A network model for learning-induced changes in odor representation in the antennal lobe

Similar documents
A model for calculation of growth and feed intake in broiler chickens on the basis of feed composition and genetic features of broilers

Multi-template approaches for segmenting the hippocampus: the case of the SACHA software

Volume measurement by using super-resolution MRI: application to prostate volumetry

Evaluation of noise barriers for soundscape perception through laboratory experiments

Generating Artificial EEG Signals To Reduce BCI Calibration Time

VS : Systemische Physiologie - Animalische Physiologie für Bioinformatiker. Neuronenmodelle III. Modelle synaptischer Kurz- und Langzeitplastizität

Reporting physical parameters in soundscape studies

On the empirical status of the matching law : Comment on McDowell (2013)

Mathieu Hatt, Dimitris Visvikis. To cite this version: HAL Id: inserm

Virtual imaging for teaching cardiac embryology.

From universal postoperative pain recommendations to procedure-specific pain management

The association of and -related gastroduodenal diseases

Optimal electrode diameter in relation to volume of the cochlea

The forming of opinions on the quality of care in local discussion networks

Estimation of Radius of Curvature of Lumbar Spine Using Bending Sensor for Low Back Pain Prevention

Characteristics of Constrained Handwritten Signatures: An Experimental Investigation

Iodide mumps: Sonographic appearance

Bilateral anterior uveitis secondary to erlotinib

Efficacy of Vaccination against HPV infections to prevent cervical cancer in France

Dietary acrylamide exposure among Finnish adults and children: The potential effect of reduction measures

Prevalence and Management of Non-albicans Vaginal Candidiasis

SIDE-SPECIFIC OLFACTORY CONDITIONING LEADS TO MORE SPECIFIC ODOR REPRESENTATION BETWEEN SIDES BUT NOT WITHIN SIDES IN THE HONEYBEE ANTENNAL LOBES

Moderate alcohol consumption and risk of developing dementia in the elderly: the contribution of prospective studies.

Usefulness of Bayesian modeling in risk analysis and prevention of Home Leisure and Sport Injuries (HLIs)

A Study on the Effect of Inspection Time on Defect Detection in Visual Inspection

Improving HIV management in Sub-Saharan Africa: how much palliative care is needed?

Extensions of Farlie-Gumbel-Morgenstern distributions: A review

Relationship of Terror Feelings and Physiological Response During Watching Horror Movie

Pharmacokinetics of caspofungin in a critically ill patient with liver cirrhosis

An Alternate, Egg-Free Radiolabeled Meal Formulation for Gastric-Emptying Scintigraphy

On applying the matching law to between-subject data

Stereotypical activation of hippocampal ensembles during seizures

SUPPLEMENTARY INFORMATION

Daily alternating deferasirox and deferiprone therapy for hard-to-chelate β-thalassemia major patients

Automatic spread of attentional response modulation along Gestalt criteria in primary visual cortex

A Guide to Algorithm Design: Paradigms, Methods, and Complexity Analysis

Enrichment culture of CSF is of limited value in the diagnosis of neonatal meningitis

The storage and recall of memories in the hippocampo-cortical system. Supplementary material. Edmund T Rolls

In vitro study of the effects of cadmium on the activation of the estrogen response element using the YES screen

STDP enhances synchrony in feedforward network

et al.. Rare myopathy associated to MGUS, causing heart failure and responding to chemotherapy.

Defining culture and interculturality in the workplace

anatomic relationship between the internal jugular vein and the carotid artery in children after laryngeal mask insertion. An ultrasonographic study.

