INTERNATIONAL GRADUATE SCHOOL OF NEUROSCIENCES (IGSN)

RUHR UNIVERSITÄT BOCHUM

THE ASSOCIATED PROTEOME

Doctoral Dissertation

David Jonathan Barbour

Department of Cell Physiology

Thesis advisor: Prof. Dr. Dr. Dr. Hanns Hatt

Bochum, Germany (30.12.05)

ABSTRACT

Olfactory receptors (OR) are G--coupled membrane receptors (GPCRs) that comprise the largest vertebrate multigene family (~1,000 ORs in mouse and rat, ~350 in human); they are expressed individually in the sensory of the nose and have also been identified in human testis and sperm. In order to gain further insight into the underlying molecular mechanisms of OR regulation, a bifurcate proteomic strategy was employed.

Firstly, the question of stimulus induced plasticity of the olfactory sensory was addressed. Juvenile mice were exposed to either a pulsed or continuous application of an aldehyde odorant, octanal, for 20 days. This was followed by behavioural, electrophysiological and proteomic investigations. Both treated groups displayed peripheral desensitization to octanal as determined by electro-olfactogram recordings. This was not due to anosmia as they were on average faster than the control group in a behavioural food discovery task. To elucidate differentially regulated between the control and treated mice, fluorescent Difference Gel Electrophoresis (DIGE) was used. Seven significantly up-regulated and ten significantly down-regulated gel spots were identified in the continuously treated mice; four and twenty-four significantly up- and down-regulated spots were identified for the pulsed mice, respectively. The spots were excised and proteins were identified using mass spectrometry. Several promising candidate proteins were identified including potential transcription factors, cytoskeletal proteins as well as calcium binding and odorant binding proteins. We propose that the dominant desensitizing factor in the continuously treated mice was down-regulation of odorant binding proteins. In the pulsed group no principle factor was evident, however, in terms of the number of proteins and degree of post-translational modifications, the pulsed group displayed greater plasticity.

Secondly, in order to reveal which olfactory receptors are expressed in human spermatozoa, and alternative proteomic strategy was developed. The optimised method employs an ‘affinity two phase partition’ system in conjunction with multi- enzymatic digestion in the presence of an organic solvent. The resultant peptides were identified using Multidimensional Protein Identification Technology (MudPIT). 222 integral membrane proteins were identified including 57 GPCRs, 35 of which were chemoreceptors, consisting of 32 ORs and 3 taste receptors. Notably, most of the peptides which resulted in GPCR identification were cleaved from transmembrane domains, thus demonstrating the efficacy of this strategy in membrane proteomics. In addition to the chemoreceptors, 23 neuronal proteins were also detected suggesting that both cell types may have more in common than usually perceived.

Both proteomic strategies afford a powerful means whereby novel protein candidates can be elucidated and thereby provide greater insight into plasticity of the olfactory receptor, its associated proteins, and the role of olfactory receptors in reproduction. TABLE OF CONTENTS

1 Introduction ______1 1.1 The ______1 1.1.1 The ______3 1.1.2 The olfactory receptor protein (OR) ______3 1.2 Project Aims______7 2 Proteomic investigation of the olfactory epithelium ______7 2.1 Olfactory stimulus induced plasticity – a background ______7 2.2 Proteomics: an introduction______11 2.2.1 Technologies for large scale proteomics______12 2.2.1.1 2D-electrophoresis ______12 2.2.1.2 Multi-dimensional protein identification technology MudPIT _____ 14 2.2.2 Neuro-proteomics______16 2.3 Conclusion ______18 2.4 Materials and Methods ______18 2.4.1 Animal and Tissue Preparation ______18 2.4.2 Behavioural study ______19 2.4.3 Microdissection of olfactory epithelium ______20 2.4.4 Electro-olfactogram (EOG) recording______20 2.4.5 Sample preparation for DIGE______21 2.4.6 Protein Labelling - Fluorescent DIGE minimal labelling ______21 2.4.7 IEF and 2-DE______22 2.4.8 Gel scanning, digitising and analysis______22 2.4.9 2D-SDS PAGE Gel total protein staining ______23 2.4.10 Trypsin digestion ______23 2.4.11 Mass Spectrometry ______24 2.4.12 Statistics ______25 2.5 Results ______25 2.5.1 Regions of interest______29 2.5.2 Pulsed Gel Analysis ______33 2.5.3 Continuous ______39 2.5.4 Protein identification ______43 2.5.5 Behavioural study ______49 2.5.6 EOG______51 2.6 Discussion ______54 2.6.1 Mice physiology______54 2.6.2 Regulation of proteins: functional categorisation & comparison______55 2.6.2.1 The cytoskeleton ______57 2.6.2.2 Intermediate early / transcription ______60 2.6.2.3 Calcium binding proteins ______62 2.6.2.4 Chaperones ______68 2.6.2.5 Lipocalins ______70 2.6.2.6 Xenobiotic & anti-oxidant metabolism______76 2.6.2.7 Energy metabolism ______77 2.6.3 Consolidating the findings: Protein to Phenotype ______78

2.6.3.1 Protein regulation in the continuously treated mice ______79 2.6.3.2 Protein regulation in the pulsed treated mice ______79 2.6.4 Conclusion ______80 3 Proteomic investigation of the human sperm membrane______82 3.1 A correlation between chemosenses and reproduction? ______82 3.2 The challenge of membrane proteomics ______83 3.2.1 In-gel methods ______84 3.2.2 In-Solution shotgun approach ______85 3.2.3 Fractionation______86 3.3 Sperm proteomics: a background ______87 3.4 Conclusion ______89 3.5 Materials and Methods ______90 3.5.1 LC/LC-MS/MS and Protein Identification ______90 3.5.2 Sperm preparation______91 3.5.3 Ca2+ imaging ______91 3.5.4 Immunocytochemistry______91 3.5.5 Sample optimisation - main strategies ______92 3.5.5.1 Gel Based Approach ______92 3.5.5.2 In Solution Strategy ______94 3.6 Results ______97 3.6.1 Gel Based Strategy______97 3.6.2 In Solution Strategy ______100 3.6.2.1 Solubilisation using organic acid______100 3.6.2.2 Solubilisation using organic solvent ______101 3.6.2.3 Vectorial labelling ______102 3.6.2.4 Lipase & Affinity Enrichment ______102 3.6.3 Validation of data ______107 3.6.3.1 Functional validation - calcium imaging ______108 3.6.3.2 Immunocytochemistry______109 3.7 Discussion ______110 3.7.1 Strategy development ______110 3.7.1.1 Gel Based Strategy______110 3.7.1.2 ‘In-solution’ based strategies ______113 3.7.2 The Solution: A final strategy______117 3.8 Identified membrane proteins ______118 3.8.1 A technical discussion on the identified proteins ______119 3.8.2 The identified proteins from a biological perspective ______120 4 Final Conclusion ______124 4.1.1 Olfactory receptor plasticity______124 4.1.2 Spermatozoa membrane proteome ______125 4.1.3 The olfactory receptor proteome ______125

ii

LIST OF FIGURES

Figure 1.1 Diagramatic representation of the rodent olfactory system Error! Bookmark not defined. Figure 1.2 Diagram depicting initial olfactory receptor ______5 Figure 2.1 Schematic representation of potential stimulus induced plasticity ______11 Figure 2.2 DIGE fluorescent CyDye minimal labelling______13 Figure 2.3 Schematic of 2D-Electrophoresis workflow ______14 Figure 2.4 Schematic of MudPIT Workflow ______15 Figure 2.5 Picture of mice odorant exposure apparatus ______19 Figure 2.6 The structure and properties of Octanal______26 Figure 2.7 Schematic representation of parameters used to calculate spot data______27 Figure 2.8 Example of two DIGE gels Upper______28 Figure 2.9 DIGE scan from 5 gels comparing the control and pulsed treated mice___ 29 Figure 2.10 Biological variation – DIGE gels ______30 Figure 2.11 Biological variation - Histogram from DIGE gels ______31 Figure 2.12 Master gels showing significantly regulated spots ______32 Figure 2.13 ‘Troubleshooting’ protein identification______44 Figure 2.14 Chart showing interaction between Treatment and Substance ______50 Figure 2.15 Frequency of found substance as function of treatment ______51 Figure 2.16 Chart showing EOG amplitudes in response Octanal ______52 Figure 2.17 Functional classification of identified proteins ______55 Figure 2.18 Alignment of identified unknown candidate protein ______72 Figure 2.19 Probability of residue phosphorylation (y-axis______73 Figure 2.20 Comparison of MUP 5 binding pheromones with octanal______75 Figure 2.21 Proposed xenobiotic clearance of octanal ______77 Figure 3.1 Organisational chart showing experimental strategies ______97 Figure 3.2 Sequence coverage of Semenogelin II and Lactotransferrin______98 Figure 3.3 Functional classification of identified 222 Integral Membrane Proteins _103 Figure 3.4 Cleaved peptides that identified GPCRs ______104 Figure 3.5 Scatter plots showing pI,MW,hydrophobicity correlation ______105 Figure 3.6 Virtual 2D gel of identified IMPs ______106 Figure 3.7 Calcium imaging of human sperm ______108 Figure 3.8 Immunocytochemistry validation of a section of neuronal proteins _____ 109 Figure 3.9 The lipid composition of human sperm ______112 Figure 3.10 Coupling of affinity ligand (WGA) to dextran ______117 Figure 3.11 Strategy to identify olfactory receptors in human sperm ______118

iii

LIST OF TABLES

Table 2.1 Most commonly encountered detection methods used in 2D-PAGE ______12 Table 2.2 Statistically significant (P<0.05) changes for pulsed experiment ______33 Table 2.3 Statistically significant (P<0.05) changes for continuous experiment _____ 39 Table 2.4 Protein identification from pulsed experiment ______45 Table 2.5 Proteins identified from continuous experiment ______47 Table 2.6 Fisher least significant difference post hoc results behavioural study _____ 50 Table 2.7 Fisher least significant difference post hoc results EOG______53 Table 2.8 List of candidate proteins ______56 Table 3.1 Olfactory proteins identified using gel based strategy ______100 Table 3.2 Integral membrane proteins identified using organic solvent strategy ____ 101 Table 3.3 Identified chemoreceptors ______107

iv

This work was financially supported by Graduiertenkolleg for “Development and Plasticity of the Nervous System: Molecular, synaptic and cellular mechanisms”. Additional support was given by the International Max Plank Research School in Chemical Biology (IMPRS-CB).

This dissertation has been completed and written independently without external assistance. It has never been submitted in this or a similar form at this or any other domestic or foreign institution of higher learning as a dissertation. The "Guidelines for Good Scientific Practice” (Leitlinien guter wissenschaftlicher Praxis und Grundsätze für das Verfahren bei vermutetem wissenschaftlichen Fehlverhaltens) according to § 9, Sec. 3 of the Promotionsordnung der International Graduate School of Neuroscience der Ruhr-Universität Bochum were adhered to.

v

ACKNOWLEDGEMENTS

First and foremost I would like to thank both of my supervisors Prof Hanns Hatt and Dr Eva Neuhaus, for giving me the opportunity to work on this project. I am grateful to Prof Hatt for his ‘open door policy’ and encouragement throughout. A special thanks goes to Dr Neuhaus for her unwavering guidance and continual support.

This project would not have been possible without the advice and direction of our collaboration partners Dr Dirk Wolters, Jun Prof Bettina Warscheid and Prof Helmut Meyer. I appreciate their enthusiasm and willingness and for giving me the opportunity to receive excellent ‘hands-on’ experience in proteomics.

Due to my collaborative experience in two proteomic laboratories I had the opportunity to work with and meet many people. In particular I would like to thank Dr Andreas Kyas for his endless support; I am also very grateful to Barbara Sitek and Nadine Stoepel for their enthusiasm and stoicism in teaching me 2D-electrophoresis. In fact, I am indebted to all of the staff working at the Medical Proteome Centre 2D-electrophoresis facility for their patience and help, including Dr. Kai Stühler.

A special thanks to Mr Harry Bartel for his expertise in constructing various application systems – his skills and affability are truly an asset to the laboratory. I am indebted to Mr Thomas Lichtleitner for his readiness to help me with computing problems and especially for his advice in desk top publishing packages. I am also grateful to Weiyi Zhang, Anastasia Mashukova and Ruth Dooley, not only for their scientific help, but also for their friendship over the past three years.

Some experiments involved demanding preparations and without the help of several people would have been arduous. In particular I am appreciative to Daniela Brunert and Heike Piechura for their abetment. I am also grateful to Dr Heike Benecke for assisting with the behavioural study.

I would also like to take this opportunity to thank the staff and students of the Cell Physiology department. When I first arrived to Bochum I was immediately impressed by a welcoming and friendly attitude from everyone. I will undoubtedly have fond memories of my time here.

Last and certainly not least I thank my family for their unconditional love, patience and support.

v i

ABBREVIATIONS

1D-SDS PAGE One dimentional sodium dodecylsulfate-polyacrylamide gel electrophoresis 2D-PAGE Two-dimensional polyacrylamide gel electrophoresis ACN Acteonitrile Ambic Ammonium bicarbonate ANOVA Analysis of variance between groups CNG Cyclic-nucleotide-gated channel DIGE Difference gel electrophoresis DTT Dithiothreitol EEG Electroencephalogram EOG Electro-olfactogram ESI Electrospray ionisation GABA Gamma-aminobutyric acid GPCR G-protein coupled receptor LC-MS/MS Liquid Chromatography-Tandem Mass Spectrometry M/T Mitral/tufted cells MALDI Matrix-assisted laser desorption ionisation MudPIT Multi-dimensional protein identification technology OB OE Olfactory epithelium OR Olfactory receptor ORN Olfactory receptor neuron OSN Olfactory sensory neuron pI Isoelectric point RMS Rostral migratory stream SDS Sodium Dodecyl Sulfate SVZ Subventricular zone TRIS Tris(hydroxymethyl) methylamine TRP Transient receptor protein VNO

v ii

1 Introduction The chemosenses are considered to be phylogenetically the oldest sense providing an important means whereby the chemical environment can communicate with an organism. This diversity ranges from the relatively primitive orientation of a unicellular organism to a chemical gradient (chemotaxis/chemokinesis), to complex behaviour demonstrated by vertebrates. Vertebrates have developed sophisticated means to detect and neuronally process a myriad of air-borne (olfaction) and liquid-borne (taste) signals facilitating emotional (anxiety, pleasure), physiological, sociological and sexual behavioural responses. Most notably, the olfactory sense is responsible for detecting aromas and flavours not to mention the detection of chemical cues integral to survival of a species e.g. danger signals (presence of toxin or predator) and pheromones involved in reproduction. Arguably the olfactory system provides an ideal model to investigate neurobiological questions pertaining to development and plasticity because of sustained neurogenesis throughout life. In terms of plasticity, its unique ability to distinguish and respond to a plethora of odours is regulated both at the behavioural and neuronal level. This dynamic processing is dependent on previous experience, the current environment and internal state. Ultimately these changes involve protein regulation, which in terms of odorant induced signal transduction, must begin at the olfactory receptor protein. The intricate mechanisms underlying receptor plasticity such as receptor activation, internalization, trafficking and recycling are poorly understood. Therefore, in order to unravel potential new mechanisms, this investigation has adopted state of the art proteomic techni- ques.

1.1 The Olfactory System The of the olfactory system has been well described for many species which in principle share a basic common primary olfactory pathway (Hildebrand and Shepherd 1997). This common pathway includes the olfactory receptor neurons (ORNs) in the nose (or insect antenna) which project to second order neurons located in the olfactory bulb (or antennal lobe) which ultimately output to the olfactory cortex (or mushroom bodies) (Wilson, Best et al. 2004). In mammals this direct pathway is networked to higher- regions including the orbitofrontal cortex and limbic system (, hippocampus, and perirhinal cortex). Interestingly, recent evidence suggests that conscious perception in mammals may not necessarily involve the thalamus, but instead, a direct pathway between the olfactory cortex and prefrontal cortex (Shepherd 2005) - this raises new questions with regard to central olfactory processing. It has also been proposed that the circuitry of

the olfactory bulb is functionally similar to the circuitry found in the retina and in parts of the neocortex (Mori, Nagao et al. 1999; Dietz and Murthy 2005). It is possible that similar odours are mapped close together in the olfactory bulb and that granule cells create centre-surround fields similar to those in the retina (Mori, Nagao et al. 1999; Aungst, Heyward et al. 2003).

In mammals olfaction is mediated by two anatomically distinct systems-:

i) the accessory olfactory system which receives inputs from the vomeronasal organ (VNO) and encodes for social and sexual behaviour – VNO receptors bind pheromones.

ii) the main olfactory system which receives inputs from the main olfactory epithelium (recognises air-borne odorants) and projects to the olfactory bulb for central processing.

Figure 1.1 Diagramatic representation of the rodent olfactory system Taken from Lledo (2005) et al. The main olfactory system is in green; the accessory in red

2

1.1.1 The olfactory epithelium The olfactory epithelium (OE) is situated dorsocaudal to the respiratory epithelium in the roof of the two nasal cavities of the nose. In adult mice it measures around 170mm2 and comprises of three principle cell types - sensory neurons, sustentacular cells and basal cells. Adjacent sustentacular cells provide metabolic and structural support to ORNs including the detoxification of xenobiotic compounds. Basal cells, on the other hand, are considered to be progenitor (stem) cells which continually regenerate the OE, in fact ORNs are continually regenerated throughout the murine lifespan (Beites, Kawauchi et al. 2005). The epithelium is kept moist by the secretions of olfactory glands which also include odorant binding proteins. These bind to hydrophobic odorants and ‘present’ them to the olfactory receptors.

1.1.2 The olfactory receptor protein (OR) In a landmark paper (Buck and Axel 1991) the olfactory receptor proteins were identified as a superfamily of G-protein coupled receptors (GPCRs). According to the latest genomic analysis it was estimated that between 1,000-1,300 ORs are expressed in mice (Zhang, Rodriguez et al. 2004) with up to 425 in humans (Armbruster and Roth 2005). They have conserved motifs consisting of around 5-10 amino acids and transmembrane (TM) variable regions which may function as ligand- binding sites. Recent molecular modeling of a mouse olfactory receptor (mOR-EG) revealed that the most critical residues involved in ligand binding are hydrophobic amino acids which form a binding pocket in the space formed by TM3, TM5 and TM6 (Katada, Hirokawa et al. 2005). Previous molecular modeling of the rat I7 olfactory receptor also predicted a similar ligand binding site and in particular, identified the residue lysine 164 as having the highest affinity for octanal (Singer 2000). It was proposed that Lysine 164 (TM4) forms an electrostatic interaction with the carbonyl oxygen of octanal.

Most vertebrate investigations have examined rodent olfaction which over time has lead to the establishment of several dogmas-:

1. One receptor one-neuron hypothesis – from an battery of independent investigations it is generally accepted that each olfactory sensory neuron expresses only one olfactory receptor (Mombaerts 2004). Recently this dogma has been challenged by convincing work which showed that in Drosophila, two functional odorant receptors are co-expressed in one

3

neuron (Goldman, Van der Goes van Naters et al. 2005). Nevertheless, in mice this hypothesis is still persuasive.

2. Zonal organization of olfactory receptors - incontestable evidence demonstrated that ORs are organised into four zones (Mori, von Campenhaus et al. 2000) which also converge into specific subsets of glomeruli (Treloar, Feinstein et al. 2002).

3. Adult neurogensis of olfactory sensory neurons – classical experiments have demonstrated that the principle regions of constitutive adult mammalian neurogenesis are the subventricular zone/olfactory bulb and hippocampal dentate gyrus (Emsley, Mitchell et al. 2005). The olfactory epithelium neural stem cell (basal cell) produces two main lineages responsible for regeneration of sustentacular cells and olfactory ensheathing cells (Beites, Kawauchi et al. 2005).

4. In the main olfactory epithelium, OR signal transducion is mediated by Golf/AC/cAMP – the signal transduction mechanism of the ORs is schematically represented below (Figure 1.2). In brief, odorant binding with the olfactory receptor leads to GPCR activation and

dissociation of the Gαolf sub unit from βγ (Menini 1999). Gαolf in turn stimulates a type III adenylate cyclase (AC) which catalyses the conversion of ATP into cAMP. The elevation in cAMP opens a cyclic-nucleotide-gated channel (cAMP increases the channel ‘open probability’), leading to an influx of Na2+ and Ca2+ and a membrane depolarisation. This depolarisation is amplified through the opening of calcium activated chloride channels leading to a Cl- efflux. At the same time the Ca2+ influx decreases cAMP synthesis and the effective affinity for CNG-cAMP, producing channel adaption (Bhandawat, Reisert et al. 2005). Ultimately, this Ca2+ influx in combination with receptor phosphorylation and hydrolysis of cAMP into 5’-AMP, leads to a ‘reset’ of the signal transduction cascade (Breer 2003). It is worth noting that in the vomeronasal organ (VNO), neurons appear to use the PtdIns(4,5)P2/phospholipase C/Ins(1,4,5)P3 pathway for signal transduction (Lucas,

Ukhanov et al. 2003). The classical olfactory transduction elements (Golf, ACIII and CNG-

channels) are replaced with Go- Gi2- subunits and gating is facilitated by a TRP (transient receptor) channel (Breer 2003).

