Olfactory Bulb In Vivo." J Neurophysiol 95: 1881-1887.
69
Wysocki, C. J., Whitney, G. and Tucker, D. (1977). "Specific anosmia in the
laboratory mouse." Behav Genet 7: 171-188.
Yan, C., Zhao, A. Z., Bentley, J. K., Loughney, K., Ferguson, K. and Beavo, J.
A. (1995). "Molecular cloning and characterization of a calmodulindependent phosphodiesterase enriched in olfactory sensory neurons." Proc
Natl Acad Sci 92: 9677-9681.
Yoshihara, Y., Kawasaki, M., Tamada, A., Fujita, H., Hayashi, H.,
Kagamiyama, H. and Mori, K. (1997). "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.
Yu, T. T., McIntyre, J. C., Bose, S. C., Hardin, D., Owen, M. C. and
McClintock, T. S. (2005). "Differentially expressed transcripts from
phenotypically identified olfactory sensory neurons." J Comp Neurol 483:
251-262.
Zhang, X. and Firestein, S. (2002). "The olfactory receptor gene superfamily
of the mouse." Nature Neurosci 5: 124-133.
Zhang, X., Rogers, M., Tian, H., Zhang, X., Zou, D.-J., Liu, J., Ma, M.,
Shepherd, G. M. and Firestein, S. J. (2004). "High-throughput microarray
detection of olfactory receptor gene expression in the mouse." Proc Natl
Acad Sci U S A 101: 14168-14173.
Zhao, H., Ivic, L., Otaki, J. M., Hashimoto, M., Mikoshiba, K. and Firestein, S.
(1998). "Functional expression of a mammalian odorant receptor." Science
279: 237-242.
Zhao, H. and Reed, R. R. (2001). "X inactivation of the OCNC1 channel gene
reveals a role for activity-dependent competition in the olfactory system."
Cell 104: 651-660.
70
Zhao, K., Dalton, P., Yang, G. and Scherer, P. (2006). "Numerical modeling of
turbulent and laminar airflow and odorant transport during sniffing in the
human and rat nose." Chem Senses. 31: 107-18.
Zheng, J. and Zagotta, W. N. (2004). "Stoichiometry and Assembly of
Olfactory Cyclic Nucleotide-Gated Channels." Neuron 42: 411-421.
Zou, D.-J., Feinstein, P., Rivers, A. L., Mathews, G. A., Kim, A., Greer, C. A.,
Mombaerts, P. and Firestein, S. (2004). "Postnatal refinement of peripheral
olfactory projections." Science 304: 1976-1979.
Zou, Z. and Buck, L. (2006). "Combinatorial effects of odorant mixes in
olfactory cortex." Science 311: 1477-81.
Zou, Z., Horowitz, L. F., Montmayeur, J.-P., Snapper, S. and Buck, L. B.
(2001). "Genetic tracing reveals a stereotyped sensory map in the olfactory
cortex." Nature 414: 173-179.
Zou, Z., Li, F. and Buck, L. B. (2005). "Odor maps in the olfactory cortex."
Proc Natl Acad Sci U S A 102: 7724-7729.
Zufall, F. and Leinders-Zufall, T. (1997). "Identification of a Long-Lasting
Form of Odor Adaptation that Depends on the Carbon Monoxide/cGMP
SecondMessenger System." J Neurosci 17: 2703-2712.
Zufall, F. and Leinders-Zufall, T. (2000). "The cellular and molecular basis of
odor adaptation." Chem Senses 25: 473-481.
71