• Keine Ergebnisse gefunden

Olfactory transduction is the creation of an electrical signal in the membrane of a sensory cell in response to the arrival of an odorant molecule. The nomenclature varies in the literature, so that transduction may refer to an initial graded change in membrane conductance or to the complete process leading to action potentials in the axon. Nevertheless, the crucial initial step in olfactory transduction is always binding of the odor to an olfactory receptor (OR).

In vertebrates and the nematode Caenorhabditis elegans, binding of an appropriate ligand to a G-protein-coupled

receptor changes the conformation of the receptor and leads to activation of G-proteins and generation of second messengers (Ha and Smith 2008). In insects the mechanisms of signal transduction are under debate. There are only moderate sequence similarities between the fruit fly’s ORs and other G-protein-coupled receptors.

However, mutations affecting different second messenger pathways (cAMP, IP3) lead to a change in olfactory perception (Wicher et al. 2008, 2009; Kain et al.

2008). In addition the presence of different G proteins or G protein subunits were shown (Boto et al. 2010; Chatterjee et al. 2009; Laue et al 1997; Schmidt et al.1998; Talluri et al. 1995; Tanoue et al.

2008)

It is thought that OR83b combines with other ORs to function as heteromeric complexes, but it remains unclear whether these heteromers operate independently as direct ligand gated ion channels, or if they are also G-protein activated (Fig. 3).

Figure
3:
Models
of
olfactory
transduction.
(a)
Direct
ligand‐gated
model
(Sato
et
al.
2008).
Direct
ligand
gated
 nonselective
cation
channels
are
activated
rapidly
by
stimulation
with
very
high
concentrations
of
odor
in
the
 absence
of
G‐protein
signaling.
(b)

G‐protein‐coupled
model
(Wicher
et
al.
2008).
In
this
model
ORx
is
a
G‐

protein‐coupled
receptor
and
OR83b
is
a
cyclic
nucleotide
gated
ion
channel.
Odor
activation
leads
to
a
faster
 ionotropic
pathway
and
a
slow
pathway
that
involves
G‐protein‐coupling
of
ORx,
leading
to
the
production
of
 intracellular
 cAMP,
 which
 activates
 OR83b.
 Abbreviations:
 AC,
 adenylate
 cyclase;
 Gs,
 stimulatory
 G‐protein;


cAMP,
cyclic
adenosine
monophosphate;
OR,
olfactory
receptor
(Nakagawa
2009,
modified)

pathway has the advantage of a very fast reaction, while a second messenger dependent (metabotropic) pathway provides a higher sensitivity to the system.

Another important feature of odorant receptors are a set of proteins in the receptor lymph that are essential for odorant-receptor binding. Relatively little is known about the detailed functional properties of these odorant binding proteins in OSNs that detect fruit odors.

They are probably essential for transferring odorants across the hydrophobic to hydrophilic interface and have been hypothesized to play important roles in olfactory neuron dynamics (Kaissling 2009).

In OSNs that detect pheromone (11-cis vaccenyl acetate) the neurons are not directly activated by the volatile ligand, but instead are activated by the pheromone- pheromone binding protein complex. A mutation in the lush gene that leads to a lack of the odorant binding protein LUSH the flies are behaviorally and electrically unresponsive to the pheromone (Xu et al. 2005).

Odor plumes

An odor is a chemical compound that animals can perceive with their olfactory

Chemical molecules become airborne as they are released from their solid or liquid structures. In a stable atmosphere without turbulence an odor concentration gradient forms around the source and could lead animals to the odor source by simply following the gradient. In the natural environment structurally complex odor plumes are created as the wind takes volatile odor molecules away from their sources (Murlis et al. 1992). These time dependent properties are crucial for insects in finding odor sources.

Figure 4: Paired laser -induced fluorescence of an odor jet. Image is false colored so that concentration increases from blue (lowest) to red (highest) (Weissburg 2000, modified).

Inside an odor plume, filaments of low and high concentrations, and all ranges in between, intersperse with areas in the plume where odor is totally absent (Vetter et al. 2006) (Fig. 4).