4

Figure 1.2 Diagram depicting initial olfactory receptor signal transduction

5. Olfactory receptors are only functionally expressed in the Olfactory Epithelium – after the identification of the olfactory receptome it was generally accepted that expression of ORs is restricted to the OE. This dogma was later disproved, now evidence exists that ORs are expressed in additional tissues including prostate, skin and spermatozoa. Parmentier suggested that ORs are expressed in mammalian germ cells (Parmentier, Libert et al. 1992) and it was purported that they may play a chemosensory role in sperm development, chemotaxis or in oocyte/sperm interaction. At least one OR protein was shown to be localized to the flagellum of mature dog sperm (Vanderhaeghen, Schurmans et al. 1993). Human sperm have the ability to detect and respond to chemotactic signals secreted by the egg through the odorant receptor hOR 17-4 (Spehr, Gisselmann et al. 2003). Several other components

5

of the odorant signalling cascade of olfactory neurons are present in sperm which include cyclic nucleotide-gated (CNG) channels (Vanderhaeghen, Schurmans et al. 1993; Weyand, Godde et al. 1994; Walensky, Roskams et al. 1995) suggesting analogous transduction mechanisms.

Over the last few years there have been significant advancements in understanding olfactory processes, nevertheless, in spite of this, there is still a salient lack in understanding some fundamental mechanisms. These include-:

• How olfactory receptor expression is controlled at the ‘zonal level’ in the OE.

• The mechanism of olfactory sensory neurons (OSN) convergence to glomeruli.

• The mechanism of olfactory receptor trafficking.

• How odours are encoded and cogitatively perceived.

• Plasticity at a cellular level – activity dependend modulation of OSN thresholds.

In order to identify novel proteins and potential protein-protein interactions which could shed new light on OR regulation, we devised a proteomic strategy. The proteomic investigation of the olfactory receptor per se considered two aspects. Firstly, identifying novel candidate proteins which are either up- or down-regulated in the olfactory epithelium. These changes are in response to odorant activation of the olfactory receptor and therefore reflect plasticity of the olfactory sensory neuron. Secondly, we paralleled this with an investigation into identifying novel candidates in human sperm. One might expect to find a population of olfactory receptors in human sperm which may be modulated by similar mechanisms as in olfactory sensory neurons. Moreover, findings from these parallel investigations could not only help to understand mechanisms in neuroscience, but also may be applied to reproductive medicine.

Both strategies require independent method development and independent experimental design, thus, both approaches are discussed in separate chapters. This dichotomy converges in the ‘final

6

conclusion section’ (chapter 4) which gives an overview of both strategies and highlights principle findings.

1.2 Project Aims The project aim was the analysis of the olfactory receptor associated proteome in order to elucidate novel protein(s) which may play a role in OR regulation in response to odorant stimulation. This goal was realised by addressing two main questions:

1) What changes occur to protein regulation in the olfactory receptor neuron as a consequence of receptor stimulation? Stimulus induced plasticity of the OSN analysing long term affects on protein regulation. Refer to chapter 2.

2) Are olfactory receptors expressed in mature human sperm? Can a parallel be identified between sperm and olfactory chemoreceptors? Characterisation of the human sperm membrane proteome with a particular emphasis on chemoreceptors. Refer to chapter 3.

2 Proteomic investigation of the olfactory epithelium 2.1 Olfactory stimulus induced plasticity – a background Investigations into olfactory plasticity have usually focused on the olfactory bulb and higher cortical regions with less of an emphasis being placed on the periphery. Moreover, the vast majority of these studies were electrophysiological. It is now generally accepted that the characteristic electrical oscillations observed in the olfactory system are gamma (30–80 Hz), beta (15–40 Hz) and theta rhythms (3–12 Hz), with the former two frequencies being associated with odour stimuli (Lagier, Carleton et al. 2004; Lledo, Gheusi et al. 2005). These oscillations are known to take part in encoding of sensory information before their transfer to higher subcortical and cortical areas. For example, there is a clear correlation between odorant induced spatial-temporal formatting in the bulb to a variety of behavioural states, such as memory or perception (Laurent, Stopfer et al. 2001; Laurent 2002).

Doving et al (Doving and Pinching 1973) were one of the first to report stimulus induced plasticity in vertebrates (rats) and revealed that a long exposure (several months) to an odorant led to specific

7

patterns of degeneration in the OB of adult rats. Later work (Vanderwolf and Zibrowski 2001) demonstrated that repeated ‘short term’ exposure to odour caused a gradual sensitisation of olfactory ß-waves in the pyriform cortex. This phenomenon occurred whilst the mucosal receptor potential was adapting (decreasing). More recently, an even shorter exposure regime (<8hrs) caused changes in the surface markers of olfactory receptor neurones (Yoshihara, Katoh et al. 1993) and expression of immediate-early genes in cells in the olfactory bulb (Sallaz and Jourdan 1993; Guthrie and Gall 1995; Baba, Ikeda et al. 1997). Buonviso et al (Buonviso and Chaput 1999) administered isoamyl acetate for 20 min pulses to rats for six consecutive days. This resulted in a decrease in responsiveness of the olfactory bulb, not only to the familiar odorant (demonstrating habituation or adaptation), but also to unfamiliar odorants. This effect was observed for up to 10 days after exposure and was attributed to an elevation in the inhibition of mitral/tufted (M/T) cell responses by periglomerular cells. Buonviso postulated that this inhibition was mediated by dopamine, or by GABA from granular cells (Buonviso and Chaput 1999). Inhibition was not only effected by different odorants, but also to a greater extent by the odorant concentration. This odorant mediated M/T cell inhibition was also observed in rat pups and was correlated to positive reinforcement in an associative learning task (Wilson and Leon 1988). Similar evidence exists in rabbits (Skarda and Freeman 1987) and, as with the Wilson study, electrophysiological recordings were associated with a learning task. We can therefore conclude that mitral-tufted cells have odorant receptive fields which are dynamic and can be modulated by odour experience (sensitisation and habitation). Arguably the dynamic nature of MT cells could directly affect OB spatiotemporal dynamics as well as cortical odour processing (Fletcher and Wilson 2003).

Several investigations have also examined the effect of odorant stimuli on ontogenetic development. Schmidt (Schmidt and Eckert 1988) compared the effect of the odorant geraniol in three groups of mice at varying stages in development. All datasets were referenced to an unexposed control. The groups were comprised of different age ranges in development, namely; Group 0-13 (postnatal day 0 to postnatal day 13 i.e. p0-p13); Group 0-6 (p0-p6) and Group 6-12 (p6-p12) – these groups received a continuous application of geraniol during the indicated periods. EEG recordings from the OB showed that Groups 0-13 and 6-12 demonstrated a significant response to geraniol whereas the control and Group 0-6 did not. Group 0-13 mice were also recorded at p70 (adults), the marked response was still evident suggesting that this period in murine olfactory development is plastic and can facilitate long lasting changes. Moreover, exposure and

8

deprivation of odours to rat pups led to a respective increase, (Rosselli-Austin and Williams 1990) and decrease, (Frazier and Brunjes 1988) in mitral and granule cell layer neurons. In fact, the spine density of granule cells was shown to be maximal at postnatal day 21 in rat pups. Continuous exposure to odourless air led to a reduction in spine density on the medial and lateral sides of the bulb whereas exposure to a single odorant, cyclohexanone, only reduced spine density on the lateral side (Rehn, Panhuber et al. 1988). Rather surprisingly, the chronic exposure of a single odorant for two months caused extensive cell shrinkage in the adult rat olfactory bulb when compared to the normal control. The shrinkage in mitral cells was even more pronounced than in rats exposed to an odourless environment. This suggests that an absence of stimulus is not as damaging as a single dominant odour (Panhuber, Mackay-Sim et al. 1987). Similar findings showed that continuous exposure of the odourant geraniol accelerated normal granule cell spine loss in the developing ferret olfactory bulb (Apfelbach and Weiler 1985). On the contrary, it was also shown that if adult mice (two months old) were exposed to an odorant enriched environment, twice as many newborn neurons survived in the main olfactory bulb. This was accompanied with an improved olfactory memory, however, it was not associated with an increased proliferation in the rostral migratory stream (RMS) and subventricular zone (SVZ) (Rochefort, Gheusi et al. 2002).

It was also shown, that repeated exposure to odorants increased peripheral olfactory sensitivity in mice (Wang, Wysocki et al. 1993) showing stimulus-induced plasticity in the receptor cell. Wang et al used inbred strains of NZB/B1NJ and CBA/J mice which are anosmic for, and sensitive to, androstenone, respectively. Shifts in sensitivity were electro-physiologically quantified by measuring field potentials from the olfactory epithelium – the electro-olfactogram (EOG) (Knecht and Hummel 2004). The male, adult mice were exposed to either androstenone or isoamyl acetate for 16 hours, daily for 2-6 weeks. EOG recordings from androstenone stimulated NZB/B1NJ mice demonstrated a 2.2-fold increase in amplitude, relative to the control. This sensitisation was not observed when the OE was stimulated with isoamyl acetate and was not demonstrated in the CBA/J androstenone stimulated mice i.e. they did not demonstrate increased sensitivity. It remains to be shown whether this effect is an androstenone specific phenomenon, nevertheless, this plasticity may be related to the continuous turnover of neurons, or alternatively it may be a form of stimulus-controlled expression. It is now known that the mechanisms of heptahelical receptor signalling are more diverse than previously thought, and that receptor stimulation can have various physiological consequences, like cytoskeletal rearrangements and gene expression (Lefkowitz 2000;

9

Hoyer, Hannon et al. 2002). These effects do not seem to be mediated by G-protein activation, but require the association of the receptors with a variety of other intracellular partners like arrestins, GRKs, SH2 proteins, small GTP-binding proteins, and PDZ domain–containing proteins (McDonald and Lefkowitz 2001; Rebois and Hebert 2003). Moreover, it is becoming increasingly clear that these proteins regulate the localization of the receptor and other key proteins in different subcellular compartments. Organization of GPCRs in supramolecular complexes was extensively studied using photoreceptor cells of the fruitfly (Harris and Lim 2001). The multivalent PDZ domain scaffolding protein INAD is required here to ensure the sensitivity and temporal resolution of the signal transduction process, most likely by coordinating the rhodopsin and other signaling molecules such as phospholipase Cb, protein kinase C, and the TRP ion channel into a multimeric complex (Harris and Lim 2001). In the last years, interactions with PDZ domain-containing proteins turned out to represent a generalized mechanism for modulating GPCR-mediated signal transduction (Zhong, Wade et al. 2003). Considering the high sensitivity and temporal resolution in vertebrate olfaction, one might expect to find a similar arrangement of olfactory receptors in supramolecular complexes in the sensory cilia of olfactory sensory neurons.

More recently there has been an increased interest in another family of scaffolding proteins, the arrestins. Two family members, β-arrestin 1 and β-arrestin 2, are ubiquitously expressed and are responsible for GPCR densensitization and internalization (Pouyssegur 2000). Arrestin binds to phosphorylated uncoupled GPCRs (βγ subunits) facilitating receptor internalization and interaction with mitogen-activated protein kinases (MAPKs). MAP kinases operate in modules performing a mini-cascade of sequential phosphorylations - MAP kinase kinase kinase (MKKK) activates MAP kinase kinase (MKK), which then activates MAP kinase which in turn has a myriad of intracellular effects including activation of c-Jun N-terminal kinase 3 (JNK3). JNK3 directly translocates to the nucleus instigating transcription (McDonald, Chow et al. 2000). Until very recently arrestin was considered to be exclusively located in the cytosol. However, exciting work has also demonstrated that β arrestin can directly translocate to the nucleus upon ligand stimulation of odorant receptors (Neuhaus et al, submitted). Nuclear translocation evidently results in histone H4 acetylation and gene transcription (Jiuhong Kang 2005).

To summarise, there have been previous reports which have investigated stimulus induced plasticity, however, these were mainly electrophysiological studies which focused on the olfactory

10

bulb. To date, there are relatively few investigations concerned with odorant induced plasticity in the olfactory epithelium and no proteomic reports. Therefore in order to address the question of differential protein regulation induced by OR activation, we applied a recent proteomic technology, fluorescent difference gel 2D-electrophoresis.

Figure 2.1 Schematic representation of potential stimulus induced plasticity of the olfactory sensory neuron What potential changes occur to protein regulation as a consequence of receptor stimulation? This question was addressed using a proteomic strategy.

2.2 Proteomics: an introduction The term “proteome” was initially coined by Wilkins who defined it as “…the entire PROTEin complement expressed by a genOME, a cell or tissue type, at a specific time in the development of the organism under specific physiological conditions” (Wilkins, Sanchez et al. 1996). Large-scale proteomics technologies have been developed to generate comprehensive, cellular protein-protein interaction maps, although “functional proteomics” in neuroscience is a relatively novel idea. Until relatively recently large scale proteomics was somewhat neglected in favour of using mass spectrometry (MS) to identify less complex samples. Interaction partners of the N-methyl-D-

11

aspartic acid (NMDA) receptor (Husi, Ward et al. 2000), of the P2X7 ATP receptor (Kim, Jiang et al. 2001), of the 5-HT2C receptor (Becamel, Alonso et al. 2002) and more recently, of mGluR5 protein complexes (Farr, Gafken et al. 2004) have been reported. The combination of classical biochemical techniques (e.g. pull-down, co-immunoprecipitation, cross-linking) with MS is clearly powerful, however, to look for changes in protein regulation in the olfactory epithelium clearly a large scale technology is required.

2.2.1 Technologies for large scale proteomics 2.2.1.1 2D-electrophoresis Two-dimensional polyacrylamide gel electrophoresis (2D-PAGE) was first described in 1975 and is still the ‘work horse’ in proteomics (Klose 1975; O'Farrell 1975). 2D-PAGE is a reproducible, reliable technique which produces high resolution protein separations. Proteins are separated in the first dimension according to protein charge as determined by a pH gradient (isoelectric point (pI)), and in the second dimension by molecular weight (SDS-PAGE). The technique has been refined (Klose and Kobalz 1995) and can now resolve up to 10,000 spots on one gel. Unfortunately, there are several criticisms of conventional 2D-PAGE. These include sensitivity, the linear dynamic range, reproducibility and accuracy of visualization methods. Moreover, it is well recognised that hydrophobic proteins, highly basic or acidic proteins cannot be easily identified using this technology.

Table 2.1 Comparison of the most commonly encountered detection methods used in 2D-PAGE NB: The silver stain provides only qualitative information about the proteins due to its narrow 10-fold linear dynamic range, while the fluorescent stain provides quantitative capabilities extending over 3-logs.

Stain Sensitivity Linear Range Coomassie Brilliant Blue G-250 30-100 ng ~10-30 X Colloidal Coomassie Brilliant Blue G-250 8-10 ng ~10-30 X Silver ~1 ng ~10 X SYPRO (fluorescent stain) ~3 ng ~103 X

Problems with reproducibly can be minimised with good standard operational procedures and handling technique, nevertheless, using conventional stains a large dataset is required. This is labour intensive and can also introduce inter-gel variability. Many of these limitations have been

12

relatively recently overcome with the emergence of a new fluorescent labelling technology, namely, DIGE - Difference gel electrophoresis (Marouga, David et al. 2005). This propriety technology (Amersham Biosciences) employs three fluorescent succinimidyl esters to specifically label proteins. These “CyDyes” (Cy2, Cy3 and Cy5) are spectrally resolvable and can therefore be mixed and run on one gel. Any differences between protein samples are directly reflected in the different ratios of the fluorescent signal. It is recommended that this method is used to compare two different biological samples, for example, treated versus control. The third dye is used to label an internal standard which allows accurate inter-gel matching. Protein solubility is maintained as this is a ‘minimal labelling’ technique with only ~1-2% of the lysine residues of the proteins being fluorescently modified. Furthermore, the singly charged CyDye ‘replaces’ the single lysine charge via a covalent NHS-ester linkage.

Figure 2.2 Fluorescent DIGE CyDye minimal labelling The CyDye minimal dyes have an NHS-ester reactive group and covalently couple to the ε-amino group of lysine of proteins via an amide linkage.

In addition, this technique is compatible with mass spectrometry, the penultimate step in protein identification. Either matrix-assisted laser desorption ionisation (MALDI) (Karas and Hillenkamp

13

1988) and/or electrospray ionisation (ESI) (Whitehouse, Dreyer et al. 1985; Fenn, Mann et al. 1989) mass spectrometry may be used to ‘fingerprint’ or sequence the protein. The resultant spectra are integrated into an algorithm which searches a protein database e.g. MASCOT or SEQUEST. Before these ‘soft-ionization’ techniques were reported, the only reasonable alternative was Edman sequencing which, due to the time and cost, was prohibitive. Proteins can now be reliably resolved using 2D-electrophoresis and subsequently identified using ESI and/or MALDI.

Figure 2.3 Schematic of 2D-Electrophoresis workflow

2.2.1.2 Multi-dimensional protein identification technology MudPIT An interesting technique for the analysis of complex protein mixtures is 'shotgun proteomics'. This approach has been facilitated by the use of multidimensional protein identification technology (MudPIT), which incorporates multidimensional high-pressure liquid chromatography (LC/LC), tandem mass spectrometry (MS/MS) with database-searching algorithms (Washburn, Wolters et al. 2001) and facilitates the concomitant identification of soluble and membrane associated proteins (Wu, MacCoss et al. 2003). In MudPIT a complex protein mix (e.g. cell lysate) is digested ‘in-

14

solution’ and separated by multidimensional capillary chromatography which is connected on-line to an ion trap mass spectrometer. The term ‘multidimensional’ is attributed to the fact that there are two dimensions in the chromatography step, however, rather than using conventional 2D-gel electrophoresis, MudPIT separates peptides in 2D-liquid chromatography. Thus, the column can be directly interfaced to a mass spectrometer for electrospray ionisation tandem MS i.e. LC/LC- ESI-MS/MS. The first dimension consists of strong cation exchange (SCX) material. This is packed ‘back-to-back’ with the second dimension, a reversed phase (RP) material inside fused silica capillaries. Peptides are therefore resolved on the column according to their charge (SCX) and hydrophobicity (RP).

Despite the power of this technique it does have limitations. There are some criticisms over the scalability and reproducibility of MudPIT (Graumann, Dunipace et al. 2004). In addition, concerning the quantification of subtle changes in protein regulation and/or in post-translational modifications, undoubtedly LC-MS techniques lag behind gel based techniques.

Figure 2.4 Schematic of MudPIT Workflow

15

2.2.2 Neuro-proteomics The proteomic methods described above can be easily applied to global profiling of many samples including neurological probes. Already early proteomic studies have attempted to map protein expression in various brain regions including the adult cerebellum from mouse (Beranova- Giorgianni, Pabst et al. 2002), rat (Taoka, Wakamiya et al. 2000) and pig (Friso and Wikstrom 1999; Morrison, Kinoshita et al. 2002). An international concerted effort (HUPO Brain Proteome Project) is now underway to differentially analyse normal and neurodegenerative pathologies, with a particular emphasis on Morbus Alzheimer, Morbus Parkinson and senescence (Michael Hamacher 2005). Several studies have used human tissue, a recent example is a comparative study of the prefrontal white matter of human alcoholics referenced to normal subjects (Alexander-Kaufman, James et al. 2005). 60 statistically different (p<0.05, ANOVA) proteins were identified, several of which were novel findings. The vast majority of neuro-proteomic studies have opted to use samples surgically removed from individual subjects. An additional approach is the use of cultures which is gaining in popularity. Recently Sitek et al stably transfected a neuroblastoma cell line (SY5Y) with TrkA or TrkB receptors (Sitek, Apostolov et al. 2005). The cells were stimulated with their respective ligands (TrkA-NGF and TrkB-BDNF) and using fluorescence two-dimensional difference gel electrophoresis (DIGE), differences in protein regulation were quantified over a variety of post-stimulation time points. Employing the same technology, another elegant study demonstrated the importance of the cytosolic protein stathmin in the migration of newborn neurons in the adult brain (Jin, Mao et al. 2004). In this case, a primary neuronal culture from p15 CD1 mice was established. ‘Immature’ neuronal cultures (cells harvested 5-6 days after plating) were compared with the ‘mature’ population (four passage cycles) – stathmin was up-regulated 7.6- fold in immature cultures. This evidence was corroborated by western blots and immunocytochemisty. Confocal images of sagital sections from neuroproliferative zones of normal adult rat brain exhibited prominent stathmin staining. Moreover, stathmin colocalized with the immature neuronal marker DCX in recently divided cells (BrdU positive). Conversely, stathmin did not colocalize with the astroglial marker GFAP, the neuropeithelial cell marker nestin, or the mature neuronal marker NeuN. Knockdown of stathmin expression with an antisense oligodeoxynucleotide (ODN) inhibited the migration of DiI-labeled cells from subventricular zone (SVZ), via the rostral migratory stream (RMS), to the olfactory bulb (OB). This comprehensive

16

study clearly exemplified the power of fluorescent DIGE in identifying candidate proteins (Jin, Mao et al. 2004).