Effets du monoxyde d azote inhalé sur le cerveau en développement chez le raton

Neural Correlates of Odor Learning in the Presynaptic Microglomerular Circuitry in the Honeybee Mushroom Body Calyx

RECIPROCITY CALIBRATION OF A SONAR TRANSDUCER FROM ELECTRICAL IMPEDANCE MEASUREMENTS IN WATER AND IN AIR : THE DELTA-Z RECIPROCITY CALIBRATION METHOD

Sparsening and Temporal Sharpening of Olfactory Representations in the Honeybee Mushroom Bodies

Interglomerular communication in olfactory processing

Realization of Visual Representation Task on a Humanoid Robot

Adaptive RR Prediction for Cardiac MRI

A new approach to muscle fatigue evaluation for Push/Pull task

ANALYSIS AND IMPROVEMENT OF A PAIRED COMPARISON METHOD IN THE APPLICATION OF 3DTV SUBJECTIVE EXPERIMENT. IEEE

Estimated intake of intense sweeteners from non-alcoholic beverages in Denmark 2005

Unusual presentation of neuralgic amyotrophy with impairment of cranial nerve XII

Synaptic Learning Rules and Sparse Coding in a Model Sensory System

Input-speci"c adaptation in complex cells through synaptic depression

Adaptive classification of mental states for asynchronous brain computer interfaces

Comments on the article by Tabache F. et al. Acute polyarthritis after influenza A H1N1 immunization,

Neural plasticity of mushroom body-extrinsic neurons in the honeybee brain

Reading Neuronal Synchrony with Depressing Synapses

Shadowing and Blocking as Learning Interference Models

HOW COST-EFFECTIVE IS NO SMOKING DAY?

Theme 2: Cellular mechanisms in the Cochlear Nucleus

In Search of the Engram in the Honeybee Brain

Modulators of Spike Timing-Dependent Plasticity

Gender differences in condom use prediction with Theory of Reasoned Action and Planned Behaviour: the role of self-efficacy and control

Ch.20 Dynamic Cue Combination in Distributional Population Code Networks. Ka Yeon Kim Biopsychology

Information Processing During Transient Responses in the Crayfish Visual System

Double dissociation of value computations in orbitofrontal and anterior cingulate neurons

Cellular Neurobiology BIPN140

Modeling of Hippocampal Behavior

LYMPHOGRANULOMA VENEREUM PRESENTING AS PERIANAL ULCERATION: AN EMERGING CLINICAL PRESENTATION?

Binaural hearing and future hearing-aids technology

AIDS IMPACT SPECIAL ISSUE The face of HIV and AIDS: can we erase the stigma?

Sensory Memory for Odors is Encoded in Spontaneous Correlated Activity between Olfactory Glomeruli

A model of the interaction between mood and memory

EVEROLIMUS IN RELAPSED HODGKIN LYMPHOMA, SOMETHING EXCITING OR A CASE OF CAVEAT mtor?

Role of Inhibition for Temporal and Spatial Odor Representation in Olfactory Output Neurons: A Calcium Imaging Study

Solving the distal reward problem with rare correlations

Artificial Emotion Generation Based on Personality, Mood, and Emotion for Life-Like Facial Expressions of Robots

Processing Stages of Visual Stimuli and Event-Related Potentials

Time-related decay or interference-based forgetting in working memory?

Usability Evaluation for Continuous Error of Fingerprint Identification

Spiking neural network simulator: User s Guide

FROM ANTENNA TO ANTENNA: LATERAL SHIFT OF OLFACTORY MEMORY IN BEES

Long-term synaptic plasticity. N500, 6 Sept., 2016

Synaptic Plasticity and Connectivity Requirements to Produce Stimulus-Pair Specific Responses in Recurrent Networks of Spiking Neurons

Rolls,E.T. (2016) Cerebral Cortex: Principles of Operation. Oxford University Press.