In addition to these biophysical properties described above, the insect itself influences the spatiotemporal sequences

of odors that are detected by the OSNs. In moth it was found that the wing beat frequencies are also playing a role in odor detection. Each down stroke of the wings accelerates the airflow over the antennae and therefore the sensilla as well (Tripathy et al. 2010).

Hence, in order to find odor sources animals need to be able follow these spatiotemporal sequences of low and high odor concentration.

Odor stimulation

Some insects are known to rely on the time dependent properties of odor plumes to find the odor source. For example, male moths following a pheromone trace leading to a female will only follow a plume that contains rapid concentration changes (Justus et al. 2005; Vickers et al.

2001; Vickers 2006). Other examples are blood feeding insects that follow CO2

plumes to find mammalian hosts. The structure of an odor plume influences the up wind flight behavior of mosquitoes, and some haematophagous insects only followed an odor plume when the CO2

pulsations were similar to those of the human breathing range (Geier et al. 1999;

Barrozo and Lazzari 2006).

Creating such complex spatiotemporal stimuli is so sufficiently difficult that many

physiological experiments ignore this aspect of the odor plume and only characterize the responses as functions of odor identity and concentration. A commonly used approach to stimulating olfactory receptors is to deliver short odor pulses and observe the responses during the stimulus and for a short time afterwards. Because of the difference between this type of stimulus and a natural odor stimulus, which may include rich spatiotemporal information, it is very likely that such experiments do not reveal the whole dynamic repertoire of OSN responses.

The difficulty of measuring odor concentration has meant that even those studies that attempt to include dynamic information have usually relied on brief pulsatile stimuli. Time duration of those stimuli ranged from 100ms to 5 sec (Hallem and Carlson 2006). However, recent advances in gas phase measurements have shown that the resultant concentration changes at the animal may be very different from those that were anticipated (Vetter et al. 2006).

In this work, recent advances in gas phase instrumentation (photo ionization detector: PID) were used to stimulate

measured, odor concentrations to changing, but as accurately as possible olfactory receptive organs with rapidly

strongly time dependent properties.

Stimuli measured by the PID ranged between 0 and 330 Hz.

Analysis methodology

A technique called linear system analysis (LSA) provides an established general method of measuring the dynamic properties of input–output systems (Bendat and Piersol 1980). LSA has been used in previous studies to investigate dynamic properties of different types of receptors e.g. mechano-, photo- and chemoreceptors (Juusola and French 1997; Justus et al. 2005). In this work LSA was used to examine the dynamic responses of Drosophila melanogaster antenna and olfactory sensilla, respectively.

When a linear system receives a sinusoidal input it produces a sinusoidal output of the same frequency, but possibly differing in amplitude (Gain) and time shift (Phase).

Any linear dynamic system can be completely characterized by the measurement of these amplitude and phase changes, known as its frequency response function. Frequency response analysis (FRA) is a technique for characterizing the dynamic properties of

yellow highlight in Fig. 5).

Figure 5: Visualization of steps necessary to calculate frequency response and view them in a bode plot.

The exact physical nature of the system is not important, only that its input and output signals can be represented as single-valued functions of time. Examples would be the height of water in a bathtub (the output signal) versus the rate of water flowing into the bathtub (the input signal), or the electrical charge on a capacitor (output signal) as a function of net current flowing into and out of it in an electronic circuit (the input signal). In a

manner that depends not only on the input (which can also vary with time) but on the properties of the physical system (e.g. shape of the bathtub) and the memory of the system (e.g. how fast water was flowing into the bathtub at all times in the past). The simplest and most common type of frequency response analysis assumes that the system is linear and stationary. Linearity means that doubling the input signal will double the output signal. Stationary means that the properties of the system (e.g. shape of the bath) do not change with time.

Almost any single valued function of time (such as the height of water in the bath versus time) can be represented by the sum of a set of sinusoidal functions of time (the Fourier theory). A sinusoid in the input signal will always produce a sinusoid of the same frequency in the output, although the amplitude of the output sinusoid and the times of its peaks and troughs (the phase) relative to the input will usually be different. Since any signal can be considered to be a sum of different sinusoids, and any sinusoidal input produces a sinusoidal output, it follows that measuring how each sinusoidal frequency is affected in amplitude and phase will allow to predict the output signal for any input signal.