Only one differential proteomic study has been reported with regard to the olfactory system. Recently, Poon examined the differential expression of proteins in the aging murine olfactory system (Poon, Vaishnav et al. 2005). Olfactory epithelium and olfactory bulb from C57BL/6 ‘young’ mice (6-week-old) were compared to OE and OB from ‘old’ mice (80-week-old) using conventional 2D electrophoresis. In the OE and OB 9 and 20 significantly regulated proteins were identified, respectively. These cytosolic proteins were categorised into three functional groups, namely, metabolic (primarily mitochondrial metabolism); transport/motility; and stress response. The alterations in the protein profile of the aging olfactory epithelium and olfactory bulb are specific to pathways involving mitochondrial and cytosolic metabolism, intracellular transport and altered response to stress. Some of these mitochondrial protein changes were already reported in microarray analysis (Kopsidas, Kovalenko et al. 2000) reinforcing the argument that oxidative stress plays an integral role in aging. This study did identify several candidate proteins; however, most of these findings were not novel with regard to senescence. Moreover, conventional 2D lacks the sensitivity and dynamic range of fluorescent labelling and therefore subtle changes in lower abundant proteins could not be identified by this methodology. Tun-Tzu Yu, on the other hand, used quantitative reverse transcriptase-PCR in order to identify differentially expressed transcripts in OSNs (Tun-Tzu Yu 2005). OSNs were purified from juvenile mice using OMP-LacZ3 transgenic mice - OSNs specifically express β-galactosidase due to LacZ being ‘under’ the promoter of olfactory marker protein – (OMP). The FACS (fluorescence-activated cell sorting) purified OSNs were compared to neighbouring cells in the OE. Of the differentially expressed transcription products, 33% had no known function, the remaining encoded for proteins involved in signalling, xenobiotic metabolism, protein metabolism and vesicle transportation. This study identified more interesting candidates than the Poon investigation, however, the paradigm of exclusively characterising proteins using their sequences is inadequate. To exemplify, with only approx. 25,000 genes in the human genome, the proteome is clearly much more complicated. Variable gene expression may result from mRNA alternative splicing, frame-shifting and the use of alternate translation start or stop sites. Proteins may undergo an array of modifications including phosphorylation, glycosylation, acetylation, nitrosation, farnesylation and sulfation. Up to 300 different post-translational modifications have been reported (Morrison, Kinoshita et al. 2002).

17

Such modifications will affect protein conformation and functionality. Furthermore, it is generally accepted that transcription does not necessary directly correlate to protein expression and protein functionality.

2.3 Conclusion The aim of this project is the analysis of stimulus induced plasticity of the olfactory epithelium (OE). To address this question, mice were exposed to either a continuous or pulsed application of an odorant, octanal, for 20 days. Juvenile mice were selected for this study (p0-p30) as this stage in olfactory development is sensitive to odorant triggered plasticity. These changes may even last into adulthood. In order to elucidate changes in protein regulation which are due to the treatment conditions, a proteomic technique compared the OE from treated mice to a sham treated control. We used fluorescent difference gel electrophoresis (DIGE) which utilises spectrally resolvable fluorescent dyes to differentially label proteins from the OE of treated and control mice. Significantly up- or down-regulated proteins can be subsequently identified using mass spectrometry.

2.4 Materials and Methods 2.4.1 Animal and Tissue Preparation All procedures were in accordance with German animal welfare act (1998). 15 timed pregnant C57BL/6 mice were purchased from Charles River, Germany. All animals were held in standard rat cages at an animal facility at room temperature (22oC) and maintained on 12-hour light/dark cycle. Water and food was given ad libitum. The dams were separated into three groups of five mice per cage – Continuous Group A, Pulsed Group B and Control Group C. Each cage was housed in a customised hood designed to deliver an isolated air stream, the ‘exhaust’ air was extracted to a fume cupboard. A picture of the apparatus follows (Figure 2.5).

The apparatus was designed in order to ensure that the vapour phase concentration of the odorant remained consistent throughout the experiment. All tubes were silicon coated to minimise odorant adsorbance and the hood was constructed from Perspex glass (Biology Werkstatt, RUB). The day of pup delivery was considered P0. The odorant regime commenced at P10 and was terminated at

P31 – the aldehyde Octanal (Sigma-Aldrich) was exclusively used at a 1:10,000 dilution in ddH20. Freshly made odorant solution was prepared every morning. For Group A (Continuous): Octanal

18

was applied for 24 hours/day for the entire duration of the experimental period. For Group B (Pulsed): Octanal was applied four times daily at the following time periods - 09:00-09:15; 12:00- 12:15; 15:00-15:15; 18:00-18:15. Group C (Control): were sham exposed to water for the entire duration of the experimental period. It was essential that handling was minimised to reduce stress, however, when absolutely necessary (e.g. new bedding required) each group was handled to the same extent. At P31 pups were taken for behavioural analysis and EOG recordings (see below). At P31 the total mean body weights for each group was as follows: pulsed 17.25g; continuous 16.18g; control 17.76g.

Figure 2.5 Picture of mice odorant exposure apparatus

2.4.2 Behavioural study P31 continuously (n=38), pulsed (n=33) and control (n=30) mice were starved for 8 hours and subjected to an exploration test as previously reported (Yamada, Wada et al. 2001; Dawson, Steane et al. 2005), with minor modifications. One control mouse was removed and placed into a test arena (rat cage containing 3 cm freshly laid bedding; Goldspan) which contained either a hidden 1.5ml vial of cheese, (stilton blue cheese, Globus, Germany) or octanal (1:500; Sigma). The vial was randomly placed in one of the four cage corners at a depth of 2 cm and the time to find the hidden food/octanal was measured. If the exploratory time exceeded 600 sec, the mouse was removed

19

and ‘failure to find’ was recorded as “600 sec” for that trial. Each control mouse (male and female) was tested for octanal and cheese; 50% of the control group (n=15) were first tested with octanal and after a short interval, cheese. The remaining 50% were first tested with cheese followed by octanal - this design eliminated any statistical bias due to the test sequence. The trials were repeated for all mice in each of the treatment groups (continuous and pulsed). The bedding was renewed ad hoc to ensure that ‘distracting’ odours such as urine and faeces were eliminated.

2.4.3 Microdissection of olfactory epithelium At P32 all pups were sacrificed by cervical dislocation followed by decapitation. The OE was dissected as previously described (Spehr, Wetzel et al. 2002). Briefly, the septal bone with the intact olfactory epithelium was dissected from the head and either used intact for EOG recording, or the epithelium was dissected from the septal bone and directly placed into 100μl ice-cooled dissection buffer in a 1.5ml eppendorf. The buffer consisted of ringers [NaCl 140 mM, KCl 5 mM, MgCl2 1 mM, CaCl2 2 mM, 10 mM Hepes, 10 mM glucose] with Roche Complete® protease inhibitor cocktail and phosphatase inhibitors [β-Glycerophosphate 10 mM, Na-orthovanadate 1 mM, NaF 9.5 mM, Na-pyrophosphate 10 mM – Sigma Aldrich]. After the sample weight was recorded the OE was snap frozen in liquid nitrogen and stored at -80oC until preparation for 2D-electrophoresis.

2.4.4 Electro-olfactogram (EOG) recording For EOG recording, the exposed and its associated epithelium was placed in a recording chamber continually perfused with oxygenated mammalian ringer (95% O2/5% CO2) (in mM) 120 NaCl, 25 NaHCO3, 5 KCl, 5 BES, 1 MgSO4, 1 CaCl2, and 10 mM glucose. Field potentials were recorded using glass pipettes (1 MΩ; tips filled with extracellular solution in 1% agar) which was placed on the airface of the epithelium. Electrophysiological signals were transduced by a unipolar silver /silver chloride (Ag/AgCl) electrode ‘housed’ in the glass pipette; the DC output was differentially amplified (P18C, Grass Instrument Co.) and referenced to an identical electrode (Ag/AgCl) grounding the bath. Odorants were delivered in the vapour phase as described previously (Spehr, Wetzel et al. 2002). A 1:10,000 dilution of octanal was prepared in ringers and soaked on a filter paper which was inserted into a motor-controlled perfusion syringe (Braun Melsungen). The odorant was injected into a constant stream of humidified air that continuously superfused the epithelium. Signals were recorded on a chart recorder and their peak amplitude was measured manually. The peak amplitudes from 1:20,000, 1:2,000 and 1:200 octanal

20

dilutions were normalised to a final application of 1:100 Henkel 100 (complex mixture of 100 odorants).

2.4.5 Sample preparation for DIGE

The OE from one mouse (single sample) was ground using a pestle and mortar under l N2 with sample buffer (30mM Tris Base, 2M Thiourea, 7M Urea, 4% CHAPS (w/v) [Biorad, Germany] – pH8.5). The amount of lysis buffer was critical –1.43 times (w/v) the sample weight. The mixture was thoroughly ground, the frozen power transferred to a 1.5mL eppendorf tube and sonicated in an ice-cooled sonication bath for 10 seconds. The sample was immediately placed in ice for another 10 seconds before repeating this procedure five more times. The sample was allowed to stand on ice for 2 minutes before centrifugation at 16,000 X g; 15min in a 4oC cooled microcentrifuge (Centrifuge 5415R, Eppendorf, Hamburg, Germany). After standing on ice for 1 min the supernatant was removed and stored on ice for protein concentration. The sediment was resuspended in X 1 (w/v) lysis buffer (i.e. equal to the weight of the original OE sample) and was deposited on the lN2 cooled mortar. The homogenisation procedure was repeated as before. The protein concentration from both aliquots was determined using the Biorad Bradford assay according to the manufactures instructions (Biorad, Munich, Germany). If a significant concentration (>15μg/μl) of protein was measured in the second ‘sediment’ supernatant the samples were pooled, otherwise, it was discarded. The typical total protein yield was between 20-35 μg/ul in the first preparation, the second preparation was usually <10 μg/ul. The samples were stored at -80oC until fluorescent DIGE labelling.

2.4.6 Protein Labelling - Fluorescent DIGE minimal labelling The samples were minimally labelled according to the manufactures instructions (Amersham Biosciences, Freibrug, Germany). Each sample, which represents the cell lysate from one mouse, was individually labelled. For statistical significance the total sample size was 5 (n=5). Briefly, stock cyanine dyes (1 nmol/μl) were diluted in freshly prepared anhydrous dimethyl formamide (DMF) p.a. to 400 pmol/μl (Sigma, St. Louis, MO), and 8 pmol dye was added per μg protein in the cell lysate. The sample was vortexed, centrifuged briefly, and left on ice for 30 min in the dark. No primary amines, DTT or carrier ampholytes were included in the lysis buffer as such components could potentially react with the NHS esters of the cyanine dyes. The labelling reaction was quenched by adding 1μl per 400 pmol dye of 10 mM L-lysine (Sigma). The treated (pulsed or

21

continuous) and untreated (control) samples were labelled with Cy5 and Cy3, respectively; the internal reference (mixture of control and treated) was Cy2 labelled. After further vortexing and centrifugation the sample was left on ice for 10mins in the dark. Thereafter, the Cy5, Cy3 and Cy2 labelled samples were mixed 1:1:1 (equating to 150μg total protein) and prepared for iso-electric focussing. To summarise, one CyDye mix comprised of: i) 50 μg treated OE from one mouse (pulsed or continuous) – Cy5; ii) 50 μg untreated control OE from one mouse – Cy3; iii) 50 μg OE from a pool of ALL sample (internal reference) – Cy2. Comparisons were made between continuous versus control and pulsed versus control. For each comparison n=5 i.e. a total of 10 gels were run.

2.4.7 IEF and 2-DE To the mixed labelled samples 10% (v/v) DTT (1,08 g/ml; BioRad) and 10% (v/v) Ampholine 2-4 (Amersham Bioscience) was added. IEF was performed using tube gels (20 cm X 0.9 mm) containing carrier ampholytes (CA-IEF) and applying a voltage gradient in an IEF-chamber produced in-house (Klose and Kobalz 1995). After IEF, the ejected tube gels were incubated in equilibration buffer (125 mM Tris, 40% (w/v) glycerol, 3% (w/v) SDS, 65 mM DTT, pH 6.8) for 10 min, after which they were washed threefold using SDS-PAGE running buffer (25 mM Tris, 192 mM Glycine, 0.2% SDS at room temperature) before transferring to the second dimension. The second dimension (SDS-PAGE) was performed on (15.2% total acrylamide, 1.3% bisacrylamide) polyacrylamide gels using a Desaphor VA 300 system. IEF tube gels were placed onto the polyacrylamide gels (20 cm X 30 cm X 0.7 mm) and fixed using 1.0% (w/v) agarose containing 0.01% (w/v) bromphenol blue dye (Riedel de-Haen, Seelze, Germany). The temperature of the running buffer was thermally regulated for the duration of the run (typically 5 hours at 200v).

NB: For protein identification by mass spectrometry a larger amount of protein was required. 500μg of total protein was subjected to 2D-electrophoresis and subsequently stained using an MS compatible (see below) stain. For protein identification DIGE staining was not employed.

2.4.8 Gel scanning, digitising and analysis The resultant 10 sample gels (5 pulsed Vs control; 5 continuous Vs control) were wiped with MEK ethanol and subsequently scanned using a Typhoon 9400 scanner (Amersham Biosciences). The appropriate excitation and emission wavelengths were chosen specifically for each of the dyes as

22

recommended by the manufacturer. Images were scanned at a resolution of 100μM and pre- processed using ImageQuantTM software (Amersham Biosciences). Processed, cropped images were imported into the Differential In-gel Analysis (DIA) software module (Amersham Biosciences). The DIA software module provides consistent and accurate co-detection of image pairs from the same gel using novel algorithms. Background subtraction, quantification, normalization and first level matching within one gel are automated allowing for high-throughput analysis with low experimental variation. The subsequent DIA analysed dataset and images were imported into the Biological Variation Analysis (BVA) module. This facilitates batch-processing, up to several hundred gel pair-images can be processed unattended for significant time savings. Images are then matched between gels using the BVA software module that looks for consistency of differences between samples across all the gels and applies statistics to associate a level of confidence for each of those differences.

Spot intensities were normalized to the internal standard (Cy2 labelled) and significantly regulated proteins were identified by Student T-test (p<0.05). Protein spots differentially expressed between the treated and untreated groups were flagged for spot excision and protein identification by MS.

2.4.9 2D-SDS PAGE Gel total protein staining In the course of optimisation a mass spectrometry incompatible stain was employed (Swain and Ross 1995). This simple sensitive procedure includes the cross-linker glutaraldehyde which is incompatible with MS. For protein identification an MS compatible silver stain (“Blum”) was employed as previously reported (Nesterenko, Tilley et al. 1994). The staining protocols were performed as described in the literature.

2.4.10 Trypsin digestion Spots which were significantly regulated by a ratio of 1.5 or more were manually excised from the preparative identification gel using a clean spot picker and transferred to clean glass mini-tubes. Spots were destained and digested with trypsin (Promega, Madison, WI) as previously described (Gharahdaghi, Weinberg et al. 1999). Briefly, spots were destained with a 1:1 working solution of 30mM potassium ferrocyanide and 100mM sodium thiosulphate (Merck, Darmstadt, Germany). A sufficient volume to cover the spots was added (30-40μL) and allowed to sit for 1-2 min. The spot was equilibrated with 10mM ammonium bicarbonate – ‘Ambic’ (Merck, Darmstadt, Germany) for

23

10 min, followed by dehydration solution (Ambic:acteonitrile [ACN] 1:1). This was repeated once more and subsequently dried by speedvac (Savant). Trypsin digestion was performed with 5± 10 ng/mL of trypsin and 10 mM ammonium bicarbonate and incubated overnight at 37oC. The peptides were extracted in 20μL extraction solution (5% Trifluoroacetic acid in ACN) and placed in a sonication bath for 10 min. This was repeated twice, the pooled extracted extracts were briefly reduced by speedvac and analysed by ESI-MS/MS.

2.4.11 Mass Spectrometry Materials and Reagents. Acetonitrile (ACN, HPLC-S gradient grade) was purchased from Biosolve (Valkenswaard, Netherlands). Formic acid (FA, Baker Analyzed) was obtained from J.T. Baker B.V. (Deventer, Netherlands), and trifluoroacetic acid (TFA, for protein analysis) was obtained from Merck (Darmstadt, Germany). Sequencing grade modified trypsin was purchased from Promega (Madison, WI, USA). Mass calibration standards were obtained from Applied Biosystems (Foster City, CA, USA). Water was used from an ELGA Labwater filtration system (Vivendi Water Systems, Ransbach-Baumbach, Germany). All other reagents were acquired from commercial sources.

Mass Spectrometric Analysis. Tandem MS (MS/MS) analyses were performed on a Bruker Daltonics HCT plus Ion Trap (Bremen, Germany) system operated in the sensitive mode. A nano electrospray ionization (nano-ESI) source (Bruker Daltonics, Germany) equipped with distal coated SilicaTips (FS360–20–10-D; New Objective) was used in conjunction with online capillary HPLC with the following mass spectral parameters: Capillary Voltage (CV), 1400 V; End Plate Offset (PO), 500 V; Dry Gas (DG), 10.0 l/min; Dry Temperature (DT), 160°C; aimed Ion Charge Control (ICC) 150000; maximal fill-time 500 ms. Online reversed-phase capillary HPLC separations were performed with the Dionex LC Packings HPLC systems (Dionex LC Packings, Idstein, Germany) as previously described by Schäfer et al (Heike Schaefer 2004).

MS spectra were a sum of seven individual scans ranging from m/z 300 to m/z 1400 with a scanning speed of 8,100 (m/z)/s. For MS/MS analyses, data-dependent software (HCT plus, Esquire Controle, Bruker Daltonics, Germany) was employed to select the two most intense, multiple-charged peptide ions detected in MS spectra. To generate fragment ions, low-energy collision induced dissociation (CID) was performed on previously isolated peptide ions by applying

24

a fragmentation amplitude of 0.6 V. Generally, MS/MS spectra were a sum of four scans ranging from m/z 100 to m/z 2200 at a scan rate of 26,000 (m/z)/s. Exclusion limits were automatically placed on previously selected mass to charge ratios for 1.2 minutes. The ion trap instrument was externally calibrated with commercially available standard compounds.

For protein identification, uninterpreted peptide MS/MS spectra were correlated with the NCBI protein sequence database (http://www.ncbi.nlm.nih.gov) restricted to Mus musculus applying the SEQUESTTM algorithm (Eng 1994). The reliability of protein identifcation via database searching with SEQUESTTM was checked manually, and sequence coverages reported correspond to the number of amino acids identified in proteins.

2.4.12 Statistics Factorial ANOVA analysis was performed using version 6.0 of Statistica software, StatSoft (Europe) GmbH., Hamburg, Germany. 3D-charts were plotted using the statistical software SPSS, SPSS GmbH Software, Munich, Germany.

2.5 Results A continuous and pulsed odorant exposure regime was designed in order to distinguish differences in protein regulation which are due to stimulus induced plasticity of the olfactory sensory neuron. Two groups of male juvenile C57BL/6 mice were exposed to octanal from postnatal day 10 (p10) for a duration of 20 days. One group received a constant supply of odorant (‘continuous group’), the other group received pulsed odorant applications - 4 x 15 minute daily pulsed treatment (‘pulsed group’). A third control group received sham continuous exposure to water, the ‘control group’. It was hypothesized that the continuous experiment induces long term adaptation whereas the pulsed treatment induces peripheral sensitization to the stimulant.

Octanal* (Figure 2.6) was selected as a stimulant for several reasons. It has been shown to activate the I7 receptor has in both mice and rat (Krautwurst, Yau et al. 1998) and was subsequently functionally characterised. This included molecular modelling of the ligand’s binding site (Singer 2000), elucidation of its receptive field (Araneda, Peterlin et al. 2004) and mapping of the glomeruli response to the stimulus (Johnson, Farahbod et al. 2004). Furthermore, data from previous work

* Octanal has several synonyms – Octylaldehyde, 1-Octanal, Aldehyde C8, Capryl aldehyde & N-Octylaldehyde

25

from our laboratory indicated that a significant proportion of wild type murine OSNs (approx. 7%) respond to a 1:10,000 dilution of Octanal (Wetzel, Brunert et al. 2005). It was also shown that rat OSNs can respond to several aliphatic aldehydes (carbon chain length C5-C10) with the greatest number of cells responding to octanal (Kaluza and Breer 2000).

Figure 2.6 The structure and properties of Octanal.

In order to identify and quantify differences in protein regulation between treated and control mice fluorescent difference gel electrophoresis (DIGE) was performed. The olfactory epithelium from treated mice were labelled and compared to the OE from control mice using DeCyderTM software suite (Amersham Biosciences); the labelling strategy follows-:

Pulsed Treated Mice (Cy5 Labelled - Red) Vs Control Untreated (Cy3 Labelled - Green)

Continuous Treated Mice (Cy5 Labelled - Red) Vs Control Untreated (Cy3 Labelled - Green)

Internal Reference/Standard (comprising all analysed samples) – Cy2 (yellow)

26

In each experiment the sample size was five (n=5). Each gel was scanned three times (Cy5, Cy3, Cy2) allowing multiplexing, dataset normalisation and gel matching. Spot volume is normalised using DIATM software. DIA analyses images from a single gel detecting and quantifying spots. Spot volumes are calculated by the pixel intensity relative to the spot boundary; this data also takes into account the volume, slope and peak height of the spot (Figure 2.7). This “co-detection” algorithm acquires spot data from each CyDye scan and reports the ‘spot ratio’ – changes between Cy5 (treated) and Cy3 (untreated) gel images relative to the internal reference (Cy2) i.e. the ratio value of a spot pair. This data is normalised so that the modal peak volume of ratios is 0. The normalized data is referred to as the ‘volume ratio’ – any increase or decrease in protein abundance is consequently reflected as an x-fold change from 1 to ∞ or from -1 to ∞, respectively. To exemplify, a two-fold up-regulation of a spot would be reported as “2.00” i.e. this two-fold increase is relative to the standard; a ‘standardised abundance’.