Sensory Memory for Odors Is Encoded in Spontaneous Correlated Activity Between Olfactory Glomeruli

Chorea as the presenting manifestation of primary Sjögren s syndrome in a child

Spatial and temporal coding in an olfaction-inspired network model

Configural Olfactory Learning in Honeybees: Negative and Positive Patterning Discrimination

Proximal tibial bony and meniscal slopes are higher in ACL injured subjects than controls: a comparative MRI study

Introduction to Computational Neuroscience

Heritability of surface area and cortical thickness: a comparison between the Human Connectome Project and the UK Biobank dataset

Sensory representation and learning-related plasticity in mushroom body extrinsic feedback neurons of the protocerebral tract

Cardiac arrhythmia induced by hypothermia in a cardiac model in vitro

The influence of clutter on real-world scene search: Evidence from search efficiency and eye movements

Spike Sorting and Behavioral analysis software

Influence of Train Colour on Loudness Judgments

Transcription:

A network model for learning-induced changes in odor representation in the antennal lobe Michael Schmuker, Marcel Weidert, Randolf Menzel To cite this version: Michael Schmuker, Marcel Weidert, Randolf Menzel. A network model for learning-induced changes in odor representation in the antennal lobe. Deuxième conférence française de Neurosciences Computationnelles, Neurocomp08, Oct 2008, Marseille, France. <hal-00331593> HAL Id: hal-00331593 https://hal.archives-ouvertes.fr/hal-00331593 Submitted on 17 Oct 2008 HAL is a multi-disciplinary open access archive for the deposit and dissemination of scientific research documents, whether they are published or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers. L archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

A network model for learning-induced changes in odor representation in the antennal lobe Michael Schmuker, Marcel Weidert and Randolf Menzel Institute for Biology Neurobiology Freie Universität Berlin Königin-Luise-Str. 28-30, 14195 Berlin Germany m.schmuker@fu-berlin.de, marcel.weidert@parexel.com, menzel@neurobiologie.fu-berlin.de ABSTRACT The antennal lobe (AL) is the insect homologue of the olfactory bulb in mammals. As such, it is the first processing station in the insect olfactory system. It has been shown previously that odorant representations change during associative odor learning [1], but contradictory findings have also been published [2]. We recorded Ca2+-activity of uniglomerular projection neurons (PNs) in the AL of the honeybee Apis mellifera during differential olfactory conditioning. Our results indicate that the activity pattern of PNs in response to odorants can change for the conditioned odor, for the unconditioned odor and for control odors which were not presented during conditioning. We designed a computational model of the glomerular network that can explain the apparent contradiction between the findings we present here and the results reported in [2]. KEY WORDS Olfaction, learning, neural coding, simulation. 1. Introduction The honeybee exhibits remarkable performance in associative odor learning. Like Pavlov's dog learned to salivate in response to a bell ringing before food delivery, the honeybee can be trained to extend its proboscis when an odor is paired with a sucrose reward. Given the accessibility and comparably low complexity of its brain, the honeybee is therefore an ideal model for studying the neural basis of olfactory coding as well as learning and memory formation. It has been shown in the honeybee that both the AL and the mushroom body (MB), the first and the second stages in the olfactory system, are involved in memory formation after olfactory reward conditioning [3]. Memory acquisition can be affected by local injections of octopamine (OA) into either one of the AL or MB [4]. OA is a neuromodulator which putatively gets released by the VUMmx1 neuron in response to sucrose reward [5]. Hence, associative learning is expected to have an effect on the plasticity of neuronal responses in the MB, but also in the AL. Indeed, Faber and coworkers have shown that Ca2+-activity in response to an odorant increases for the learned odors, and that the correlation between response patterns for rewarded and unrewarded odors is reduced [1]. However, when Peele et al. analyzed odor responses of PNs (which form the output of the AL and project to the MB), they did not observe any significant changes in odor responses before, during and after conditioning [2]. Apart from the incongruity in experimental findings, it is also unclear how downstream neuron populations can be able to reliably identify odor stimuli when their representation in the AL changes. In this contribution, we present data which supports the hypothesis that the glomerular activity pattern is changing. In addition, we provide a computational network model that can explain the apparent contradiction in experimental results. The model also resolves the problem how activity patterns can change in response to conditioning, while at the same time enabling downstream populations to reliably identify odor stimuli. 2. Results 2.1 Differential olfactory conditioning We recorded Ca2+-activity from uniglomerular PNs in the AL of harnessed honeybees during differential olfactory conditioning. We used four odors during our conditioning experiments: octanol, linalool, limonene and hexanal. In the pre-phase of each experiment, we presented each odor three times. For conditioning, we paired the presentation of one odor (the conditioned odor, CS+) with sucrose reward. Another odor (the unconditioned odor, CS-) was also presented in the conditioning phase, but was not paired with sucrose solution. CS+ and CS- were presented five times during the conditioning phase. In the post-phase, we presented the CS+, the CS- and the two remaining odors (Ctrl1 and Ctrl2) several times until the signal degraded. 2.2 Response patterns before, during and after differential conditioning PN response was assessed as the integral of Ca2+activity during odor presentation. Depending on the preparation, 4 to 11 glomeruli were analyzed in an animal. Hence, the activity pattern in response to an odor was an n-dimensional vector, with n ranging between 4 and 11. In order to visualize changes in the response patterns, we performed principal component analysis (PCA) on the n-dimensional response patterns. Fig. 1 shows the