Measuring the frequency response of an

unknown system can simply be done by subjecting it to a series of sinusoidal inputs at different frequencies, measuring the resultant outputs and calculating the change in amplitude and time shift versus frequency. Stimulating an unknown system with “white noise” (a random unpredictable signal containing sinusoids of all possible frequencies within a certain frequency band) allows the entire frequency response to be measured from a single experiment. The different sinusoidal frequencies in the input and output signals are then separated by the Fourier transform and the changes from input to output compared as if they had each been used separately.

An example for a low-pass frequency response would be a bath, where rapid sinusoidal fluctuations of the water flowing into the bath (e.g. 100 cycles per second) would make very small changes in the height of the bath water versus time. Slow fluctuations (e.g. 0.001 cycles per second) would produce large changes in the height versus time. The bath frequency response giving a smaller output as the frequency rises. The frequency response is a representation of the system's response to a sinusoidal input at different frequencies. The output of the system to the input is a sinusoid of the same frequency but possibly with a

to view the frequency response of a system is via a Bode plot (Fig. 5).

In the present work the unknown system was the fly. How does the antenna or single sensilla respond to odor stimulation at different frequencies?

To investigate this question random noise was used as the input signal and the recorded biological signal (EAGs or action potentials, respectively) as the output signal. To avoid aliasing the action potentials were digitally filtered by convolution with a sinc function (sin(x)/x) (French and Holden 1971). To limit the bandwidth, the signal was down sampled to 100 Hz. The sampled data (tracer gas concentration as input and action potentials as output) were transferred from the time domain into the frequency domain via the Fast Fourier transform (FFT) (Cooley and Tukey 1965) into segments of sample pairs. Response functions between the PID voltage and action potentials were calculated by direct spectral estimation and plotted as Bode plots of phase and log gain versus log frequency (Bendat and Piersol 1980).

The upper part of the Bode plot shows the relation of the output to the input signal.

It shows how strong the response is at different frequencies. A response that is fitted by a low pass filter function indicates that its power remains the same

becomes less sensitive at higher frequencies. A band pass filter response indicates that the sensitivity of the system peaks at certain frequencies and is lower at lower and higher frequencies. The filter function time constants indicate at which frequencies the changes in the rising or falling of the response occurs. The lower part of the Bode plot shows the phase shift (time shift) between input (stimulus) and the output (action potential) signal.

Aims of the thesis

It is clear that temporal changes in odor concentration are vitally important to many animals, but this raises many questions. How are such temporal changes detected? How does temporal sensitivity vary amongst the large number of odors that an animal may detect?

Which stages of the odor detection cascade are time-dependent? Which stages limit the temporal changes that can be detected? How are dynamic signals coded into action potentials and processed in higher centers?

Answering these questions requires accurate methods of delivering dynamically changing spatiotemporal odor stimuli and quantitatively characterizing the resultant neural responses. My thesis describes how an accurate dynamic

odorant stimulation and analysis system was developed and tested in the fruit fly.

Then the system was used and to provide an initial characterization of primary odorant receptor cells of D. melanogaster at the single neuron level. These are early, but crucial, steps in answering some of the important questions about kinetics in insect odor detection posed above.

Literature

Barrozo RB, Kaissling KE. Repetitive stimulation of olfactory receptor cells in female silkmoths Bombyx mori L. J Insect Physiol 2002; 48(8): 825-834.

Barrozo RB, Lazzari CR. Orientation response of haematophagous bugs to CO2: the effect of the temporal structure of the stimulus. J Comp Physiol A 2006;192: 827-31.

Bau J, Justus KA, Loudon C, Carde RT.

Electroantennographic resolution of pulsed pheromone plumes in two species of moths with bipectinate antennae. Chem Senses 2005; 30:

771-80.

Bendat JS, Piersol AG. Engineering

applications of correlation and spectral analysis. New York: John Wiley, 1980;

1–302.

Benton R, Sachse S, Michnick SW, Vosshall LB. Atypical membrane topology and heteromeric function of Drosophila odorant receptors in vivo.

PLoS Biol 2006a; 4(2): 20.