Figure 2.7 Schematic representation of parameters used to calculate spot data using the DIA (Differential In-gel Analysis) Module

27

.

Figure 2.8 Example of two DIGE gels. Upper: Pulsed treated mouse (n=1) Lower: Continuously treated mouse (n=1). In both cases treated and control mice were labelled using Cy5 (red) and Cy3 (green), respectively.

28

Figure 2.8 above demonstrates two multiplexed images (Cy3, Cy5 & Cy2) from the pulsed and continuous experiments (n=1). Fluorescent green and red spots indicate control and treated labelled samples, respectively. Yellow spots reflect no significant change in protein regulation. Significant changes (+/- 1.5 fold) in spot intensity were manually validated. For the continuous experiment a total of 18 significantly regulated spots were identified: 11 down-regulated and 7 up- regulated. For the pulsed experiment a total of 29 significantly regulated spots were identified: 4 down-regulated and 25 up-regulated

2.5.1 Regions of interest To illustrate the power of DIGE in differential analysis, two regions from the continuously (region A) and pulsed treated (region B) mice were selected for elaboration. To reiterate, 5 gels were run for each experiment (n=5) ensuring that any changes are reproducible and statistically significant.

Figure 2.9 DIGE scan from 5 gels comparing the control (green) and pulsed (red) treated mice. Underneath each fluorescent image is the corresponding 3D-intensity chart which schematically represents changes in intensity.

29

From Figure 2.9 it can be seen that the magnitude of the control signal amplitude is greater than the treated signal i.e. there was a significant down-regulation of this region as a consequence of pulsed octanal treatment. This striking effect was demonstrated across all gels, however, it was only statistically significant in 4/5 gels. Thus, these spots were ‘flagged’ for excision and identification by mass spectrometry. Area of interest B (Figure 2.10), on the other hand, demonstrates non- reproducible and insignificant changes for one spot across five ‘pulsed’ gels. In 2 of the pulsed gels this spot was up-regulated. This was not the case for the remaining pulsed treated mice where either there was no significant change or a moderate down-regulation.

Figure 2.10 Biological variation. DIGE scan from 5 gels comparing the control (green) and pulsed (red) treated mice. This spot was not significantly regulated

Spot #1 (Figure 2.11) demonstrated the most striking difference, a significant up-regulation. From the standardized log abundance chart this effect was not as marked as in the other pulsed samples.

30

#1 was analysed by ESI-MS/MS and identified as sulfotransferase family, cytosolic, 1C, member 1 - SULT1C1 (Swiss-Prot Accession Number: Q80VR3). This demonstrates the power of DIGE in detecting subtle changes in protein regulation.

Figure 2.11 Biological variation. Histogram shows differences in standardized log abundance across pulsed gels. Gel images and 3D diagram are from one gel - Spot #1 (region of interest B)

31

Figure 2.12 Master gels showing statistically significantly up- and down-regulated spots. The spots were arbitrarily numbered & excised for protein identification using mass spectrometry. Magnified area on pulsed gel was significantly up-regulated and identified as two cytoskeltal proteins, K8 & K18, which are clearly subjected to post-translational modifications (PTMs)

32

2.5.2 Pulsed Gel Analysis Significantly regulated proteins were excised (Figure 2.12), tryptically digested and peptide sequenced using ESI-MS/MS. Using the SEQUESTTM algorithm the resultant spectra were used to search the public NCBI Mus musculus database. The following table (Table 2.2) indicates the spot number and its degree of regulation for the pulsed treated samples. The corresponding gel images and differential abundances are also shown.

Table 2.2 Statistically significant (P<0.05) changes for pulsed experiment.

CONTROL PULSED SPOT No./Fold STANDARDIZED LOG Change Ratio ABUNDANCE (Cy3) (Cy5)

1

7.98

2

2.21

3

2.2

33

CONTROL PULSED SPOT No./Fold STANDARDIZED LOG Change Ratio ABUNDANCE (Cy3) (Cy5)

4

2.56

5

-2.83

6

2.49

7

3.89

8

2.38

34

CONTROL PULSED SPOT No./Fold STANDARDIZED LOG Change Ratio ABUNDANCE (Cy3) (Cy5)

9

4.04

10

4.78

11

7.87

12

4.90

13

3.74

35

CONTROL PULSED SPOT No./Fold STANDARDIZED LOG Change Ratio ABUNDANCE (Cy3) (Cy5)

14

2.85

15

3.87

16

4.51

17

4.88

18

4.37

36

CONTROL PULSED SPOT No./Fold STANDARDIZED LOG Change Ratio ABUNDANCE (Cy3) (Cy5)

19

2.81

20

3.06

21

2.24

22

2.24

23

37

CONTROL PULSED SPOT No./Fold STANDARDIZED LOG Change Ratio ABUNDANCE (Cy3) (Cy5)

24

2.46

25

2.58

26

-16.69

27

-14.79

28

2.28

38

CONTROL PULSED SPOT No./Fold STANDARDIZED LOG Change Ratio ABUNDANCE (Cy3) (Cy5)

29

-19.76

2.5.3 Continuous Similarly, significantly regulated proteins from the continuously treated mice were analysed as described for the pulsed treated samples. The following table (Table 2.3) indicates the spot number and its degree of up- or down-regulation for the continuously treated samples. The corresponding gel images and differential abundances are also shown.

Table 2.3 Statistically significant (P<0.05) changes for continuous experiment.

SPOT CONTINUOUS CONTROL STANDARDIZED LOG No./Fold ABUNDANCE Change Ratio (Cy5) (Cy3)

1

-6.10

39

SPOT CONTINUOUS CONTROL STANDARDIZED LOG No./Fold ABUNDANCE Change Ratio (Cy5) (Cy3)

2

-5.45

3

-7.95

4

-5.97

5

-1.73

6

1.42

40

SPOT CONTINUOUS CONTROL STANDARDIZED LOG No./Fold ABUNDANCE Change Ratio (Cy5) (Cy3)

7

1.44

8

-4.92

9

1.48

10

1.50

11

1.68

41

SPOT CONTINUOUS CONTROL STANDARDIZED LOG No./Fold ABUNDANCE Change Ratio (Cy5) (Cy3)

12

1.42

13

-7.42

14

-3.46

15

-2.90

42

SPOT CONTINUOUS CONTROL STANDARDIZED LOG No./Fold ABUNDANCE Change Ratio (Cy5) (Cy3)

16

1.46

17

-6.80

19

-27.90

2.5.4 Protein identification In order to identify proteins, DIGE gels containing 150 μg total protein were scaled up to preparative gels where a total of 500 μg protein was loaded. A greater amount of protein is required for MS detection. For individual proteins the detection limit using silver stain (1-2ng) may give rise to tryptic fragments, nevertheless, greater amounts of protein increase the chances for MS identification. Rather unexpectedly, scaling up using identical 2D conditions led to poor protein resolution and streaking (Figure 2.13, bottom image). In order to overcome this, 20% DTT was added to the first dimension (reducing streaking) and a 4% stacking gel was employed in the second dimension (improving SDS-PAGE resolution). The resultant empirically optimised gel (Figure 2.13, top) allowed identification of 96% of the significantly regulated proteins (45/47).

43

Figure 2.13 ‘Troubleshooting’ protein identification. Blum silver stained preparative gels (500 μg total protein) – Bottom: using identical conditions as with DIGE; Top: using optimised conditions

44

The spots of interest were excised, digested and the peptides subjected to ESI-MS/MS. Spectra were used to search the NCBI mouse genome using the SEQUESTTM algorithm. The results for both odorant conditions are tabulated below (Table 2.4). In some cases two proteins were identified from one spot. Without independent validation, for example using western blotting, the significantly regulated protein is therefore unknown.

Table 2.4 Protein identification from pulsed experiment. For each protein the corresponding theoretical isoelectric point (pI) molecular mass (kD) & degree of up- or down-regulation (fold change) are indicated. All regulated spots were significant as determined by the Student T-test.

Spot Protein Name Theoretical Fold change T-test P

No pI/Mr ratio value

1 UNIDENTIFIED n/a 7.98 0.023

2 Tumor rejection antigen gp96 4.7/92432 2.21 0.0022

3 Gelsolin precursor (Actin-depolymerizing factor) (ADF) 5.8/80751 2.20 0.011

4 Hydroxyacyl-Coenzyme A alpha subunit 9.2/82617 2.56 0.045

Poly(A) binding protein, cytoplasmic 1 9.5/70655

5 Enolase 2, gamma neuronal 5.0/47267 -2.83 0.013

6 Cytokeratin endo B 5.3/47460 2.49 0.0047

7 Cytokeratin 5.2/47509 3.89 0.0019

8 Cytokeratin endo B 5.3/47460 2.38 0.0073

9 Cytokeratin endo A 5.7/54531 4.04 0.0019

10 Cytokeratin endo B 5.3/47460 4.78 0.00025

11 Vimentin protein 5.1/53641 7.87 0.048

45

Spot Protein Name Theoretical Fold change T-test P

No pI/Mr ratio value

12 Cytokeratin endo B 5.3/47460 4.90 0.00028

13 UNIDENTIFIED n/a 3.74 0.00048

14 Cytokeratin endo A 5.7/54531 2.85 0.00056

15 Cytokeratin endo A 5.7/54531 3.87 0.002

16 Smooth-muscle alpha tropomyosin 4.7/32655 4.51 0.0022

17 Cytokeratin endo A 5.7/54531 4.88 0.00093

18 Cytokeratin endo A 5.7/54531 4.37 0.00084

19 Cytokeratin endo A 5.7/54531 2.81 0.000082

20 Annexin A5 4.8/35730 3.06 0.0012

21 Carbonyl reductase 2 9.1/25941 2.24 0.0047

22 Carbonyl reductase 2 9.1/25941 2.24 0.00074

23 Calbindin 2 4.9/31352 2.33 0.017

24 Alpha-globin 8.0/15117 2.46 0.023

Carbonyl reductase 2 9.1/25941

25 CCG1-interacting factor B 5.8/22436 2.58 0.00072

26 Adenylate kinase 1 5.7/23101 -16.69 0.005

27 ATP synthase, H+ transporting, mitochondrial F0 5.5/18737 -14.79 0.0032 complex, subunit d

28 Cu/Zn superoxide dismutase 6.0/15751 2.28 0.00074

46

Spot Protein Name Theoretical Fold change T-test P

No pI/Mr ratio value

29 Odorant binding protein Ia 5.2/16759 -19.76 0.013

To summarize, the majority of spots were successfully identified - only 2 spots were unidentified due to low protein abundance. Spots 4 and 24 each contained 2 proteins which require validation to ascertain which protein is being regulated. Validation can be performed using a variety of quantitative techniques, for example, western blotting, immunocytochemistry and semi-quanitative PCR.

In an identical approach to the pulsed gels, significantly regulated spots from the continuous gels were excised, trypically digestsed and the peptides sequenced by ESI-MS/MS. These results are tabulated below (Table 2.5)-:

Table 2.5 Proteins identified from continuous experiment. For each protein the corresponding theoretical isoelectric point (pI) molecular mass (kD) & degree of up- or down-regulation (fold change) are indicated. All regulated spots were significant as determined by the Student T-test

Spot Protein Name Theoretical Fold change T-test P

No pI/Mr ratio value

1 RIKEN cDNA 5430413K10 5.5/62306 -6.1 0.019

2 RIKEN cDNA 5430413K10 5.5/62306 -5.45 0.025

NADH dehydrogenase (ubiquinone) Fe-S protein 1 5.5/79725 (Ndufs1)

3 RIKEN cDNA 5430413K10 5.5/62306 -7.95 0.02

47

4 RIKEN cDNA 5430413K10 5.5/62306 -5.97 0.02

5 hsp70 homolog cytosolic form 5.3/73469 -1.73 0.047

6 Alpha-cardiac actin 5.3/41919 1.42 0.0025

7 Actin, gamma, cytoplasmic 1 5.3/41765 1.44 0.0009

8 Albumin 1 5.7/64960 -4.92 0.0026

9 Put. beta-actin (aa 27-375) 5.8/39160 1.48 0.017

Aminoacylase 1 5.8/45766

Dhdh protein 5.9/36617

10 Carbonyl reductase 2 9.1/25941 1.50 0.02

11 Carbonyl reductase 2 9.1/25941 1.68 0.0037

12 Carbonyl reductase 2 9.1/25941 1.42 0.0059

13 PREDICTED: similar to odorant binding protein 2B 8.4/29655 -7.42 0.021

14 Cytokeratin endo A 5.7/54531 2.85 0.00056

15 Lipocalin 13 6.0/19983 -2.90 0.035

16 Major urinary protein 5 4.8/20992 1.46 0.01

Myosin regulatory light chain 2, smooth muscle isoform 4.8/19716 (Myosin RLC)

17 Lipocalin 13 6.0/19983 -6.8 0.026

18 S100 calcium binding protein A5 5.2/10805 -27.20 0.0009

48

In this case proteins from all of the significantly regulated spots were successfully identified, however, spots 2 and 16 each contained 2 proteins, and 3 proteins were identified from spot 9. Once again further validation is required in order to ascertain which protein is actually being regulated.

2.5.5 Behavioural study

Male and female P31 continuously (n=38), pulsed (n=33) and control (n=30) mice were subjected to an exploration task. In order to evaluate whether there was significant habituation towards octanal the mice were placed in an ‘open arena’ where either one vial of octanal (1:500) or one vial of cheese was hidden. The time taken to find the vial was recorded. To ensure there was no bias due to the trial order, 50% of randomized mice were firstly exposed to cheese followed by octanal, for the remaining 50% this order was reversed. Each independent trial lasted for a maximum of 10 minutes, if the mice failed to locate the vial a time of 600 sec was recorded. This data was statistically analyzed by ANOVA (Figure 2.14) which showed that there was a high degree of significance within the substance group i.e. between the times taken to identify either cheese or octanal [Factorial ANOVA F(1, 194)= 99.2496, p<0.001]. Interestingly, the pulsed and continuously treated groups found the cheese specifically quicker than the control suggesting some kind of behavioural sensitization. From Figure 2.14 it can be clearly seen that all treatment groups were ‘longer’ at finding octanal, although this was only significant between the continuously treated group and control as determined using the Fisher least significant difference (LSD) post-hoc test (p<0.05). LSD is based on the T-statistic and thus can be considered a form of T-test. It compares all possible pairs of means after the F-test (difference of group means as determined by ANOVA) rejects the null hypothesis that groups do not differ. It is evident that the vast majority of interactions were significant (Table 2.6).

If the data is transposed into either “found” (<600 sec) or “not found” (>600 sec) the distribution is markedly different (Figure 2.15). There is now a significant difference between the continuous octanal condition and all other groups, in fact, no continuously treated mouse found the octanal vial within the allocated timescale.

Notably, there was no statistical difference between sexes or importantly, between the trial order.

49

650

600 * 550

500

450

400

350

300

* Latency Time (sec) 250 ** 200

150

100 substance oct 50 substance continous pulsed control cheese Treatment

Figure 2.14 Chart showing interaction between Treatment and Substance (Least Square Means) with respect to latency (time to find substance). Current effect: F (2, 194)=9.5855, p=.00011. Vertical bars denote 0.95 confidence intervals. * = p<0.05; ** = P<0.01

Table 2.6 Below: LSD post hoc results for the above data which was analysed using factorial analysis of variance (ANOVA). Red, orange and grey boxes indicate P<0.001; p<0.01; and not significant, respectively.

Treatment Contin Pulsed Control Substance Octanal Cheese Octanal Cheese Octanal Cheese Contin Octanal 0.000000 0.090636 0.000000 0.015804 0.000003 Cheese 0.000000 0.000000 0.596768 0.000000 0.000260 Pulsed Octanal 0.090636 0.000000 0.000000 0.449821 0.002432 Cheese 0.000000 0.596768 0.000000 0.000000 0.002080 Control Octanal 0.015804 0.000000 0.449821 0.000000 0.024840 Cheese 0.000003 0.000260 0.002432 0.002080 0.024840

50

100 90 80 70 60 50 40 30

% successful finds successful % 20 10 0 Cheese Oct Cheese Oct Cheese Oct Continuous Pulsed Control Treatment Not Found Found

Figure 2.15 Score of successfully found substance against unsuccessful finds as a function of treatment regime. The latency time data was transposed into a discriminatory task – Found or Not Found within a 10 min period.

To summarise, the treated mice were quicker than the control mice at finding the food stimulus (cheese). Conversely, the treated mice were slower at locating octanal as compared to the control group, however, this was only statistically significant between the continuous and control groups.

2.5.6 EOG In order to correlate the identified proteins with a phenotype, electro-olfactogram (EOG) recordings were conducted. The EOG records field potentials from the olfactory epithelium (OE) directly reflecting the activity of responding olfactory sensory neurons. This is independent of any modulation from higher olfactory structures e.g. the olfactory bulb. The OE from control, pulsed and continuously treated mice were sequentially stimulated using Low (1:20,000), Medium (1:2,000)

51

& High (1:200) octanal concentrations. This was followed by a final application of an odorant mix, Henkel 100 (1:200), which was chosen to evoke a maximal response from the OE. Due to the variability in amplitudes between individual mice, the octanal amplitudes were normalized to the Henkel 100 response. This data was plotted and statistically analyzed (Figure 2.16). Rather unsurprisingly there was a significance difference due the concentration of octanal applied [Factorial ANOVA F(2, 42)= 2.146, p<0.001]. The medium and high octanal concentrations evoked greater responses; however, there was no significant difference between the medium and high applications suggesting that a saturation level was attained. There was also statistical significance with regard to treatment regime [Factorial ANOVA F(2, 42)= 4.091, p<0.05]. For the low octanal application there was a significant decrease in amplitude for both the continuously and pulsed treated mice with respect to the control.

* *

Figure 2.16 Chart showing EOG amplitudes in response to Low (1:20,000), Medium (1:2,000) & High (1:200) octanal concentrations. Amplitudes were normalised to a reference (application of the odorant mix Henkel 100, 1:200) and categorized according to experimental group (continuous, pulsed & control).

52

Details of any significant interactions are indicated on the following LSD Table 2.7

Table 2.7 The Fisher least significant difference (LSD) post hoc results for the above data which was analysed using factorial analysis of variance (ANOVA). Red, yellow, orange and grey boxes indicate P<0.001; p<0.01; p<0.05 and not significant, respectively. treatment Control Contin. Pulsed Conc High Med Low High Med Low High Med Low High 0.2128 0.0141 0.5937 0.9738 0.0000 0.3327 0.6367 0.0000 Control Med 0.2128 0.0004 0.0703 0.1829 0.0000 0.0266 0.0797 0.0000

Low 0.0141 0.0004 0.0382 0.0115 0.0342 0.0967 0.0332 0.0403 High 0.5937 0.0703 0.0382 0.5994 0.0002 0.6447 0.9486 0.0000 Contin. Med 0.9738 0.1829 0.0115 0.5994 0.0000 0.3262 0.6447 0.0000

Low 0.0000 0.0000 0.0342 0.0002 0.0000 0.0006 0.0001 0.6552 High 0.3327 0.0266 0.0967 0.6447 0.3262 0.0006 0.5994 0.000 Pulsed Med 0.6367 0.0797 0.0332 0.9486 0.6447 0.0001 0.5994 0.0000

Low 0.0000 0.0000 0.0403 0.0000 0.0000 0.6552 0.0002 0.0000

Thus, one can infer that at the low octanal concentration (1:20,000) both treated groups’ demonstrated octanal densensitization with respect to the control. This difference was not reflected at higher octanal applications (medium/high).

53

2.6 Discussion This investigation, to the best of our knowledge, is the first to apply a proteomics strategy to analyse odorant triggered plasticity of the olfactory epithelium. The DIGE technique proved to be a highly sensitive and dynamic method which facilitated the detection of significantly regulated proteins from pulsed and continuously octanal treated mice. Rather exceptionally, some of these proteins were regulated to a great degree. For example, the odorant binding protein Ia and the calcium binding protein S100A5 were down-regulated 16-fold and 27-fold in the in the pulsed and continuous treated mice, respectively. Such findings are rarely reported in large scale differential analysis and can only be attributed to the experimental design.

2.6.1 Mice physiology The pulsed and continuous octanal regimes not only had a significant impact on protein regulation, but also on mice phenotype, as determined using behavioural analysis and EOG recordings. From the behavioural investigation continuous octanal treatment had a dramatic effect on the mice’s ability to locate the octanal filled vial in an exploration task. This was not due to any anosmic effect as the continuous group was on average quicker at locating cheese as compared to the control. Similarly, the pulsed group was also quicker than the control at cheese discovery, however, these mice showed a greater interest in the octanal as compared to the continuous group. In terms of latency time (Figure 2.14) this was insignificant. Thus, one can conclude that: i) octanal exposure did not have an anosmic effect, in fact, it seemed to sensitise the mice in the food discovery task; and, ii) octanal latency for the continuously treated mice was significantly different from the control. This suggests a behavioural desensitisation to octanal in the continuous group, however, this was not reflected in the pulsed group (p=0.45).