scores of PN response patterns on the first two principal components (PCs) in one animal before, during and after conditioning. The response pattern of the CS+ clearly shifts along PC 1. A similar shift can be observed for the CS-, although in a slightly different direction. The control odors (Ctrl1 and Ctrl2) also exhibit some shift, but to a lesser degree and in different directions. We observed similar shifts for the majority of 32 animals that we recorded from. These observations indicate that response patterns of PNs are not invariant, but can be affected by olfactory conditioning. PN1 will receive less inhibition and its activity can increase. Fig. 2: Model of glomerular architecture, schematic. Fig. 1: Principal component (PC) scores for the first two PCs of PN response patterns before, during, and after conditioning. Symbols indicate the role of the odor (+: CS+, : CS-, : Ctrl1, : Ctrl2), shades of gray show the phase of the experiment ( : pre, : training, : post). 2.3 A model of the glomerular network Based on anatomical findings from the honeybee AL, we designed a model for glomerular network architecture that can resolve the apparent contradiction of our data with those from [2]. Since there are several PNs innervating each glomerulus [6], we formed the hypothesis that not all PNs inside a glomerulus respond with the same strength to upstream activity, but their responses can differ when a reward is delivered. In consequence, measurements of Ca2+-activity in a glomerulus could yield different results depending on the PN subpopulation that has been stained, probably because of a slight variation in the injection site or amount of injected dye. Sparser stainings may reveal differential effects that cannot be observed when staining all PNs inside a glomerulus, because the differences could average out. Figure 2 shows a schematic of our model. An olfactory receptor neuron (ORN) drives two PNs, and two inhibitory local neurons (iln and iln-vum). Both PNs also drive the iln and receive recurrent inhibition from it. The PNs project their output to the MB. The VUMmx1 neuron (VUM) is active upon reward delivery and modulates the activity of a subpopulation of ilns (iln-vum), which project to a subpopulation of PNs in the glomerulus. This setting allows for the VUMmx1 neuron to modulate in an associative way the activity of PNs via the iln-vum. At the same time, the overall activity can be kept constant through action of the iln if PN2 decreases its activity, iln activity decreases as well. In order to demonstrate the function of the proposed network, we implemented it using PyNN (http://neuralensemble.org/trac/pynn). We used the IF_curr_alpha spiking neuron model (integrateand-fire neurons with alpha-shaped postsynaptic currents); ORN firing was modeled as a poisson process, using PyNN's SpikeSourcePoisson model. Figure 3 shows spike trains from an exemplary simulation run. As proof of concept, we focused on the behavior of the network in the case that asymmetric PN activity is caused by activity of the VUMmx1 neuron when a reward is delivered, and did not yet consider plasticity in the network. Hence, the simulation only represented the action of the network during the conditioning phase. Fig. 3: Spike trains from one simulation run. Two conditions are shown, with and without reward. We modeled the reward dependent modulatory activity of the VUM neuron as a decrease in firing threshold of the iln-vum unit. In the above example, the threshold was at -35 mv in the no reward condition, leading to 11 spikes being fired. When a reward was given, the threshold decreased to -55 mv, resulting in 54 spikes. In both conditions, the ORN delivered 46 spikes per odorant presentation. In the no-reward condition, PN1 fired 29 spikes while PN2 fired 26 spikes. But when the VUM-neuron was active in response to a reward, PN2 received more inhibition via the iln-vum and decreased its response to 21 spikes, while PN1 received less inhibition via iln and increased its response to 33 spikes. Hence, when averaging the response of PN1 and PN2, the response magnitude when reward is given (27.5 spikes) would be hardly distinguishable from the no-