Benton R. On the ORigin of smell: odorant receptors in insects. Cell Mol Life Sci 2006b; 63(14): 1579-85.

Benton R, Vannice KS, Vosshall LB. An essential role for a CD36-related receptor in pheromone detection in Drosophila. Nature 2007; 450(7167):

289-93.

Benton R. Molecular basis of odor detection

in insects. Ann NY Acad Sci 2009a;

1170: 478-81.

Benton R. Eppendorf winner. Evolution and revolution in odor detection. Science 2009b; 326(5951): 382-3.

Benton R, Vannice KS, Gomez-Diaz C, Vosshall LB. Variant ionotropic

glutamate receptors as chemosensory receptors in Drosophila. Cell

2009c; 136(1): 149-62.

Boto T, Gomez-Diaz C, and Alcorta E.

Epression analysis of the 3 G-protein subunits, Gα, Gβ, Gγ, in the olfactory receptor organs of adult Drosophila melanogaster. Chem Senses 2010;

35, 183-193.

Chatterjee A, Roman G, and Hardin PE.

Go contributes to olfactory reception in Drosophila melanogaster. BMC Physiol 2009; 9(22): 1-7.

Clyne PJ, Warr CG, Freeman MR, Lessing D, Kim J, Carlson JR. A novel family of divergent seven-transmembrane proteins: candidate odorant receptors in Drosophila. Neuron 1999; 22(2):

327-38.

Cooley JW, Tukey JW. An algorithm for the machine calculation of complex Fourier series. Mathematics of computation JSTOR 1965; 19(90):

297-301.

Couto A, Alenius M, Dickson BJ. Molecular, anatomical, and functional

organization of the Drosophila olfactory system. Curr Biol 2005;

15(17): 1535-47.

De Bruyne M, Clyne PJ, Carlson JR. Odor coding in a model olfactory organ: the Drosophila maxillary palp. J Neurosci 1999; 19(11): 4520-32.

De Bruyne M, Foster K, Carlson JR. Odor coding in the Drosophila antenna.

Neuron 2001; 30(2): 537-52.

French AS, Holden AV. Alias-free sampling of neuronal spike trains. Kybernetik 1971; 8(5): 165-71.

Gao Q, Yuan B, Chess A. Convergent projections of Drosophila olfactory

neurons to specific glomeruli in the antennal lobe. Nat Neurosci 2000;

3(8): 780-5.

Geier M, Bosch OJ, Boeckh J. Influence of odour plume structure on upwind flight of mosquitoes towards hosts. J Exp Biol 1999; 202 (Pt 12): 1639-48.

channeling scent. Cell 2008; 133(5):

761-3.

Hallem EA, Ho MG, Carlson JR. The molecular basis of odor coding in the Drosophila antenna. Cell 2004;

117(7): 965-79.

Hallem EA, Carlson JR. Coding of odors by a receptor repertoire. Cell 2006;

125(1): 143-60.

Hildebrand JG. Analysis of chemical signals by nervous systems. Proc Natl Acad Sci U S A. 1995; 92:67-74.

Hildebrand JG, Shepherd GM. Mechanisms of olfactory discrimination: converging evidence for common principles across phyla. Annu Rev Neurosci 1997; 20:

595-631.

Justus KA, Carde RT, French AS. Dynamic properties of antennal responses to pheromone in two moth species. J Neurophysiol 2005; 93(4): 2233-9.

Juusola M, French AS. The efficiency of sensory information coding by mechanoreceptor neurons. Neuron 1997; 18(6): 959-68.

Kain P, Chandrashekaran S, Rodrigues V, and Hasan G. Drosophila mutants

in phospholipid signaling have reduced olfactory responses as adults and larvae. J Neurogenet 2008; 23(3):

303-312.

Kaissling K. Olfactory perireceptor and receptor events in moths: a kinetic model revised. J Comp Physiol A Neuroethol Sens Neural Behav Physiol 2009; 195(10): 895-922.

Keil TA. Contacts of pore tubules and sensory dendrites in antennal chemosensilla of a silkmoth: demonstration of a possible pathway for olfactory molecules. Tissue Cell 1982; 14(3):

451-62.