In order to correlate these findings with functional changes in the olfactory epithelium we performed electro-olfactogram (EOG) recordings. It was unsurprising that higher octanal concentrations led to activation of more OSNs (perhaps more than >1012). This difference was not demonstrated between the high and medium octanal applications suggesting that a saturation level was already attained. Of greater note was the difference between treatment groups. The control mice displayed greater sensitivity to odorant treatment across all octanal concentrations; however, this was only statistically significant when a low concentration (1:20,000) was applied. Thus, there was a peripheral desensitization in both treated groups and statistically significant habituation in the

54

continuously treated mice – in light of the identified differentially regulated proteins, how may this phenotype be explained?

The proteins were therefore functionally classified and the most promising were selected for future investigation. A brief review of their function is listed below.

2.6.2 Regulation of proteins: functional categorisation & comparison The proteins listed in Tables 2.4 and 2.5 were functionally classified according to their classically interpreted role. This data is represented as a pie chart below (Figure 2.17). A total of 16 proteins were identified as candidate proteins of interest (Table 2.8). This was based on i) the extent to which the protein was up- or down-regulated and ii) a detailed investigation into their currently known roles as described in peer reviewed scientific literature.

Continuous Chaperone Homeostasis Lipocalins 1↑ 1↓ 1↑ 3↓ Energy metabolism 2↑ 1↓

1↑ Xenobiotic/ antioxidant 1↓ 5↑ Cytoskeletal Calcium binding

Pulsed Homeostasis Lipocalins

1↓ Transcription factor 1↓ 2↑

Chaperone 1↑ Cytoskeletal 6↑

Energy metabolism 1↑ 3↓

2↑ 2↑ Xenobiotic/ Calcium binding antioxidant

Figure 2.17 Functional classification of proteins identified from pulsed and continuously octanal treated mice. For each category the number of up- & down- regulated proteins are indicated by arrows.

55

Table 2.8 List of candidate proteins

Spot Fold Change Experiment Protein Name Comments No Ratio

Pulsed 2 Tumor rejection antigen 2.21

Pulsed 3 Gelsolin precursor (Actin-depolymerizing 2.20 Immunohistochemistry factor) (ADF) 11 Vimentin protein 7.87 Immunohistochemistry Pulsed 20 Annexin A5 3.06 Pulsed 23 Calbindin 2 2.33 Immunohistochemistry Pulsed 25 Hypothetical protein LOC76491 2.60 Pulsed 26 Adenylate kinase 1 -16.69 Pulsed 29 Odorant binding protein Ia -29.76 Pulsed 1 RIKEN cDNA 5430413K10 -6.10 Continuous 3 RIKEN cDNA 5430413K10 -7.95 Continuous 4 RIKEN cDNA 5430413K10 -5.97 Continuous 5 hsp70 homolog cytosolic form -1.73 Continuous 13 PREDICTED: similar to odorant binding -7.42 Continuous protein 2B 15 Lipocalin 13 -2.90 Continuous Continuous 16 Major urinary protein 5 1.46 2 proteins identified from 1 spot 18 S100 calcium binding protein A5 -27.19 Continuous

56

The greatest proportion of up-regulated proteins from the treated mice were cytoskeletal and moreover, these were usually Cytokeratin endo A and Cytokeratin endo B. Interestingly, in the continuously treated mice a greater number of lipocalins (odorant binding proteins) were regulated, the majority of these were down-regulated. Rather uniquely, two transcription factors in the pulsed experiments were up-regulated.

2.6.2.1 The cytoskeleton 2.6.2.1.1 The cytokeratins The greatest proportion of up-regulated proteins from the continuous and pulsed groups were cytoskeletal intermediate filaments. The vast majority of these were the “simple keratins”, cytokeratin endo A (K18) and cytokeratin endo B (K8), which are the major structural component of simple epithelial cells (Dewi W. Owens 2003). Epithelial cells maintain contact with the basal lamina of the extracellular matrix, throughout the body they form the lining of secretory and absorptive organs (e.g. intestine, kidney) including the pseudostratified olfactory epithelium. The intermediate filament ‘primary keratins’ K18 and K8 are highly conserved, their expression precedes all other cytoplasmic keratins and are the first intermediate filaments expressed in mouse embryonic development (Lu, Hesse et al. 2005). They form obligate heterodimers whose primary role is to provide structural stability to the epithelial cell (Yamada, Wirtz et al. 2003) and have been implicated in several other cellular processes. These include maintaining cell polarity, as potential modulators of apoptosis and protection against mechanical and non-mechanical stress (Dewi W. Owens 2003). Unlike the actin and cytofilament families, the keratins are induced in a heat shock like manner (Zhong and Omary 2004).

In the pulsed gel K8 and K18 were up-regulated in several spots and to a similar degree. This is to be expected given the fact that they are both obligate noncovalent heteropolymers. Both forms are required for filament stabilisation, over expression of either K8 or K18 leads to degradation of the overly abundant protein (Kulesh, Cecena et al. 1989). The change in mass (kDa) and isoelectric point (pI) in the differentially regulated spots indicates that both K18 and K8 were likely subjected to post-translational modifications (PTMs). It has been reported that all intermediate filament proteins may be phosphorylated and glycosylated (Ku and Omary 1994) and that the frequency of PTMs increase upon mitotic arrest. Moreover, previous work has demonstrated that increased phosphorylation (Ku and Omary 1995) may be related to filament reorganisation and that

57

cytoskeletal glycosylation can also play a regulatory role (Kearse and Hart 1991). In addition, increased phosphorylation of K8/K18 by the stress activated protein kinases, p38 and JNK, parallels up-regulation of heat shock protein 70 (hsp 70) in toxin treated hepatocytes – this has a cellular protective effect (Fausther, Villeneuve et al. 2004). Interestingly, in the continuously treated mice, hsp70 was down-regulated and only K18 was marginally up-regulated.

Thus, intermediate filament proteins may play a role in stimulus induced plasticity of the olfactory epithelium. Nevertheless, despite that fact that K8 has been identified in OSNs (Arnold, Smutzer et al. 1998), it is highly unlikely that K18 and K8 were up-regulated in the OSN, but instead, in surrounding supporting cells (Suzuki and Takeda 1991; Othman, Klueber et al. 2003) and/or in the keratin positive basal cells (Satoh and Takeuchi 1995) – K18/K8 are not mature neuron markers. This could be a consequence of increased OSN turnover; in order to verify this further immunohistochemical investigations are required.

2.6.2.1.2 Gelsolin precursor (actin-depolymerising factor) – up-regulated 2.20 fold in pulsed gel Gelsolin was first reported in 1979 when is was observed that it could activate the ‘gel-sol’ transformation of actin filaments in a calcium dependent manner (Yin and Stossel 1979). It is now known that gelsolin belongs to a large superfamily of actin binding proteins responsible for actin organization by severing filaments, capping filament ends and nucleating actin assembly (McGough, Staiger et al. 2003). Presently, there are least six addition members of this family: villin, adseverin, capG, advillin, supervillin and flightless I (Silacci, Mazzolai et al. 2004). In addition to their role in filament remodelling, these proteins have been implicated in various cellular processes including cell motility, control of apoptosis and regulation of phagocytosis. Recent evidence also suggests that gelsolin proteins may even be involved in intracellular signalling and gene expression regulation (Archer, Behm et al. 2004). For example, it is possible that Gelsolin may facilitate androgen receptor (AR) nuclear translocation and enhance AR transactivation (Hsu, Chen et al. 2005). In spite of the fact that the gelsolin knock-out (KO) mouse phenotype is quite mild, ablation of the gene results in retardation of migratory cell movement (e.g. fibroblasts). In addition, the retraction of temporary filopodia from neurite exploratory growth cones is also hindered (Stuart K. Archer 2005).

58

When the transcription factor, serum response factor (SRF), was knocked-out in the mouse fore- brain (Franziska F. Wiebel 2002), a deficiency in neuronal migration was reported (Alberti, Krause et al. 2005). As already shown that SRF regulates genes encoding for cytoskeletal proteins (Norman, Runswick et al. 1988), including gelsolin and colfilin, it was therefore unsurprising that gelsolin was down-regulated upon SRF depletion. This directly influenced actin filament recycling – polymerisation/depolymerization (Nicholson-Dykstra, Higgs et al. 2005) - and consequently, cell motility. This defect was clearly manifested in a reduction in neuron migration from the subventricular zone to the olfactory bulb along the rostral migratory stream. The authors did not report an olfactory phenotype as the animals did not usually live beyond P21 and moreover, SRF- null mice typically demonstrate an embryonic-lethal phenotype (Weinhold, Schratt et al. 2000).

Interestingly, Gelsolin has also been linked to neuronal morphogenesis in the rabbit retina (Legrand, Ferraz et al. 1991). Gelsolin localized to the inner segment of photoreceptors and not to the continually renewable outer segment. The authors also speculated that Gelsolin may play a role in the ensheathing of neurons and raises the interesting question whether it may have a parallel function in OSNs.

2.6.2.1.3 Vimentin Protein – up-regulated 7.87 fold in pulsed gel Vimentin is an intermediate filament whose primary role is to provide structural support to cells. Until relatively recently this support was presumed to be rather static, with intermediate filaments thought to act as non-motile scaffolds. Research has shown that vimentin filaments are dynamic, moving along microtubules towards and away from the cell centre (Clarke and Allan 2002). In vivo, vimentin is expressed in mesenchymal and endodermic cells in embryos, but only in mesenchyma-derived cells in adults (A. Privat 2003). During the development of the nervous system most, if not all, of the proliferating neuroepithelium of the CNS (neuronal and glial precursors) and of the migrating neural crest, contain abundant vimentin. When neuroblasts cease proliferation, vimentin expression ceases, and neurofilament proteins accumulate in the vast majority of PNS and CNS neurons. For a brief period during embryonic development, vimentin and neurofilament proteins are neuronally co-expressed; this is short lived, as vimentin is eventually down-regulated. Thus, in post-mitotic neurons, vimentin is down-regulated and neurofilament proteins accumulate. Consequently, vimentin is often considered a neuroepithelium stem cell marker which, along with other intermediate filament proteins, may play a role in OSN

59

regeneration (Chung-Liang Chien 1998). Conversely, there is also evidence that vimentin may have a role in the mature olfactory sensory neuron. It is expressed in the of mature OSNs and in some cases, even the cell bodies (Schwob, Farber et al. 1986; Kharen L. Doyle 2001).

Vimentin has been depleted from cells in tissue culture (Sarria, Panini et al. 1992) and knocked-out from the mouse genome (Colucci-Guyon, Portier et al. 1994). In both cases no strong phenotype was evident indicating that vimentin is not critical for growth, division or development, at least under normal physiological conditions. In spite of this rather benign phenotype, contrasting evidence suggests that vimentin does play a role in several cellular mechanisms, including: maintaining cellular shape through organising microfilaments and tubules; cellular migration; membrane trafficking; cell cycle (it is an immediate early gene) and metabolism of lipid derived cholesterol (Sarria, Panini et al. 1992; A. Privat 2003).

2.6.2.1.4 Do cytoskeletal proteins have non-classical functions in OSNs? Vimentin and gelsolin (Figure 2.17) were ‘functionally classified’ as ‘cytoskeletal’ proteins i.e. as their classical or conventionally understood role. As already mentioned, these proteins may be implicated in other cellular roles, for example, they may even directly effect transcription. This raises the question, why did mice which received pulsed octanal treatment up-regulate more cytoskeletal related proteins? Perhaps some of them may be involved in additional ‘non-structural’ processes? The continuously treated mice only mildly up-regulated actin and this was in fewer spots indicating fewer PTMs. These observed changes in the cytoplasmic matrix may be related to activity dependent structural plasticity of the neuron.

2.6.2.2 Intermediate early genes / transcription It has been previously shown that the immediate early genes (IEG) arc (activity-regulated cytoskeleton-associated protein) and c-fos play a role in the modulation of long-term changes in neuronal function, including those in response to a novel stimulus. Novel odour exposure to adult rats induces up-regulation of c-fos and arg in the olfactory bulb (Guthrie, Rayhanabad et al. 2000). This effect is reversed when a familiar or habituated odorant is employed (M. Montag-Sallaz 2002). Odorant up-regulation of c-fos has also been reported in the OE of three week old C57BL/6J male mice. The expression pattern also varied between the structurally different odorants octanal, trimethyl phosphate and cyclohexanone.

60

Neither c-fos or arc was detected in either of the continuously or pulsed treated mice. However, an effector protein for transcription, CCG1-interacting factor B, was up-regulated in the pulsed treated mice.

2.6.2.2.1 CCG1-interacting factor B (CIB - cell cycle arrest in G1) – up-regulated 2.60 in pulsed gel This novel protein was first identified using a yeast two-hybrid system (Padmanabhan, Kuzuhara et al. 2000) which screened for factors which interact with CCG1†. The human CCG1 gene encodes for CCG1/TAFII250/p250/TFIID250, an essential subunit to the general transcription factor, TFIID, which recognises the TATA element and therefore binds to promoter DNA. CCG1 has also been shown to play an important role in cell cycle (Hisatake, Hasegawa et al. 1993). In searching for additional novel CCG1 protein-protein interaction candidates CIB was identified and crystallised (Padmanabhan, Kuzuhara et al. 2004). It is likely that CIB is a serine esterase, the structure of which is closely related to α/β domain of prolyl aminopeptidase from Serratia marcescens‡ which catalyzes the removal of amino-terminal proline residues from peptides. In terms of transcription, it is feasible that the acetylation of histones is regulated by CIB, however, additional biochemical evidence have failed to demonstrate CIB as an acetyltransferase or deacetylase, thus, its function is still uncertain (Padmanabhan, Kuzuhara et al. 2004). Nevertheless, by virtue of the fact that CIB directly interacts with CCG1, it may be implicated in gene expression, chromatin organization, cell cycle, kinase, and ubiquitinase activities.

Undoubtedly CIB is directly involved in transcription although little is known about its function. Notably these proteins were exclusively up-regulated in the pulsed treated mice although the direct functional relevance of this is not immediately apparent.

2.6.2.2.2 A potential role for MAPK? It is known that GPCRs can also activate the extracellular signal-regulated kinase (ERK)-mitogen- activated protein kinase (MAPK) [MAPK/ERK] cascade (Gudermann 2001). These important kinases have been found in abundance in higher olfactory structures (Dorothy G. Flood 1998) and in the olfactory sensory neuron (Watt and Storm 2001). Watt and Storm were the first to suggest that long term plasticity of the OSN may be modulated by induction of the ERK/MAP kinase

† GenBankTM/EBI accession number Q96IU4. ‡ Gram negative, facultative anaerobic, motile bacillus which occurs naturally in the soil, water and intestines. Usually non-pathogenic.

61

pathway. This activation was sufficient to induce cAMP-response element (CRE)-mediated transcription in OSNs (Watt and Storm 2001). The odorant responding OSNs were rescued from apoptosis and therefore survived longer in the olfactory epithelium and moreover, this rescue was blocked upon inhibition of CREB transcription. This provided a good argument for odorant induced sensitivity which is mediated by prolonging the life of responding OSNs. In addition, there is also strong evidence that ERK1/2 and CREB phosphorylation in the olfactory bulb plays a role in aversive olfactory learning in juvenile rats (Zhang, Okutani et al. 2003). Indeed, it is now generally accepted that the MAPK pathway is of vital importance in hippocampal long-term potentiation (LTP), a potential mechanism for mammalian associative learning (Atkins, Selcher et al. 1998).

In the MAPK/ERK cascade CREB is the first step in transcription. It interacts with a DNA pre- initiation complex (PIC), this includes a class of proteins called “general activators” or the TFII’s (Holle 2004). Of notable interest, CREB has been shown to interact with TFIID, the resultant PIC facilitates positioning of RNA polymerase II, initiating transcription (Felinski and Quinn 1999). More recently it has been suggested that TFIID may have a role in neuronal differentiation (Brunkhorst, Karlen et al. 2005), in fact, neuroblastoma cells contain high levels of endogenous TAF4, a sub-unit of TFIID (Brunkhorst, Neuman et al. 2004). As mentioned above, CIB, a novel protein shown to interact with TFIID, was up-regulated in the pulsed gel. This raises interesting questions - was this in response to CREB phosphorylation? Could this facilitate long term adaptation of the olfactory sensory neuron? This would contrast its previous suggested role in olfactory sensitization. Moreover, of central importance, how does this relate to the intracellular calcium concentration?

2.6.2.3 Calcium binding proteins Calcium concentration is important in any biological cell and this is especially true for excitatory cells. Odorant activation of the OSN results in an elevation in intracellular Ca2+ and is associated with depolarization of the cell. Elevated intracellular calcium is facilitated by a Ca2+ influx via CNG channels and from intracellular stores, the latter may also lead to long term effects such as adaptation (Gomez, Lischka et al. 2005). Two ‘classical’ calcium binding proteins were up- regulated in the pulsed experiments (annexin A5, calbindin 2) and one was significantly down- regulated in the continuously treated mice (S100 calcium binding protein A5). In addition, one

62

cannot ignore the up-regulation of gp96 in the pulsed experiments, this classical chaperone may also play a role in calcium binding and buffering.

2.6.2.3.1 Annexin A5 – up-regulated 3.06 fold in pulsed gel Annexins are a universally conserved superfamily (27 subfamilies) of calcium binding proteins containing a distinctive four-repeat architecture, the so called ‘annexin repeats’ or type II binding sites (Brachvogel, Dikschas et al. 2003). They are involved in a multitude of intercellular and extracellular functions including apoptosis (Reutelingsperger, Dumont et al. 2002; Kietselaer, Hofstra et al. 2003), homoeostasis and membrane scaffolding and trafficking (Hayes, Rescher et al. 2004; Gerke, Creutz et al. 2005). In particular, annexin A5 has been shown to physically insert into the membrane and directly interact with phosphatidylethanolamine and also bind to phosphatidylcholine. In this Ca2+ dependent process, A5 assembles as a 2D ‘crystal’ (ordered array) on the membrane (Richter, Lai Kee Him et al. 2005). The function of this lattice is unknown; however, it may be involved in export and import mechanisms (e.g. ‘blebbing’) and signal transduction. It has even been hypothesised that annexin A5 could sufficiently perturb the membrane in order to electroporate the membrane and facilitate Ca2+ influx (Demange, Voges et al. 1994). From an evolutionary perspective annexins were considered to originate from a common Ca2+ ion channel (Morgan, Martin-Almedina et al. 2004) and thus can conduct calcium across the membrane. This theory is controversial and there is a lack of evidence. It is perhaps more reasonable to suggest that certain annexins (esp. A7) mediate intracellular Ca2+ release (Baker, Witherdin et al. 2005). There is strong evidence for the regulation of ion channels by annexins A2, A4 and A6, although to date, A5 has not been implicated. Of particular interest is annexin A2 which interacts with another calcium binding protein, S100A10. There are several reports of this complex interacting with NaV1.8 Na+ channel, the two-pore acid-sensitive K+ channel-1, the transient receptor potential-V5 (TRPV5) and TRPV6 epithelial Ca2+ channels (Gerke, Creutz et al. 2005). Finally, several annexins have been shown to interact with the cytoskeleton. Annexin 1 binds to actin filament bundles perhaps ‘directing’ polymerisation and, coupled with their ability to bind to the inner membrane envelop, regulate the membrane-cytoskeleton (Hayes, Rescher et al. 2004). For example, it is possible that annexin 6 could regulate microfilament architecture in neurites in response to calcium signals.

63

In recent years the use of in vitro systems has greatly improved our understanding of annexin function. Annexin A5 still remains one the least poorly understood and, to compound matters, A5- null mice apparently have no clear behavioural or biochemical defects (Brachvogel, Dikschas et al. 2003). Nevertheless, depending on the intercellular location of A5, it is apparently involved in a plethora of mechanisms ranging from nuclear translocation (interaction with helicase and demethylase), membrane/vesicle trafficking, exocytosis and embryogenesis (Knowlton, Chan et al. 2003).

2.6.2.3.2 Calbindin 2 or Calretinin (CR) - up-regulated 2.33 fold in pulsed gel Calbindin 2 or calretinin (CR) is a calcium binding protein which belongs to the calbindin family. It shares 25% homology to calmodulin and possesses 6 EF-hand domains – a common calcium binding motif in animal cells (Nelson, Thulin et al. 2002). In neuronal cells Calbindin 2, and other EF-hand proteins (incl. calbindin D-28k (CB) and parvalbumin (PV)), are classically considered to be calcium buffers (Billing-Marczak and Kuznicki 1999). More recently their role as a calcium modulator has also been suggested. Their potential modulatory role is accentuated in the cerebellum where, CB, CB-2 and PV are enriched. Here, calbindin-2 is principally expressed in the granule cells, whereas PV and CB are located in the axons, soma and dendrites of Purkinje cells (Schwaller, Meyer et al. 2002). This differential expression is thought to regulate calcium pools which are important in synaptic plasticity. In immunohistochemistry investigations of the murine olfactory system, calbindin-2 was identified in the olfactory sensory neuron, and in the vomeronasal and septal chemoreceptor neurons (Kishimoto, Keverne et al. 1993). The vast majority of PV and CB immunostaining was restricted to the VNO. In the rat, distinctive and localised calbindin and calbindin-2 staining was reported in bundles, with staining being restricted to external fibres (Bastianelli and Pochet 1994). Building on this work, Bastianelli et al also examined the expression pattern of six calcium binding proteins in post-natal developing rat olfactory epithelium. Calmodulin, calretinin, calbindin-D28k and neurocalcin were localised to the mature sensory neuron and not to basal cells. This expression pattern sequentially intensified from P1 to P20 whereupon it attained a maximal level equivalent to that seen in the adult. A similar increase in ontological expression was also reported in the rainbow trout (Porteros, Arevalo et al. 1997). The authors suggested that this degree of differential calcium-binding organization may play a role in odour discrimination.