reward condition (27 spikes). Only when analyzing PN1 and PN2 separately, the effect of reward presentation can be observed as an increase in asymmetry between PN responses. The above scheme has also advantages when considering downstream neuron populations which should reliably identify the presented odor. If the signals of all PNs in one glomerulus are averaged, the response pattern is invariant to reward presentation. Simultaneously, the differential signal from PNs is available to indicate if the odor has been rewarded. 2.4 Adding plasticity to the model So far, we only considered the effect of neuromodulation during the conditioning phase. As we have shown in section 2.2, appetitive conditioning causes changes in odor coding which outlast the conditioning phase. Hence there must be a mechanism which induces longer-lasting modifications of some synaptic weights. For reward-based learning, both the input signal and the reward signal are required. In our model, both signals are available at the VUM-iLN: the ORN provides the input signal, while the reward signal is provided by modulatory input from the VUM neuron. If the synaptic weight between ORN and VUM-iLN is subject to plasticity, the VUM-iLN's odor response, and therefore its impact on odor coding by PN subpopulations, can be modified according to the temporal overlap between ORN input and reward signal. Hence, the synapse between ORN and iln-vum is the ideal location to induce synaptic plasticity during the conditioning phase. Proper function of the network depends on the combination of synaptic weights. Not all weight combinations are equally well suited. If we introduce synaptic plasticity at a specific synapse, one weight is changing in an activity dependent manner. The remaining weights must be chosen such that a weight change at that single synapse is tolerated and does not jeopardize the function of the entire network. We addressed this issue by scanning a large number of parameter combinations by their usefulness for the performance, or fitness, of the model. In a useful model, the average spike count of both PNs should stay as constant as possible in rewarded and unrewarded trials. We assessed this criterion by equation (1): f mean= PN1r PN2 r 1, PN1nr PN2nr (1) with PN1r and PN2r the PN's spike counts for rewarded trials, PN1nr and PN2nr the spike counts for unrewarded trials, and fmean the fitness criterion indicating how well the mean spike count from both PNs is preserved during rewarded and unrewarded trials. In addition, the difference in spike counts should be small in unrewarded trials and large in rewarded trials. This was quantified by equation (2): f diff = 1 PN1nr PN2 nr, 1 PN1r PN2r (2) where fdiff is the fitness criterion assessing how much PN responses diverge during rewarded trials. Fitness increases if both criteria approach zero. The overall fitness f of a particular combination was calculated as f= f mean f diff 2. (3) Table 1 displays the weights we used for probing the multidimensional weight space. Negative weights indicate fitness assessment. We will use the abbreviations given in the third column for reference in the remainder. Table 1: Weights used for fitness assessment. source target abbr. weights ORN PN1, PN2 OP 20 25 30 35 40 ORN iln Oi 2 4 6 8 10 PN1, PN2 iln Pi 6 8 10 12 14 iln PN1, PN2 ip iln-vum PN2 ORN ivp iln-vum OiV -20-17.5-15 -12.5-10 -10-8.5 5 7.5-7 -5.5-4 10 12.5 15 Our goal was to identify, out of all possible 15625 weight combinations, those in which OiV could vary the most with the least impact on model fitness. To this end, we assessed model fitness with each combination of weights from Table 1 for all values of OiV and calculated the mean and the variance of f. The ideal weight combination should exhibit small mean fitness, and low variance for different values of OiV. Figure 4 shows the distribution of mean fitness and its variance. Fig 4: Mean vs. variance of fitness over different values of OiV. The gray circle marks the weight combination that we chose for the remaining simulations. Most weight combinations yield good mean fitness around 0.15, with low variance around 0.02. For the remaining simulations, we chose the combination closest to the origin. The corresponding weight values were OP: 35, Oi: 6, Pi: 12, ip: -15, ivp: -4. We enabled spike-timing dependent plasticity (STDP, [7]) at the synapse between ORN and iln-vum (OiV). Weights were modified in an additive manner using PyNN's AdditiveWeightDependence STDP rule with minimum and maximum weights wmin = 5 and wmax = 15. The maximum amount of weight change was bounded by A+ = 0.005 for increasing and A- = 0.0025 for decreasing weights. The temporal window for STDP was defined by τplus = 5 ms and τminus = 7 ms for potentiation and depression, respectively.