Keil TA. Fine structure of the pheromone-sensitive sensilla on the antenna of the hawkmoth, Manduca sexta. Tissue Cell 1989; 21(1): 139-51.

Laue M, Maida R, and Redkozubov A. G-protein activation, identification and immunolocalization in

pheromone-sensitive sensilla trichoidea of moths.

Cell Tissue Res. 1997; 288: 149-158.

Laissue PP and Vosshall LB. The Olfactory Sensory Map in Drosophila. Adv Exp

Med Biol 2008; 628: 102-114.

Lundin C, Käll L, Kreher SA, Kappd K, Sonnhammerb EL, Carlson JR, von

Membrane topology of the

Drosophila OR83b odorant re-ceptor. FEBS Lett2007; 581: 5601–4.

Murlis J, Elkinton JS, Cardé RT. Odor plumes and how insects use them.

Annual Review of Entomology 1992;

37: 505-532.

Nakagawa T, Vosshall LB. Controversy and consensus: noncanonical signaling mechanisms in the insect olfactory system. Curr Opin Neurobiol 2009;

19(3): 284-92.

Park S, Shanbhag SR, Dubin AE, de Bruyne M, Wang Q, Yu P, Shimoni N, D'Mello S, Carlson JR, Harris GL, Steinbrecht RA, Pikielny CW.

Inactivation of olfactory sensilla of a

single morphological type

differentially affects the response of Drosophila to odors.

J Neurobiol 2002; 51(3): 248-60.

Randall D, Burggren W, French K. Eckert Animal Physiology. Mechanisms and Adaptations, 4th Ed. WH Freeman and

Company, New York. Cap_tulos, 1997.

Robertson HM, Warr CG, Carlson JR.

Molecular evolution of the insect chemoreceptor gene superfamily in Drosophila melanogaster. Proc Natl Acad Sci U S A 2003; 100 Suppl 2:

14537-42.

Rutzler M, Zwiebel LJ. Molecular biology of insect olfaction: recent progress and conceptual models. J Comp Physiol A Neuroethol Sens Neural Behav Physiol 2005; 191(9): 777-90.

Sanes JR, Hildebrand JG. Structure and development of antennae in a moth, Manduca sexta* 1. Developmental Biology 1976; 51(2): 282-299.

Sato K, Pellegrino M, Nakagawa T, Vosshall LB, Touhara K. Insect olfactory receptors are heteromeric ligand-gated ion channels.

Nature 2008; 452(7190): 1002-6.

Sato K, Touhara K. Insect olfaction:

receptors, signal transduction, and behavior. Results Probl Cell Differ 2009; 47: 121-38.

Schmidt CJ, Garen-Fazio, S, Chow YK, and Neer EJ. Neuronal expression of a newly identified Drosophila melanogaster G protein alpha o

subunit. Cell Regul 1998; 1(1): 134.

Serizawa S, Miyamichi K, Sakano H.

Negative feedback regulation ensures the one neuron-one receptor rule in the mouse olfactory system. Chem Senses 2003; 30 Suppl 1: i99-100.

Shanbhag SR, Muller B, Steinbrecht RA.

Atlas of olfactory organs of Drosophila melanogaster:: 1. Types, external organization, innervation and

distribution of olfactory sensilla.

International Journal of Insect Morphology and Embryology 1999;

28: 377-397.

Shanbhag SR, Muller B, Steinbrecht RA.

Atlas of olfactory organs of Drosophila melanogaster 2. Internal organization and cellular architecture of olfactory sensilla. Arthropod Struct Dev 2000;

29(3): 211-29.

Shields VDC, Hildebrand JG. Recent advances in insect olfaction specifically regarding the morphology and sensory physiology of antennal sensilla of the female sphinx moth Manduca sexta.

Microscopy research and technique 2001; 55(5): 307-329.

Smart R, Smart R, Kiely A, Vargas MBE, Carraher C, Kralicek AV, Christie DL, Chen C, Newcomb RD, Warr CG. Drosophila odorant receptors

are novel seven transmembrane domain proteins that can signal independently of heterotrimeric G proteins. Insect Biochem Mol Biol.

2008; 38: 770–780.