64

2.6.2.3.3 S100 calcium binding protein A5 (S100A5) – down-regulated 27.10 fold in continous gel S100 is a large multigene family of proteins which contain two EF-hand domains and bind two Ca2+ / Cu2+ ions per molecule and one Zn2+ ion per molecule. There is general consensus that S100 monomers dimerise in a calcium (or divalent cation) dependent process, with the resultant dimer mediating protein-protein interactions. The only known exception to this hypothesis is the monomeric calbindin D9K isoform. S100 calcium binding proteins are expressed in a tissue specific manner and are implicated in an array of cellular processes including phosphorylation, differentiation, modulation of the cytoskeleton and enzymes, inflammation, transcription and calcium homeostasis (Rosario Donato 2003). With regard to the adult nervous system, the S100 proteins are known to play a role in embryonic and post-natal development (Chan, Xia et al. 2003).

Of a total of 18 members comprising the S100 family, S100A5 is the least reported in the literature. Relatively recently there has been a moderate increase in investigations. S100A5 has restricted expression patterns in the kidney (Teratani, Watanabe et al. 2002) and brain, with prominent immunohistochemistry labelling only occurring in the olfactory bulb, the brainstem, and the spinal trigeminal tract in adult Harlan Sprague-Dawley rats (Schafer, Fritschy et al. 2000). The most intense olfactory bulb staining was in the olfactory nerves and in a subset of glomeruli. In addition, S100A5 was not found in any glial cells or in brain neuronal soma. In an elegant paper, Schafter et al purified and biochemically characterised this protein demonstrating that the S100A5 dimer binds Zn2+ in addition to Ca2+ and, rather uncommonly, also binds Cu2+. The authors did not refer to the olfactory epithelium. Subsequent work reported that S100A5 is important in embryonic development of the hippocampus and temporal cortex (Chan, Xia et al. 2003), and that it is postnatally expressed in a subpopulation of neurons in dog cochlea (spiral ganglion) (Coppens, Kiss et al. 2001).

More recently, another investigation differentially compared the transcriptome of olfactory sensory neurons with neighbouring cells and identified 54 differentially expressed transcripts (Tun-Tzu Yu 2005). From these transcripts S100A5 was selected as a candidate for further investigation, its expression was restricted to a subset of OSNs in what was described as “a mosaic expression in about 70% of OSNs” (Tun-Tzu Yu 2005). The authors refrained from discussing any functional implication this may have. Nevertheless, one could speculate that the additional copper sequestering ability may enable the protein to act as an enzyme co-factor, or perhaps even protect

65

the OSN against oxidative damage. Moreover, it has been previously suggested that the olfactory receptor protein may even be a metalloprotein (Wang, Luthey-Schulten et al. 2003).

2.6.2.3.4 The significance of calcium binding proteins in the OSN Considering the plethora of roles calcium binding proteins may have in the olfactory sensory neuron and besides that, in the entire olfactory epithelium, drawing conclusions can only be conjecture. One can accurately infer that at least two of the identified regulated calcium binding proteins (S100A5 & calretinin) were regulated in the mature olfactory sensory neuron. The chaperone and calcium binding protein, gp96, may have been up-regulated in non-neuronal cells and is discussed later. The importance of intracellular calcium concentration in the OSN cannot be understated.

It is generally accepted that odorant triggered elevation of cAMP (AC3) is the primary, although not exclusive, olfactory signal transduction pathway. Cross-talk between cAMP and other pathways is an important means of olfactory plasticity e.g. mediating long term potentiation (LTP) and memory (Wang and Storm 2003) not to mention receptor desensitization/sensitisation. With regard to short-term receptor adaptation, a series of related events have been elucidated – central to 2+ these is [Ca ]i. Upon OSN depolarization, AC3 is phosphorylated by Calmodulin (CAM) dependent protein kinase II (CAM-Kinase II), inactivating the enzyme. This slow adaptation is also strengthened by concomitant processes involving two kinds of phosphodiesterases (PDE) - CAM- PDE, which has a low affinity for cAMP, and cAMP-PDE which can hydrolyze both cAMP and cGMP. Both PDEs hydrolyse cAMP, although cAMP-PDE is not dependent on CAM : Ca2+. This results in reduced cAMP which also contributes to cellular adaptation (Bradley, Reisert et al. 2005). The phosphodiesterases are ‘kept in check’ through phosphorylation by (PKA). PKA, which is activated by odorant triggered elevation in intracellular calcium, also phosphorylates the odorant GPCR leading to its desensitization. This effect is potentiated by additional receptor phosphorylation by G-protein coupled receptor kinase III, GRK3, in an arrestin 2 dependent manner (Mashukova et al, submitted). In addition, the sensitivity of the CNG channel is also regulated by phosphorylation and by Ca2+: CAM levels i.e. elevated Ca2+ actually reduces its sensitivity to cAMP, thereby reducing the channels release probability (Zufall and Hatt 1991; Schleicher, Boekhoff et al. 1993). These rather complex regulatory processes are thought to mediate short-term receptor adaptation (seconds), but what of longer term effects?

66

Leinders-Zufall & Zufall proposed that long lasting adaption (LLA) of the OSN is mediated by cGMP (Zufall and Leinders-Zufall 1997). The OSN CNG channel has a higher affinity for cGMP

(Nakamura 2000) than cAMP – the K1/2s for cAMP and cGMP are ca. 3 and ca. 2 μM, respectively – thus, lower concentrations of cGMP can illicit a long term gating of the channel. This results in long lasting (6 min in salamander OSNs) independent adaptation of the neuron. It was proposed that cGMP is generated by nitric oxide (which incidentally may also directly gate the channel) (Schmachtenberg, Diaz et al. 2003), or by carbon monoxide (CO) production in the neuron (Zufall and Leinders-Zufall 1997). Furthermore, evidence also exists for involvement of IP3 as an alternative or perhaps more accurately, complementary secondary messenger. OR activation of membrane bound PLC could lead to hydrolysis of PIP2 and production of IP3 and 2DG. IP3 can directly open intracellular calcium stores potentiating long lasting adaptation. Evidence for these mechanisms is weak but nonetheless, cannot be disregarded.

Why was calretinin and annexin A5 up-regulated in the pulsed treated mice, whereas, S100A5 was down-regulated in the continuously treat mice?

Calretinin has a relatively fast Ca2+-binding kinetic which has been shown to reduce transmitter release probability (Felmy and Schneggenburger 2004). As previously mentioned, CR is expressed pre-synaptically in glutamatergic OSNs. One can therefore speculate that an elevation would reduce postsynaptic excitation (↓ excitatory postsynaptic currents - EPSCs), having an effect on short term facilitation. Membrane bound annexin A5 may also ameliorate this affect, although it is more likely it has a neuron-protective role in preventing apoptosis. There is evidently a ‘thin line’ between its role as promoter or impeder of apoptosis. Annexin V has been shown as a marker for neuronal apoptosis in the rat brain (Walton, Sirimanne et al. 1997) and in respiratory sensory neurons, where apoptosis was induced by sustained calcium influx through activated TRPV1 channels (Agopyan, Head et al. 2004). Nevertheless, annexin A5 has also been shown to have a neurotrophic effect (Takei, Ohsawa et al. 1994), including protection against oxidative damage, on cultured cortical neurons (Han, Zhang et al. 2004).

However, in the continuous experiments, the only regulated calcium binding protein, S100A5, was significantly down-regulated (-27.20 fold). There are reports of up- and down- regulation of S100 proteins in transformed cells (Feng, Xu et al. 2001; Suzuki, Lareyre et al. 2004) where down-

67

regulation may protect the cell against apoptosis. This protection may also be facilitated through interaction with Annexins (Roesch Ely, Nees et al. 2005). One can merely speculate that down- regulation in the olfactory sensory neuron may prevent apoptosis. Its mosaic expression in the olfactory epithelium must be of significance, it may even be related to odorant conditioning. Further investigations are required in order to unravel its role in the OE.

2.6.2.4 Chaperones Chaperones are ubiquitous proteins that are often localized to the endoplasmic reticulum. Here they govern protein folding which culminates in the export of the synthesized protein. In the case of olfactory receptors, little is known about these mechanisms, a fact which is exacerbated by the notorious difficulty to functionally express ORs in heterologous expression systems (Lu, Echeverri et al. 2003). In order to improve heterologous expression we have recently reported that the coexpression of a testicular heat shock protein, hsc70t, significantly improves transfection rates (Neuhaus, et al; submitted).

One ‘classical’ chaperone was regulated in both the pulsed and continuously treated groups, tumor rejection antigen gp96 and hsp70. They are briefly discussed below.

2.6.2.4.1 Tumor rejection antigen gp96 – up-regulated 2.21fold in pulsed gel Tumor rejection antigen gp96 has several synonyms, namely, Endoplasmin [Precursor], Endoplasmic reticulum protein 99 (ERP99), 94 kDa glucose-regulated protein (GRP94) and Polymorphic tumor rejection antigen 1. It is a single copy gene (Tra-1§), and is found in some multicellular organisms such as C. elegans, plants and in all vertebrates, however, it is absent from drosophila and from unicellular organisms like bacteria and yeast (Argon and Simen 1999). Its importance in vertebrate development is apparent given the fact that mouse knock-outs (KO) are lethal at the embryonic stage of mesoderm formation (Dollins, Immormino et al. 2005).

It is a stress or ‘heat shock protein’ (although is not significantly induced by heat shock) whose primary role is proposed to be folding and chaperoning of nascent polypeptides (Nicchitta, Carrick et al. 2004). Having said this, unlike many other classical chaperon proteins, gp96 physically interacts with comparatively few proteins – these include; Ig chains, MHC class II, thyroglobulin,

§ UniProtKB/Swiss-Prot entry P08113

68

erbB2 (epidermal growth factor receptor), a herpes virus glycoprotein, apolipoprotein B, collagen and protein C (Argon and Simen 1999). It is a member of the HSP 90 family and is specifically located in the endoplasmic reticulum, accounting for 5-10% of the lumenal contents. Here, in the endoplasmic reticulum, gp96 plays an integral role in calcium buffering having a low-affinity/high- capacity binding for calcium (Argon and Simen 1999).

Previous reports have demonstrated that up-regulation is induced by low glucose levels, prolonged anaerobiosis, acidosis, expression of mutated or transfected proteins and viral infections. More recently it has been suggested that gp96 plays a role in immunity (antigen presentation) and even in establishing pathogen virulence (Descoteaux, Avila et al. 2002). There are no reports with regard to the olfactory system, nevertheless, recently gp96 has been implicated in glial cell (Gye Sun Jeon 2004) and neuronal cell (Bando, Katayama et al. 2003) protection in response to stress.

This protein was up-regulated in the pulsed treated mice. Naturally without relevant follow-up investigations the implications of this cannot be inferred. Nevertheless, one may speculate that up- regulation of gp96 may have: i) a neuro-protective effect in response to stress and/or apoptosis ii) a role signal transduction through Ca2+ buffering iii) a role in receptor chaperoning.

2.6.2.4.2 Hsp70 homolog cytosolic form (Mortalin) – down-regulated 1.73 in continuous gel Hsp 70 has several other synonyms including 75 kDa glucose regulated protein (GRP 75), Peptide- binding protein 74 (PBP74), P66 MOT and Mortalin. It is a member of the heat shock protein 70 family and is differentially expressed in normal and cancerous cells. Using various biochemical techniques (northern/western blots, in situ hybridisation, immunohistochemistry) Kaul et al reported that mortalin is differentially expressed in various rat tissues with the highest levels being detected in skeletal muscle and brain (Kaul, Matsui et al. 1997). Intracellularly, it may be found in mitochondria, endoplasmic reticulum, cytoplasmic vesicles and the cytosol (Kaul, Taira et al. 2002). Its functions range from stress response, muscle activity, mitochondrial biogenesis, trafficking, control of cell proliferation, differentiation, cell fate determination and tumorigenesis. With regard to neurons, mortalin has been suggested to have neuro-protective effects and is also up-regulated in injured dorsal root ganglion (DRG) axons (Willis, Li et al. 2005). However, in the continuous treated experiment hsp 70 was down-regulated. A similar down-regulatory effect ( ≈ -4.3 fold) was also reported in a proteomic study into the effect of heat stress and the NMDA channel blocker,

69

MK801, on primary cortical neuronal cultures (Bruno P. Meloni 2005). Rather surprisingly, hsp 70 & 90 were not induced during heat stress, in fact, hsp 70 was down-regulated. MK801 has been previously shown to illicit a neuro-protective effect when administered to cultures prior to assault (Tremblay, Chakravarthy et al. 2000); in this case it induced down-regulation of hsp 70 (≈ -7.0) along with several other cytoskeletal, mitochondrial and chaperone proteins. Feasibly hsp 70 down-regulation may also have a neuro-protective effect on the olfactory epithelium.

2.6.2.5 Lipocalins The lipocalins are a diverse family of proteins which bind to small hydrophobic molecules such as fatty acids, steroids, odorants, and pheromones and can directly interact with cell surface receptors (Flower 1996). Selective ligand binding is facilitated by a distinctive β-barrel structure (a ‘cup’) which contains a loop at one end - the molecular ‘entrance’. This permits the specific entry and binding of retinoids, steroids, fatty acids and a range of aromatic and aliphatic compounds which include odorants (Schlehuber and Skerra 2005).

All but one lipocalin was down-regulated in the pulsed and continuously treated mice. The exception, major urinary protein 5, was slightly up-regulated (1.46) in the continuous gel. Each of these potential ligand binding proteins (lipocalin and/or lipocalin like) are discussed below.

2.6.2.5.1 Odorant binding protein Ia – down-regulated 29.76 fold in pulsed gel Odorant binding proteins (OBPs) are also members of the lipocalin family consisting of >100 proteins which transport physiological hydrophobic molecules including vitamins, steroids and various metabolites (Lacazette, Gachon et al. 2000; Lobel, Strotmann et al. 2001). They are also found in tear and salivary secretions. As the name suggests, OBPs are located in the nasal epithelium and selectively bind odorants. They are produced by the lateral and septal glands of the nasal epithelium and bind hydrophobic odorants which otherwise could not diffuse across the aqueous mucus environment and bind to the olfactory receptor (Park, Shanbhag et al. 2000). Conversely, it has also been suggested that instead of chaperoning odorant molecules, OBPs may actually bind to and inactivate dominate odours which would otherwise desensitise the OSN (Lacazette, Gachon et al. 2000). Thus, the exact role of odorant binding proteins still remains elusive.

70

Despite reports that the overall sequence homology within and between species is low (Pes, Mameli et al. 1998), there is striking homology between some rat and mice forms (Tegoni, Pelosi et al. 2000). Murine OBP Ia is related to its homologue OBP II, which shows 80% to the rat form OBP I (Ohno, Kawasaki et al. 1996). In rat, OBP expression increases markedly from P1 till a maximum level around P30, in fact, the expression appears to parallel ontological development of the (Hamil, Liu et al. 2003). It is likely that in mice olfactory development OBP expression is similar to the rat, however, no observations have been published.

2.6.2.5.2 Lipocalin 13 (Lcn-13) – down-regulated 2.90 & 6.8 fold in continuous gel Lcn-13 is secreted into the epididymis from around P21 with a maximal expression level at adulthood i.e. approx. week 7 or 8. Lcn-13 production ceases upon castration and cannot be recovered using testosterone supplements. The possible involvement of epididymal secreted lipocalins in sperm development has already been suggested, where they could transport ligands to and from lumen epithelial cells or perhaps even to the developing spermatozoa (Thompson, Gibson et al. 1997). To date Lcn-13 has not been described in the olfactory system.

2.6.2.5.3 RIKEN cDNA 5430413K10 – down-regulated in 3 spots; 6.1, 7.95 & 5.97 fold The role of this protein is unknown and consequently a Basic Local Alignment Search Tool (BLAST) search was performed**. The closest homologue was rat potential ligand binding protein (RYF3)†† with an E-score (expected score) of e-121 i.e. a low probability that this significant match has arisen by chance. To illustrate, the BLAST search was re-aligned using CLUSTAL_X (Dear, Boehm et al. 1991), the windows interface of the common multiple alignment tool CLUSTALW. The rat potential ligand binding protein was first cloned in 1995 from olfactory mucosa (Mechref, Ma et al. 1999). The authors report that no homology was found between RYF3 and known proteins, however, a current BLAST reveals that it has over 50% homology with rat vomeromodulin – a possible pheromone and odorant transporter (Novotny, Ma et al. 1999). It therefore appears likely that RIKEN cDNA 5430413K10 is a potential odorant binding protein which was identified in the pulsed, continuous and control gels as three isoforms. These isoforms were only down-regulated in the continuous gels and displayed a discernable change in mass and pI

** EXPASY blastp server: http://www.expasy.org/tools/blast/. Standard search settings across all species was employed: Comparison matrix - auto-select; E threshold - 10; Filter for low complexity sequences and Gapped alignment enabled. †† Accession Number : Q05703

71

(in results section, area of interest A). In order to verify this, the spots were excised and identified from two separate experiments using both ESI and MALDI. Using phosphorylation prediction software (Netphos 2.0‡‡) it seems likely that the shift in pI arose from phosphorylation. The prediction results are indicated in Figure 2.19 below.

Figure 2.18 Alignment of identified unknown candidate protein with rat potential ligand binding protein.

Unfortunately, due to lack of time the phosphorylation sites were not mapped. Phosphorylation can also, to a lesser degree, account for the proteins shift in mass. Nevertheless, it is probable that RIKEN cDNA 5430413K10 was also subjected to additional post-translational modifications such as N- or O-glycosylation.

‡‡ http://www.cbs.dtu.dk/services/NetPhos/

72

Figure 2.19 Upper chart: probability of residue phosphorylation (y- axis) as a function of amino acid sequence (x-axis). Lower table: postulated change in molecular weight and pI as a consequence of phosphorylation.

2.6.2.5.4 Predicted similar to odorant binding protein – down-regulated 7.42 fold in continuous gel This protein shows significant homology to the rat odorant binding protein 2b (OBP2b or RY2G12). The functional relevance of this class of proteins has already been discussed above.

The above presumed odorant binding proteins were markedly down-regulated in the pulsed and continuously treated mice. As the exact role of these proteins are currently unknown, and without supporting biochemical and anatomical data, discussing the reason for this striking change can be only speculative. Nevertheless, taking the data at ‘face value’, a down-regulation in OBP presumably indicates a reduction in the binding capacity for relatively hydrophobic odorants. As a consequence, it is envisioned that olfactory sensitivity for these compounds is diminished.

73

In the behavioural and electrophysiological recordings there was an apparent desensitization to octanal – could this be a reflection of diminished lipocalin expression? On the other hand, arguably this down-regulation could also be attributed to a saturation of the lipocalins due to octanal binding – i.e. the lipocalins source was exhausted or not replenished quickly enough.

Interestingly, one lipocalin was up-regulated, major urinary protein 5.

2.6.2.5.5 Major urinary protein 5 (MUP 5) – up-regulated 1.46 fold in continuous gel Major urinary proteins are also lipocalins whose function is to bind to pheromones which are secreted in male urine. The mouse genome possesses around 35 MUPs, several of which have been shown to illicit strong behavioural responses (in conjunction with the VNO system) such as puberty (Timm, Baker et al. 2001) and male aggression (Cavaggioni and Mucignat-Caretta 2000). The type and amount of excreted MUP varies between mice strain and even between individuals (Cavaggioni and Mucignat-Caretta 2000; Novotny 2003). Moreover, their affinity to odorant molecules is quite biased - MUPs display a preferential affinity for the pheromones 2-sec-butyl-4,5- dihydrothiazole (DHT) and dehydro-exo-brevicomin (DHB) (Makio Utsumi 1999). Strong evidence also demonstrated that these ‘pheromones’ (DHT & DHB) also induce aggression in male mice.

Interestingly they are also synthesised in other tissues including the liver, mammary glands and main olfactory epithelium (Makio Utsumi 1999). MUPs 4 and 5 are apparently selectively expressed in the nose, in particular, the lateral nasal glands (L1 and L3 regions) and nasal glands of the septum (Leinders-Zufall, Lane et al. 2000). Nasal MUPs may therefore have a dual role: in the male murine nose eliciting aggression; in the female nose, a priming pheromone response accelerating or effecting sexual responses.

It has already been shown that mouse VNO receptors which respond to these ligands do so with a high affinity, and a narrow tuning range (Novotny, Ma et al. 1999). It is therefore feasible that their corresponding ligand binding protein, MUP 5, may also possess a relatively narrow tuning range. In the continuous experiment MUP 5 was the only up-regulated lipocalin. Given its affinity for the pheromones DHT (thiazole) and DHB (heterocyclic compound), and considering their structural dissimilarity with octanal (Figure 2.20), it is therefore likely that MUP 5 cannot strongly bind to octanal. This raises the question: why was this [presumed] non-octanal binding protein up-

74

regulated? As these mice (P31) are adolescences’ approaching sexual maturity, one could speculate that an up-regulation of a pheromone binding protein may accelerate puberty (Chen and Vandenbergh 1993). This acceleration should be accompanied by a concomitant increase in the uterine and total body weight (Tun-Tzu Yu 2005), this was not observed and therefore the reason for this moderate up-regulation is not immediately apparent. Furthermore, it is possible that MUP5 was up-regulated by the vomeronasal organ (VNO) as the olfactory epithelium preparation probably contained tissue from the VNO.