In order to mimic the data shown in 2.2, we organized our simulation runs into three phases: pre-, training- and post-phase, with 3, 5 and 3 trials respectively. The only difference between the phases is that during the training phase we lower the threshold of the iln-vum neuron to mimic a rewarded odor, in accordance to the model description in 2.3. Each trial was 500ms long, and trials were separated by intervals of 100 ms length during which no input activity was delivered. Figure 5 depicts spike trains from one simulation run. Fig. 5: Spike trains from one simulation run with plasticity. The gray rectangle marks the training phase, where the threshold of the iln-vum neuron is reduced. In the pre-phase, the iln-vum neuron fired only few spikes, while in the training-phase its firing rate increased due to the decreased firing threshold, mimicking neuromodulatory input by the VUMmx1 neuron. In the post phase, the iln-vum exhibited an increased response to the stimulus as a result of the weight increase that has been brought about by STDP. In consequence of the increased iln-vum response, PN2 decreased its firing rate during the training- and post-phase of the experiment, while the response of PN1 increased. This behavior can be observed more specifically in Figure 6, which shows the relative spike counts of PN1 and PN2 during the experiment. from PN1 and PN2. During the training-phase, this difference increased and it was still large in the postphase, demonstrating that synaptic plasticity at the OiVsynapse can effectively influence PN responses, making them inhomogeneous for rewarded odors. It must be noted that in the post-phase, the difference was somewhat smaller than during training, in slight incongruence with the original data. However, this incongruity could be addressed by using a different set of weights, which yield smaller, i.e. better values for fdiff, and thus a larger difference in the post-phase. In order to demonstrate that average PN output actually stays the same, Figure 6 also depicts how the mean spike count evolved throughout the experiment. We normalized the mean values such that the value from the first trial equals one. Since the ORN input spike train is generated by a poisson process, it is subject to stochastic fluctuations, which are reflected by the variance in mean spike count over trials. In total, mean spike count was on the same level in the pre- and postphase, with an apparent increase during the training phase. In the above settings, the ORN always fired with with a spike rate of approximately 100 Hz. Because glomerular activity patterns are not all-or-nothing signals, but rather use the entire range of activities, we tested how the model performs for a range of input spike rates. Figure 7 depicts how the synaptic weight between ORN and iln-vum evolved for ORN rates between 10 and 130 Hz. Fig. 7: Evolution of the weight between ORN and ilnvum for input rates between 10 and 130 Hz. Fig. 6: Spike counts of PN1 and PN2, relative to the mean of both in each trial. Light gray crosses depict how the mean spike count varies, relative to the mean in trial 1. The gray rectangle marks the training phase. In each trial, spike counts are scaled such that the mean spike count of both PNs equals one. In the pre-phase, there was only a small difference between spike counts The higher the input rate, the more drastic was the change in synaptic weight. Hence, in multi-glomerular activity patterns, the glomerulus exhibiting the strongest response will change its response the most. Another observation that can be made in Figure 7 is that the weight change is strongest during the training phase, and that there is relatively small or no change in the post-phase. This behavior indicates that the parameter setting we chose for the STDP rule is well suited to elicit synaptic potentiation mostly during training trials, and prevents that the weight changes erroneously during non-rewarded trials. However, it must be noted that for the largest input rate, the synaptic weight is maxed out, so no further potentiation can take place. Possibly, this behavior could be avoided by tuning the parameters