Steinbrecht RA. Pore structures in insect olfactory sensilla: a review of data and concepts. International Journal of Insect Morphology and Embryology 1997; 26(3-4): 229-245.

Stengl M, Hatt H, Breer H. Peripheral processes in insect olfaction. Annu Rev Physiol 1992; 54: 665-81.

Stocker RF. The organization of the chemosensory system in Drosophila melanogaster: a review. Cell Tissue

Res 1994; 275(1): 3-26.

Stocker RF. Drosophila as a focus in olfactory research: mapping of olfactory sensilla by fine structure, odor specificity, odorant receptor expression, and central connectivity. Microsc Res Tech 2001; 55(5): 284-96.

Talluri S, Bhatt A, and Smith DP.

Identification of a Drosophila G protein alpha subunit (dGq alpha-3) expressed

in chemosensory cells and central neurons. PNAS USA 1995; 92: 11479.

Tanoue S, Krishnan P, Chatterjee A, and Hardin PE. G protein-coupled

receptor kinase 2 is required for rhythmic olfactory responses in Drosophila. Curr. Biol. 2008; 18, 787–

794.

Tichy H, Hinterwirth A, Gingl E. Olfactory receptors on the cockroach antenna signal odour ON and odour OFF by excitation. Eur J Neurosci 2005;

22(12): 3147-60.

Touhara K, Vosshall LB. Sensing odorants and pheromones with chemosensory receptors. Annu Rev Physiol 2009; 71:

307-32.

Tripathy S, Peters OJ, Staudacher EM, Kalwar FR, Hatfield MN, Daly KC.

Odors pulsed at wing beat frequencies are tracked by primary olfactory networks and enhance odor detection.

Front Cell Neurosci 2010; 4: ahead of print.

Van der Goes van Naters W, Carlson JR.

Receptors and neurons for fly odors in Drosophila. Curr Biol 2007; 17(7):

606-12.

Venkatesh S, Singh N, et al. Sensilla on the third antennal segment of Drosophila melanogaster Meigen (Diptera:

Drosophilidae). International Journal of Insect Morphology and Embryology 1984;13(1): 51-63.

Vetter RS, Sage AE, Justus KA, Carde RT, Galizia CG. Temporal integrity of an airborne odor stimulus is

greatly affected by physical

aspects of the odor delivery system. Chem Senses 2006; 31(4):

359-69.

Vickers NJ, Christensen TA, Baker TC, Hildebrand JG. Odour-plume

dynamics influence the brain's olfactory code. Nature 2001;

410(6827): 466-70.

Vickers NJ. Winging it: moth flight behavior and responses of olfactory neurons are shaped by pheromone plume dynamics. Chem Senses 2006; 31(2):

155-66.

Vosshall LB, Amrein H, Morozov PS, Rzhetsky A, Axel R. A spatial map of olfactory receptor expression in the

725-36.

Vosshall LB, Wong AM, Axel R. An olfactory sensory map in the fly brain.

Cell 2000; 102(2): 147-59.

Vosshall LB, Stocker RF. Molecular architecture of smell and taste in

Drosophila. 2007.

Weissburg MJ. The fluid dynamical context of chemosensory behavior. Biol Bull 2000; 198(2): 188-202.

Wicher D, Schafer R, Bauernfeind R, Stensmyr MC, Heller R, Heinemann SH, Hansson BS.

Drosophila odorant receptors are both ligand-gated and cyclic-nucleotide-activated cation channels. Nature 2008; 452(7190): 1007-11.

Wicher D, Schafer R, Bauernfeind R, Stensmyr MC, Heller R, Heinemann SH, Hansson BS.

Or83b--receptor or ion channel? Ann NY Acad Sci 2009; 1170: 164-7.

Xu P, Atkinson R, Jones DNM, Smith DP.

Drosophila OBP LUSH is required for activity of pheromone-sensitive neurons. Neuron 2005; 45(2): 200.

Yao CA, Ignell R, Carlson JR.

Chemosensory coding by neurons in the coeloconic sensilla of the Drosophila antenna. 2005.

Zufall F and Leinders-Zufall T. The Cellular and Molecular Basis of Odor Adaptation. Chem Senses 2000;

25: 473-481.