Figure 2.20 Comparison of MUP 5 binding pheromones with octanal

To conclude, the down-regulation of nasal lipocalins may consequently reduce the sensitivity of the peripheral olfactory epithelium or VNO to the relatively hydrophobic aldehyde, octanal. This could in part explain the phenotypic desensitization in the continuously and pulsed treated mice.

75

2.6.2.6 Xenobiotic & anti-oxidant metabolism In the pulsed treated mice carbonyl reductase 2 (CR2) and Cu/Zn superoxide dismutase were up- regulated and are presumably responsible for xenobiotic and anti-oxidant metabolism, respectively. Carbonyl reductase 2 was also up-regulated in the continuously treated mice.

In the olfactory epithelium CR2 is only expressed in the sustentacular cells and is known to have a role in the clearance of odorants which, in effect, are xenobiotic compounds (Bohren, von Wartburg et al. 1987). It has been isolated in various tissues including the liver, lung and brain and has relatively broad substrate specificity. It is an aldo-keto reductase catalyzing the transfer of the pro-4S-hydrogen atom of the nicotinamide ring of NADPH to the carbonyl substrate (Ghosh, Sawicki et al. 2001). Metabolically, carbonyl reductase, along with other short chain dehydrogenases/reductases, catalyse critical steps in prostaglandin and leukotriene metabolism. This essentially involves the activation (-OH oxidation) and inactivation (C=O reduction) of steroids, vitamins, prostaglandins, and other bioactive molecules (Oritani, Deyashiki et al. 1992). Furthermore, porcine lung isolated CR2 was also shown to reduce a range of aliphatic and aromatic compounds including octanal, the proposed reaction is shown in Figure 2.21 (Ling, Gu et al. 2004).

It is proposed that the more polar product, octanol, is subsequently metabolized by the well reported microsomal cytochrome P450 (CYP) enzymes of the olfactory mucosa (Ling, Gu et al. 2004). Notably, none of these abundant enzymes were evidently regulated which suggests that the administered concentration of octanal was non-toxic. Decreased expression of cytochrome P450 enzymes, in particular CYP2A/2G, is associated with pathological or toxicological conditions (Forrest and Gonzalez 2000).

More recently is has also been suggested that CR2 may also play a role in free radical modulation i.e. in the clearance of reactive oxygen species from the cell (Fujii, Myint et al. 1995). Currently there is little evidence to suggest this, however, Cu/Zn superoxide dismutase (SOD) was also up- regulated in the pulsed treated mice. Superoxide dismutases scavenge superoxide radicals protecting cells against oxidative stress (Janssen, Dzeja et al. 2000).

Nevertheless, it seems highly probable that up-regulated CR2 was responsible for xenobiotic metabolism of octanal in the olfactory epithelium.

76

Figure 2.21 Proposed xenobiotic clearance of octanal by CR2 in the sustentacular cells

2.6.2.7 Energy metabolism Some enzymes involved in mitochondrial energy metabolism were up-regulated nevertheless, there was a notable down-regulation in the pulsed treated mice. This included adenylate kinase 1 (AK1) which was significantly down-regulated and subsequently highlighted as a candidate protein of interest.

2.6.2.7.1 Adenylate kinase 1(AK1) – down-regulated 16.69 fold in pulsed gel Adenylate kinase catalyses the reaction ATP + AMP ↔ 2ADP and is obviously a key enzyme in the synthesis, equilibration and regulation of adenine nucleotides. It is predominately located in skeletal and brain playing an integral role in intercellular energy transfer, and has even been implicated in phosphoryl transfer from mitochondria to myofibrils (Inouye, Seo et al. 1998). Thus,

77

one would expect this enzyme to mirror metabolic demand. Previous work comparing the expression level of AK1 during neuronal differentiation in mouse embryonic carcinoma P19 cells, and in rat brain primary cultured cells, demonstrated a fivefold up-regulation of AK1 in differentiated neurons (Inouye, Seo et al. 1998). This up-regulation was specific for neuron like cells, and not to glial cells suggesting an important role in neuronal development. Subsequent work compared expression levels in the developing rat brain where AK1 was localised in the olfactory bulb and olfactory nerve layer (Yue Ge 2003). AK1 expression increased rapidly during early postnatal days with expression lagging in the cerebellum. In the entire brain (western blot) levels increased from E18 to P29 in accordance with energy demand. The exceptionally high expression level in the bulb and evidently, in the olfactory sensory neurons, could be attributed to continual turnover of OSNs. However, in the pulsed experiment, there was a significant down-regulation which contradicts these findings. It seems unlikely that energy demand waned as the olfactory system was developing. Previously a fivefold down-regulation of AK1 has been reported in the hindlimb muscles of mdx mice (a model for severe Duchenne muscular dystrophy (DMD)) using large scale 2D-electrophoresis (Piast, Kustrzeba-Wojcicka et al. 2005). It was suggested that this deficiency was compensated by increased creatine kinase activity (alternative phosphotransfer system) although this data was not published from the proteomic dataset. It is therefore feasible that OSNs utilised alternative metabolic pathways e.g. β oxidation of free fatty acids which could also, in theory, metabolise octanal.

2.6.3 Consolidating the findings: Protein to Phenotype To recap, the pulsed and continuously treated mice displayed peripheral desensitization to octanal with the continuously treated mice also exhibiting habituation. Despite observations that the pulsed mice also demonstrated habituation, this was not statistically significant. Desensitization/habituation was not due to anosmia as both treated groups displayed what appeared to be sensitization towards the food stimulus.

Behavioural habituation, a form of implicit memory, is more strongly modulated by higher brain regions. Indeed, it is probable that this also occurred with both treated groups. There is strong evidence that cortical adaptation is faster and more selective than seen in either OSNs or mitral cells (Wilson 2000). Nevertheless, from the EOG data it is incontestable that desensitization occurred at the level of the OE.

78

This interesting outcome was unexpected: It was hypothesized that the pulsed treated mice would be sensitized to octanal whereas the continuously treated mice would be habituated. Evidently long term adaptation occurred in both groups. Given the differences in protein regulation, the underlying molecular mechanisms must be different.

2.6.3.1 Protein regulation in the continuously treated mice Desensitization in the continuously treated mice may have been influenced by down-regulation of lipocalins. Octanal is relatively hydrophobic and therefore cannot easily diffuse across the mucosal layer without binding to a suitable carrier protein. In addition, there is evidence that odorant binding proteins have odorant selectivity, thus regulation of particular sub-types could have a filtering effect (Breer 2003). The continuously treated mice down-regulated more lipocalins than the pulsed treated group. This included a novel candidate ligand binding protein, RIKEN cDNA 5430413K10, which was also subjected to post-translational modifications (PTMs). In addition, we have also reported the down-regulation of lipocalin 13, which until now was only described in the reproductive system.

Of particular note was the dramatic down-regulation of S100A5, a calcium binding protein that is normally expressed in around 70% of mature olfactory sensory neurons. Its current role is unknown, however, from limited reports one could postulate that its down-regulation may play a role in neuro-protection. This argument is further strengthened by the evident down-regulation of hsp70 and moderate up-regulation of actin, which have also been implicated in neuro-protection (Bruno P. Meloni 2005). The exact mechanisms of this proposed response is unknown; however, it seems likely that these proteins make promising candidates for future investigation.

2.6.3.2 Protein regulation in the pulsed treated mice Evidently lipocalin down-regulation was not the principle mechanism of desensitization in the pulsed treated mice. Indeed, protein regulation in these mice was more dynamic, involving more proteins with a greater number of PTMs (e.g. the cytokeratins). In addition to the ‘simple cytokeratins’, vimentin and gelsolin were also up-regulated. This may be related to an up-regulation in OSN production from basal cells; nevertheless, one cannot ignore the possible ‘non-classical roles’ of these proteins which even include transcription. Moreover, a recently identified transcription factor binding protein, CIB, was also up-regulated.

79

Up-regulation of the calcium binding protein calretinin may have influenced the synaptic release probability which, in all likelihood, is also coupled with cytoskeletal rearrangements. The calcium binding proteins annexin A5 and gp96 were also up-regulated, the former may have a role in preventing apoptosis and it is very likely that latter played an integral role in intracellular calcium buffering. It is known that gp96 is up-regulated upon endoplasmic calcium store depletion (Yang and Li 2005). Furthermore, one cannot exclude that gp96 may also act as an olfactory receptor chaperone. Interestingly, the number of known client proteins for gp96 is significantly lower than its homologue chaperone, HSP90. The list of currently known client proteins include MHC class II, insulin-like growth factor–II, p43, ErbB2 and integrins (Yang and Li 2005). We therefore suggest that gp96 is a very promising candidate for future investigations into receptor trafficking and its potential role in calcium modulation.

2.6.4 Conclusion To conclude, inferences can be drawn from the regulated proteins which in part explain peripheral desensitization in the continuously and pulsed octanal treated mice. For the continuously treated mice we suggest that the greatest single factor affecting desensitization was lipocalin regulation. A clear dominant factor influencing the pulsed group cannot be easily identified.

However, connecting these peripherally regulated proteins with behavioural sensitization to the food stimulus is less obvious. To date there is little evidence to indicate that odorant specific desensitization leads to a general increase in responsiveness to non-familiar odorants. In fact, application of a dominate odorant for 20 min pulses to rats for six consecutive days led to a decrease in responsiveness of the olfactory bulb, not only to the familiar odorant, but also to unfamiliar odorants (Buonviso, Gervais et al. 1998). This so called ‘self-adaptation’ (Dalton 2000) did not occur with the treated mice. Nevertheless, odorant exposure may indeed lead to a general sensitization as previously described in humans who work in ‘olfactory challenging’ environments, such as wineries (Bende and Nordin 1997) or perfumeries (Hummel, Guel et al. 2004). This could be mediated by increased peripheral sensitivity to novel odorants, or more likely, is due to plasticity in higher brain loci including the olfactory bulb.

Thus, in the treated mice it is likely that there was adaptation, including cross-adaptation to octanal in the olfactory bulb. One can envisage that those M/T neurons within octanal’s receptive field

80

were laterally inhibited by GABAergic granule cells (GC). This could be possibly mirrored with an inhibitory decrease in spike frequency of the (Davis 2004) or indeed, in other tertiary neurons such as the , the perirhinal cortex and the amygdala. Notably, after long exposure to octanal it is also feasible that the M/T cell receptive field shifted from octanal towards other non-optimal odorants which have different carbon chain lengths (Fletcher and Wilson 2003). This could possibly improve odorant discrimination and in part explain the enhanced detection of the food stimulus. The food stimulus, a rather pungent cheese, is clearly a complex and rich odour source which may include octanal analogues and undoubtedly includes hydrophilic and structurally unrelated compounds. The treated mice olfactory bulb, and to a lesser extent their OE, may have been tuned to non-octanal aldehydes with different carbon lengths. Moreover, hydrophilic and structurally unrelated compounds would not be susceptible to a deficit in odorant binding proteins – the result? – increased sensitivity to the food stimulus.

81

3 Proteomic investigation of the human sperm membrane 3.1 Rationale – a correlation between chemosenses and reproduction? Excitable cells are conventionally considered to be muscle-, nerve- and sensory transducer cells. These cells are equipped with specialized sensors and voltage-sensitive channels which enable membrane potential modulation in order to transduce signals.

Spermatozoa have never been classically considered as ‘excitable cells’. One need only consider the morphological differences between a spermatozon and neuron to illustrate the common belief that both cells have very little in common. Of even greater note are the intracellular differences: neurons are diploid transcriptionally active cells whereas haploid spermatozoa do not even possess golgi apparatus nor endoplasmic reticulum (Meizel 2004). Nevertheless, accumulating evidence suggests that despite these salient dissimilarities, spermatozoa and neurons possess similar membrane receptors and ion channels – indeed, spermatozoa have been facetiously described as “neurons with a tail” . Several ‘neuronal’ receptors have been identified in human sperm including those for neurotransmitters and neuromodulators e.g. ionotropic AMPA and NMDA receptors (Shimshek, Bus et al. 2005), metabotropic GABA receptors (He, Hu et al. 2001) and cannabinoid receptors (Schuel and Burkman 2005). In addition, an array of ion channels have been found in mature spermatozoa which have been implicated in the acrosome reaction (AR) and in chemotaxis (Darszon, Labarca et al. 1999). Many of these channels are conserved throughout species ranging from marine life to mammals.

It has been previously shown that olfactory receptors are transcribed in mammalian testis (Parmentier, Libert et al. 1992) and expressed in mature canine sperm (Vanderhaeghen, Schurmans et al. 1993). This suggests a role for chemoreception in fertility, most likely with respect to chemotaxis and chemokinesis. This argument was strengthened by pioneering work which functionally characterised a human testicular olfactory receptor, hOR17-4, on human sperm (Spehr, Gisselmann et al. 2003). The hOR17-4 agonist, bourgeonal, not only activated sperm in calcium imaging studies, but also acted as a strong chemo-attractant in behavioural investigations. Within the field of reproductive biology these findings encourage debate regarding which molecular mechanism transduces the chemotactic signal. Two signal tranduction pathways have been proposed: i) a pathway initiated with a guanyl cyclase receptor (Kaupp, Solzin et al. 2003) or ii) a pathway initiated with an olfactory receptor (Spehr, Schwane et al. 2004). It is possible that sperm

82

contains both systems (Eisenbach 2004) although support for the olfactory mechanism is increasing. With regard to neuroscience, these findings suggest that neurons have more in common with spermatozoa than a mere ‘excitatory’ phenotype. Could olfactory sensory neurons (OSNs) and sperm cells possess similar receptors?

The purpose of this study was to proteomically profile human sperm membrane proteins in order to identify chemoreceptors to assist in the elucidation of their function in fertility. Conventional techniques such as in-situ hybridization or PCR could not be employed as mature spermatozoa are transcripitionally inactive.

3.2 The challenge of membrane proteomics The membrane proteome may be described as those proteins which are strongly associated with or integral to the membrane. In eukaryotic cells this also includes intracellular organelles. The importance of resolving the membrane proteome is axiomatic – membranes serve more than a mere means for compartmentalization, they facilitate many fundamental functions, including signal transduction, cell-cell interaction and transportation. Moreover, it has been estimated that around 20-30% of the human genome is represented by membrane proteins of which up to 70% are currently targeted by drugs (Kroeze, Sheffler et al. 2003). It is speculated that one of the most important classes, G-protein coupled receptors, account for around 50% of marketed drugs (Santoni, Molloy et al. 2000; Wu and Yates 2003). Thus, there have been many efforts to analyse membrane proteins, nonetheless, their proportion has been grossly under-represented in proteomic datasets. This is mainly due to the amphipathic nature of transmembrane proteins (hydrophilic and hydrophobic regions) and the fact that they are normally in low abundance. Consequently, the proteomic technique of choice must possess a large dynamic range with high sensitivity whilst facilitating reproducible resolution and solubilisation. Such a methodological ‘holy grail’ has not yet been described although several developments have been made with regard to sample fractionation and the identification of membrane proteins using mass spectrometry.

It is well recognised that traditional IEF/2D-electrophoresis possesses inherent limitations in identifying extremely acidic, basic and hydrophobic proteins (Gorg, Obermaier et al. 1997). With regard to hydrophobicity, the reason for this shortcoming is primarily two-fold: firstly, the isoelectric focusing buffer cannot contain ‘strong’ ionic detergents, a prerequisite for solubilisation;

83

and secondly, solubilised membrane proteins often precipitate at their isoelectric point (pI). Furthermore, most hydrophobic proteins tend to have pIs in the alkaline range, but due to electroendoosmotic movement, isoelectric focusing above pH 8 significantly deteriorates. In some cases membrane proteins can be identified by 2D-electrophoresis (Molloy 2000), however, invariably these proteins do not possess multiple transmembrane regions (Bunai and Yamane 2005). Consequently, two alternative strategies have evolved. Non-IEF gel-based methodologies and ‘in-solution’ shotgun approaches e.g. MudPIT.

3.2.1 In-gel methods Successful gel-based techniques are either based on 1D-SDS PAGE (Laemmli 1970) or 2D- electrophoresis employing ‘stronger’ ionic detergents in the first dimension e.g. the 2D-BAC/SDS- PAGE (2-DB) system (Hartinger, Stenius et al. 1996).

1D-SDS PAGE is a well known and routinely used technique worldwide. Quaternary, tertiary and secondary protein structures are denatured in laemmli buffer which contains a ‘strong’ ionic detergent, SDS. Unfolded protein chains are surrounded by negatively charged SDS molecules which form elongated micelles. To ensure adequate denaturisation it is recommended that the sample is heated and a reducing agent is added, for example, β-mercaptoethanol [HS-CH2-CH2- OH]. This reduces disulphide bonds [-S-S-] and therefore protein disulphide bridges. The sample is subsequently loaded into an SDS-PAGE gel, the distance of protein migration is proportional to the molecular weight (Mathews 1990). It is likely that the resolved protein bands contain several proteins, these are excised and digested in a likewise manner as described for 2D-electrophoresis. However, due to the sample complexity the use of matrix-assisted laser desorption mass spectrometry (MALDI-MS) is not recommended.

2-DB, on the other hand, employs two dimensions and therefore facilitates better resolution, although this is significantly lower than with traditional IEF. Proteins in the first dimension are separated according to molecular weight in an acidic discontinuous PAGE system (pH 4.0–1.5) using cationic benzyldimethyl-n-hexadecylammonium chloride (BAC) as detergent (Macfarlane 1989; Hartinger, Stenius et al. 1996). As proteins in the first BAC-PAGE dimension show slightly different migration properties compared to that seen in conventional SDS-PAGE, the resolution can be improved by a subsequent second-dimension SDS-PAGE (René-Peiman Zahedi 2005).

84

Using this technique the resolution of extremely hydrophobic proteins has been reported including those with up to 10 transmembrane domains (Isam Rais 2004). More recently, a double SDS- PAGE (Dirk Claeys 2005) has also been described which may also be combined with blue native gel electrophoresis (Schagger and von Jagow 1991) in order to identify membrane protein complexes. These methods are undoubtedly powerful, nevertheless, they are still restricted by a limited dynamic range (staining) and moreover, in reality, relatively few highly hydrophobic multi- transmembrane proteins (e.g. GPCRs) have been actually identified.

3.2.2 In-Solution shotgun approach Using ‘shotgun’ methods complex protein mixtures (e.g. cell lysate) are enzymatically digested and directly analysed using liquid chromatography/mass spectrometry (LC/MS). MudPIT (introduced before) is a powerful example of this technology. As with gel systems, the major hurdle is protein solubilisation – three main strategies are currently applied, solubilisation via: organic solvents; organic acids and detergents. Naturally these conditions must be compatible with enzymatic digestion and mass spectrometry, thus extremely denaturing conditions or high detergent concentrations are prohibited. The classical MudPIT protocol used a combination of organic acid (formic acid) and chemical cleavage (cyanide bromide) in order to solubilise and identify saccharomyces cerevisiae membrane proteins (Washburn, Wolters et al. 2001). A total of 131 membrane proteins with three or more predicted transmembrane regions were identified. This large scale technique was refined for total brain membrane proteins which not only facilitated concomitant analysis of membrane and soluble proteins, but also their post-translational modifications (Wu, MacCoss et al. 2003). Soluble and peripheral membrane proteins were removed by washing a crude membrane preparation in a high pH solution (100mM sodium carbonate). High pH (11.5) induces the formation of membrane ‘sheets’ or fragments which allow improved digestion of integral membrane proteins using the non-specific protease, proteinase K. A total of 1,610 brain proteins were identified, 28.2% of which were predicted membrane proteins. Blonder also employed a stringent carbonate washing step on bacterial (Deinococcus radiodurans) membrane proteins, however, digestion was achieved using trypsin in the presence of 60% methanol (organic solvent) (Blonder, Goshe et al. 2002). 503 proteins were identified, 135 were calculated as hydrophobic using grand average of hydropathicity (GRAVY) values for proteins (Kyte and Doolittle 1982). In subsequent work, detergent-resistant membrane rafts from African green monkey kidney (Vero) cells were analysed using identical digestion conditions, revealing a

85

total of 380 proteins - 24% were classified as integral α-helical membrane proteins, of which 56% were predicted to have multiple transmembrane domains (Goshe, Blonder et al. 2003; Blonder, Hale et al. 2004).

‘In-solution’ digestions with detergents have also been reported. Han et al compared the microsomal fractions from undifferentiated and differentiated (induced by 12-phorbol 13-myristate acetate) human HL-60 cells (Han, Eng et al. 2001). Microsome fractions were dissolved in 0.5% SDS and labelled using the commercially available sulfhydryl-specific ICAT (isotope-coded affinity tags) reagent. ICAT is analogous to DIGE in that it is also a quantitative proteomic technique, facilitating differential analysis of samples. Whereas DIGE is gel based, ICAT relies on MS for quantification. The latest generation of ICAT reagents (Applied Biosystems) have been improved by the addition of a biotin cleavable tag which enriches for labelled proteins. Using this combination of detergent and ICAT labelling Han et al. identified 491 proteins determining their relative abundance in the microsomal fractions of native and in vitro−differentiated human HL-60 cells. Despite the author’s enthusiasm for ICAT technology, it has significant limitations including limited accuracy and proteome coverage.