of the STDP mechanism, or simply avoiding the use of input rates higher than 100Hz in the model. 3. Conclusion We presented data showing changes in the representation of odorants in the AL induced by olfactory associative learning. These data conflict with previously published findings, and also raise the question how downstream neurons can identify odor stimuli when their representation in the AL is changing. Proposing a connectivity scheme for glomeruli that can resolve that conflict, we constructed a computational model of the glomerular network, in which the response of PNs in one glomerulus is not homogeneous, but diverges when an odor is rewarded or associated with a reward. In a computational conditioning experiment, we employed STDP at one synapse in the network to acquire this assymetry in PN response during the training-phase when odor presentation is paired with activity in the VUMmx1 neuron (i.e., when the odor is rewarded), and showed that it was preserved in later trials without VUM activity. We showed that the network model is capable of generating asymmetric PN responses, while at the same time keeping the average response relatively constant. Hence, downstream neurons could obtain reliable information about the stimulus by integrating the signal of all PNs from one glomerulus, or obtain information about which odor was rewarded by analyzing the differential signal. The model makes the prediction that responses of PNs in the same glomerulus are not homogeneous when odors are rewarded. This prediction can be tested experimentally, e.g. by recording the activity of PNs belonging to the same glomerulus during olfactory conditioning. The experimental data provides us not only with data from one glomerulus, but from several glomeruli giving the entire activity pattern in the frontal aspect of the antennal lobe in response to odorants. Our goal is to use several model glomeruli in order to reproduce our experimental findings and analyze the effect of plasticity in the antennal lobe on odorant response patterns. References [1] T. Faber, J. Joerges, & R. Menzel, Associative learning modifies neural representations of odors in the insect brain, Nature Neuroscience, 2(1), 1999, 74-78. [2] P. Peele, M. Ditzen, R. Menzel & C.G. Galizia, Appetitive odor learning does not change olfactory coding in a subpopulation of honeybee antennal lobe neurons, Journal of Comparative Physiology A, 192(10), 2006, 1083-1103. [3] R. Menzel & M. Giurfa, Cognitive architecture of a mini-brain: the honeybee, Trends in Cognitive Sciences, 5(2), 2001, 62-71. [4] M. Hammer & R. Menzel, Multiple sites of associative odor learning as revealed by local brain microinjections of octopamine in honeybees, Learning & Memory, 5(1-2), 1998, 146-156. [5] M. Hammer, An identified neuron mediates the unconditioned stimulus in associative olfactory learning in honeybees, Nature, 366, 1993, 59-63. [6] J. Rybak, Die strukturelle Organization der Pilzkörper und synaptische Konnektivität protocerebraler Interneuronen im Gehirn der Honigbiene, Apis mellifera (Ph.D thesis, Freie Universität Berlin, 1994). [7]G. Bi & M. Poo, Synaptic modification by correlated activity: Hebb's postulate revisited, Annual Review of Neuroscience, 24, 2001, 139-166.