More recently (Nan Li 2004), as much as 1% SDS was used to solubilise and facilitate tryptic digestion of detergent resistant membrane microdomains (RAFTs) in a monocyte cell line (THP- 1). This ‘in-solution’ strategy was compared to the conventional ‘in-gel’ 1D-SDS PAGE approach. 71 lipid raft proteins were revealed of which 45 were not detected using in-gel digestion. Nevertheless, employing detergent for ‘in-solution’ digestions will undoubtedly retard enzyme digestion and moreover, can cause significant problems for subsequent mass spectrometry, indeed additional sample ‘clean-up’ steps are essential.

3.2.3 Fractionation Most of cited ‘in-solution’ methodologies enzymatically digested relatively crude membrane pellets which were stringently washed in order to remove contaminating proteins. Using a gel based method the proteins are also fractionated by their molecular weight and/or isoelectric point. However, using an in-solution technique without protein pre-fractionation produces a complex peptide mixture; this may contain contaminants or hinder identification of low abundant proteins. Strategies have therefore been developed in order to enrich for cellular structures, organelles and

86

low abundant proteins (Yates III, Gilchrist et al. 2005). In a land-mark paper, silica-coated-beads were used to isolate lung endothelial cell surface proteins from normal and transformed tissue. MudPIT & 1D-SDS were employed in combination with a bioinformatic and a subtractive proteomic method (Andersen, Wilkinson et al. 2003) resulting in the identification of Annexin A1 as a potential new target for tumour therapy (Oh, Li et al. 2004). As an alternative means for isolation of the plasma membrane proteome, cell surface biotinylation has been suggested. N- Hydroxysulfosuccinimide (Sulfo-NHS) esters of biotin are the most popular type of biotinylation reagent. NHS-activated biotins react efficiently with primary amine groups (-e.g. NH2 groups from lysine side chain) in pH 7-9 buffers to form stable amide bonds. Theoretically, only extracellularly exposed proteins are labelled – so called “vectorial labelling” (Vener and Stralfors 2005). Labelled surface proteins may be affinity enriched using avidin beads, this procedure has recently gained in popularity and has been applied to several cell lines (Zhang, Zhou et al. 2003; Zhao, Zhang et al. 2004) although to date, this technique has not demonstrated a convincing advantage over ‘cruder’ membrane preparations.

3.3 Sperm proteomics: a background As mentioned before, the nuclear chromatin in mature mammalian sperm is highly condensed and therefore transcriptionally inactive i.e. PCR screening of sperm cDNA could not be employed. Thus, in order to identify novel proteins in the mature cell, only proteomic methodologies may be utilised.

Several investigations have employed non-fluorescent difference gel electrophoresis to compare fertile sperm with samples from infertile patients in an attempt to elucidate whether failed fertilization events are a consequence of aberrant protein expression (Pixton, Deeks et al. 2004; Rajeev and Reddy 2004). The investigations focussed on the soluble cell lysate. Shetty, on the other hand, resolved surface sperm proteins using 2D-electrophoresis and transferred the gels to nitrocellulose membranes i.e. performed western blots (Shetty, Naaby-Hansen et al. 1999). The blots were incubated with clarified serum from 15 infertile male subjects (sperm auto-antibodies) and 6 infertile female subjects (sperm iso-antibodies) that contained antisperm antibodies (ASA). 98 sperm auto- and iso-antigenic protein spots were recognized; however, protein identification was not reported. Employing a similar strategy Shibahara and colleagues identified ‘sperm immobilizing antigens’ which may be considered as the ideal targets in developing immunocontraceptives

87

(Shibahara, Sato et al. 2002). This work was established on notable work which vectorially labelled human sperm in order to enrich for surface proteins (Naaby-Hansen, Flickinger et al. 1997). Sperm surface proteins were labelled using biotin or 125I and “surface proteins” were detected by western blots and autoradiography, respectively. For protein separation, isoelectric focusing (IEF) and nonequilibrium pH gradient electrophoresis (NEPHGE) in the first dimension was coupled with polyacrylamide electrophoresis in the second dimension i.e. IEF/PAGE, NEPHGE/PAGE, respectively. This strategy identified 1397 spots of which 228 were surface biotinylated and 181 were 125I labelled – these proteins were presumably membrane.

A novel human sperm membrane antigen, SAMP32 (sperm acrosomal membrane-associated protein 32) was discovered from biotinylated human sperm proteins isolated by Triton X-114 phase partitioning and resolved by 2D-electrophoresis (Hao, Wolkowicz et al. 2002). Electron microscopy indicated that SAMP32 is localized to the equatorial segment in the inner acrosomal membrane of capacitated sperm, thus, surface biotinylation was either a result of spontaneous acrosomal reaction or cell damage.

Recently, a novel epididymis glycoprotein was identified in Lewis rats by means of 2-D gel electrophoresis, which was followed by immunoblotting with rat hyperimmune sera raised against isologous rat sperm (Rao, Herr et al. 2003). It was suggested that this epididymal secreated protein (E-3) may have a role in sperm maturation, sperm-egg binding, or as a defensin to protect sperm from bacterial infection.

Bohring also employed a 2D-electrophoresis and immunoblotting strategy in order to identify ASA from ‘highly enriched’ human sperm membrane proteins (Bohring and Krause 1999; Bohring, Krause et al. 2001). Several proteins including HSP70 and HSP70-2 were identified as potential ASA. The majority of candidate proteins were not membrane associated, a predicable outcome given the fact that sperm cells and membrane fraction were rather crudely prepared using the swim- up procedure (Edwards, Bavister et al. 1969) followed by sonication and ultracentrifugation. As inadequate washing steps and ‘standard’ IEF-2D was employed it is likely that the majority of resolved spots were non-membrane. Having said this, this strategy was successful at identifying potential antisperm antibodies which are directly relevant to fertilization, opening up opportunities for diagnosis and treatment of immune infertility (Domagala and Kurpisz 2004).

88

There are older reports describing techniques whereby one may ‘dissect’ and differentially analyse regions of the sperm cell body e.g. acrosomal region, head and tail. These include ‘chemical dissection’ of mammalian sperm using various strategies, such as manipulating buffer osmolarity, proteolytic cleavage and the use of reducing agents (Millette, Spear et al. 1973; Gall, Millette et al. 1975; Meur, Dhoble et al. 1988). Using a combination of mechanical dissociation and density gradient centrifugation, guinea pig sperm acrosomes were enriched (A. Stojanoff 1987). None of these rather outdated physicochemical investigations employed 2D-electrophoresis or any other means of large-scale proteomics. Fractions were analysed by microscopy and biochemical assays in order to evaluate structural integrity and purity.

More recently, there have been efforts to analyse the phosphoproteome of sperm. Bailey investigated the total cell lysate from capacitated Boar sperm using a combination of immunoprecipitation and western blot analysis employing phosphotyrosine antibodies (Bailey, Tardif et al. 2005). Using western blots transferred from 1D and 2D polyacrylimide gels at least nine tyrosine phosphoproteins were identified, four of which varied with capacitation. A similar study was also conducted in human sperm which clearly demonstrated that capacitation is associated with cholesterol release from the sperm membrane and the activation of protein kinase A and tyrosine kinase secondary messengers (Osheroff, Visconti et al. 1999). An elegant approach was employed by Ficarro who investigated the phosphoproteome of capacitated human sperm (Ficarro, Chertihin et al. 2003). A 2D gel electrophoresis/western blot approach was used and once again antibodies against phosphotyrosine were employed. However, unlike many previous studies an additional novel step was utilised, namely, IMAC - immobilized metal affinity chromatography (Ficarro, McCleland et al. 2002). Using this technique Ficarro successfully identified 5 tyrosine, 3 threonine and 56 Serine phosphorylation sites. Reassuringly some of these candidate proteins were already known (e.g. a kinase-anchoring proteins) and importantly new candidate proteins were also identified including valosin-containing protein (VCP), a homologue of the SNARE interacting protein NSF.

3.4 Conclusion There is a clear lack of investigations into the human sperm membrane proteome, moreover, any large scale studies have employed 2D-electrophoresis as the preferred technology of choice. The power and reproducibility of this technique is self-evident in examining soluble lysates (e.g. first half

89

of this manuscript); this is not the case for membrane proteomics. Acknowledging the challenge of membrane proteomics, and in particular, the human sperm membrane (Martinez and Morros 1996), we set about to empirically optimise the necessary conditions to identify olfactory receptors in human sperm.

3.5 Materials and Methods 3.5.1 LC/LC-MS/MS and Protein Identification LC-ESI MS/MS and Protein Identification. A fused-silica microcapillary column was laserpulled (Model P-2000, Sutter Instrument Co., Novato, CA) and successively packed with 10 cm of 5-μm C18 reverse-phase material (XDBC18, Hewlett Packard) and 4 cm of 5-μm strong cation exchange material (SCX-Partisphere, Whatman, Clifton, NJ) as previously described (Gatlin, Kleemann et al. 1998). The sample was bomb-loaded and analyzed using fully automated online LC/LC-MS/MS instrumentation (HPLC LC Packings Ultimate, MS, Finnigan LTQ). A Dual Gradient System HPLC pump (Dionex, Amsterdam) including a Famos auto sampler and Switchos was connected to a Finnigan LTQ ion trap mass spectrometer (Thermo Electron Corp., San Jose, CA). The LTQ was operated via Instrument Method files of Xcalibur to acquire a full MS scan between 350 and 2000 m/z followed by full MS/MS scans of the three most intensive ions from the preceding MS scan. The heated desolvation capillary was set to 180°C. The relative collision energy for collision induced dissociation was set to 35%, dynamic exclusion was enabled with a repeat count of 2, a repeat duration of 0.5 min, and a 3 min exclusion duration window. Samples were loaded onto a 15 cm fused silica column as described elsewhere (Wolters, Washburn et al. 2001). The column flow rate was set to 0.15-0.25 µL/min and a spray voltage of 1.8 kV was used. The buffer solutions used for the chromatography were 5% ACN, 0.1% Formic Acid (buffer A), 80% ACN, 0.1% Formic Acid (buffer B); 500 mM Ammonium Acetate, 5% ACN, 0.1% Formic Acid (buffer C). Either a 6- step or 12-step salt gradient (usually chosen to resolve complex sample mixtures) was performed.

The SEQUEST algorithm was used to interpret MS/MS spectra. Results were interpreted on the basis of a conservative criteria set, i.e., only results with DCn scores greater than 0.08 were accepted, cross-correlation scores (Xcorr) of single charged, double charged or triple charged ions had to be greater than 1.8, 2.5, or 3.5 respectively. Spectra for proteins identified by less than three peptides were manually evaluated to match the following criteria: Distinct peaks with signals clearly

90

above noise levels, differences of fragment ion masses in the mass range of amino acids, and fulfilment of consecutive b and y ion series.

3.5.2 Sperm preparation Human sperm were either freshly obtained from young healthy donors or obtained from a fertility clinic as 50 ml frozen aliquots. For fresh sperm preparations a PercollTM density gradient centrifugation was performed after liquefaction (30 min at 35.5°C) to isolate mature and motile sperm (Moohan and Lindsay 1995). In brief, liquefied semen was overlaid on a two-layer Percoll (Amersham Biosciences) density gradient consisting of 80% and 55% isotonic PercollTM in Ham’s F-10 medium (Invitrogen). After 40-min centrifugation at 500 X g at room temperature the pellet was collected, washed in standard Ringer solution (140 mM NaCl, 5 mM KCl, 2 mM CaCl2, 2 mM

MgCl2, 10 mM Hepes, 10 mM glucose), and again centrifuged for 15 min. For frozen sperm preparation, samples stored at -70°C were thawed on ice and PercollTM purified as described above with the exception that all steps were performed at 4°C. The frozen or fresh sperm pellets were immediately used for sample optimisation strategies.

3.5.3 Ca2+ imaging Fresh PercollTM purified sperm cells from a single donor were prepared as described above (sperm preparation, 3.5.2). Cells were incubated for 30 min at 37°C in the loading buffer containing 3 mM Fura2 (Molecular Probes) and standard Ringer’s solution. After 30 min cells were rinsed once with Fura2 free Ringer’s solution and then incubated in Fura free Ringer’s solution. The final concentration of Papaverine (Sigma) and Strychnine (Sigma) were 500µM and 1mM, respectively. The positive control, Progesterone (Sigma), was applied at a concentration of 30μM. Ca2+ imaging experiments were performed on the stage of a Zeiss inverted microscope equipped for ratiometric imaging with a xenon arc lamp, a multi-wavelength illumination system POLYCHROME II for excitation, a charge-coupled device (CCD) camera, and WinNT based T.I.L.L.-Vision software to collect and quantitate spatiotemporal Ca2+-dependent fluorescence signals (f340/f380 ratio). Cells were viewed with 32 X magnification.

3.5.4 Immunocytochemistry The following primary antibodies were used: (a) rabbit polyclonal antibody against NCAM (b) rabbit polyclonal antibody against GluR1 (c) mouse monoclonal antibody against Syntaxin 3 [kind

91

gifts from Prof. Dr. A. Faissner)] and (d) rabbit polyclonal antibody against P2X3 (Alomone Laboratories, Jerusalem, Israel). For immunofluorescence secondary goat anti-rabbit or goat anti- mouse conjugated to Alexa 488 or Alexa 568 from Molecular Probes were used. Enriched spermatozoa were plated on polylysine-coated coverslips (15 min/room temperature). Coverslips were washed in Ringer solution and fixed (30 min/room temperature) in 3% paraformaldehyde in Ringer solution containing 10 mM glucose. After washing with PBS, cells were permeabilized and non-specific binding sites blocked with 0.1% Triton X-100 in PBS containing 1% cold water fish skin gelatin (Sigma) (overnight/4oC). Cells were incubated with primary antibodies diluted 1:100 in PBS/gelatin/Triton X-100 for 1.5 hr, washed, and incubated with fluorescently labelled secondary antibodies diluted 1:500 in PBS/gelatin/Triton X-100 for 45 min; nuclei were stained with DAPI (Molecular Probes) in this incubation step. Unbound antibodies were washed with PBS, and coverslips were mounted in ProLong Antifade (Molecular Probes). All fluorescence images were obtained with a confocal microscope (LSM510 Meta; Zeiss) using a 40 X 1.4-numerical aperture objective (pinhole set to one Airey unit) and further processed with Photoshop (Adobe Systems Inc., San Jose, CA).

3.5.5 Sample optimisation - main strategies Assortments of methods were attempted, the details of which are beyond the scope of this manuscript and thus, only the main strategies are listed below. These were based on either a gel based approach, or on an in-solution strategy.

3.5.5.1 Gel Based Approach 3.5.5.1.1 1D-SDS PAGE ‚In-Gel’ digestion Spermatozoa were enriched as described above and resuspended in standard Ringer’s solution with protease inhibitors (Roche Complete® protease inhibitor mixture). Cells were lysed by homogenization (Ultraturax; 1200 units/min for 1 min) and subsequent sonification for 10 s at 40 watts (Sonifier B12, Branson Sonic Power Co.). Cell debris and nuclei were removed by centrifugation (1,000 X g, 10 min), the pellet was discarded, and the supernatant was centrifuged for 2 h at 100,000 X g. The resultant membrane pellet was solubilised using 1% CHAPSO (Sigma). A sample aliquot (100 μg total protein) in Laemmli buffer (30% glycerol, 3% SDS, 125 mM Tris/Cl, pH 6.8) was resolved by 8% SDS-PAGE. Coomassie-stained bands were excised, pooled, and washed thrice in 100 mM NH4HCO3 (pH 8.5) in 50% acetonitrile (ACN) for 20 min.

92

Colourless gel pieces were washed in 100% ACN for 10 min and dried by vacuum centrifugation for 15 min at room temperature. Gel pieces were reduced with 20 mM dithiothreitol (DTT) in 100 mM NH4HCO3, 5% acetonitrile for 1 h at 55 °C, washed (100 mM NH4HCO3 for 10 min), dehydrated (100% ACN 20 min), and alkylated (100 mM iodoacetamide/100 mM NH4HCO3 for 30 min (dark/room temperature). After further washing, dehydration, and drying, the pieces were rehydrated in digestion buffer (40 mM NH4HCO3/9% ACN containing proteomics grade trypsin (Promega) at an enzyme:substrate ratio of 1:50) and incubated overnight at 37°C. Peptides were extracted using 1% trifluoroacetic acid in 50% acetonitrile for 20 min. The gel pieces were vortexed for 5 min before recovering the supernatant. The extracted peptides were concentrated by vacuum centrifugation and subjected to LC/LC-MS/MS.

3.5.5.1.2 1D-SDS PAGE with Additional Detergents The above protocol (3.5.5.1.1) was also performed with the exception that CHAPSO was substituted with the following detergents (1%): N-Lauroylsarcosine, Sodium taurocholate, Triton X-114, Triton X-100, Nonylphenyl Polyethylene Glycol NP-40, 1% (w/v) L-α- lysophosphatidylcholine and ASB-14. MudPIT was not performed in each case; instead, each detergent condition was directly compared to an olfactory epithelia membrane preparation which was also solubilised in parallel. Both samples were loaded unto one mini-gel. Identical initial protein concentrations were used, this ranged from 80μg to 200μg total protein.

3.5.5.1.3 Olfactory epithelium preparation for detergent comparison Adult CD1 mice were obtained from the animal facility, Ruhr University Bochum. They were sacrificed by CO2 asphyxiation. The OE was dissected as previously described (Spehr, Wetzel et al. 2002). Briefly, the septal bone with the intact olfactory epithelium was dissected from the head and the epithelium was dissected from the septal bone and directly placed into 100μl ice-cooled dissection buffer in a 1.5ml eppendorf. The buffer consisted of ringer’s [NaCl 140mM, KCl 5mM,

MgCl2 1mM, CaCl2 2mM, 10mM Hepes, 10mM glucose] with Roche Complete® protease inhibitor cocktail. The OE membrane pellet was prepared in an identical fashion to the sperm membrane pellet as described above (1D-SDS PAGE ‚In-Gel’ digestion, 3.5.5.1.1), the choice of detergent paralleled the sperm experiment (3.5.5.1.2).

93

3.5.5.2 In Solution Strategy 3.5.5.2.1 Organic Acid Solubilisation: Formic acid Cell lysis buffer was standard ringers solution (140 mM NaCl, 5 mM KCl, 2 mM CaCl2, 2 mM MgCl2, 10 mM Hepes, 10 mM glucose) containing protease inhibitors (Roche Complete® protease inhibitor mixture). Cells were lysed by homogenization (Ultraturax; 1200 units/min for 1 min) and subsequent sonification for 10 s at 40 watts (Sonifier B12, Branson Sonic Power Co.). Cell debris and nuclei were removed by centrifugation (1,000 X g, 10 min), the pellet was discarded, and the supernatant was centrifuged for 2 h at 100,000 X g. Essentially this crude membrane pellet was solubilised and digested as previously described (Washburn, Wolters et al. 2001). Briefly, 50–100 μL of a CNBr solution (500mg/mL in 96% formic acid - Sigma) was directly added to the pellet and incubated at RT, in the dark, overnight. The solution was neutralized with ice-chilled 30% ammonium hydroxide and carefully adjusted to pH 8–8.5 using ammonium bicarbonate

(NH4HCO3- Sigma). The sample was solubulized using 8M urea, reduced (20 mM DTT for 15 min, RT), alkylated (25 mM iodoacetamide, incubated in the dark at RT for 20 min) and subsequently digested using endoproteinase Lys C (1:100 enzyme:sample – Roche Diagnostics) at 37oC for a minimum for 8 hrs in the dark. The digest was diluted to 2M urea, 50mM ammonium bicarbonate, pH 8.5, to which 100 mM CaCl2 to a final concentration of 1 mM was added. The sample for digested further using sequencing grade trypsin (1:100 enzyme:sample - Promega) with shaking (600 rpm) at 37°C (Eppendorf thermomixer) for a minimum of 8 h in the dark. The resultant complex peptide mix was centrifuged (Eppendorf 5451R) at 20,000 X g for 2 min to remove particulates. Failure to do so blocks the column. The sample was acidified using 0.5% acetic or 0.5% formic acid (pH<3). LC/LC-MS/MS was performed as described above.

3.5.5.2.1.1 PROTEINASE K & STRINGENT WASHING This protocol was based on previous work (Wu, MacCoss et al. 2003) with minor modifications. Briefly, slow thawed sperm were PercollTM purified, washed, resuspended in cell lysis buffer and homogenised as described in strategy above (organic acid solubilisation, 3.5.5.2.1). Crude membrane pellets were suspended in high pH solution (100mM sodium carbonate, pH 11.5), vortexed and sonicated using an ice cooled sonication bath (Bandelin, Sonorex TX 20) for 15 min. The sample was allowed to rest on ice for 1hr before adding urea to 8M and reduced and alkylated as described above (organic acid). 5 μg Proteinase K (Sigma) was subsequently added to the

94

suspension and incubated at 37oC for 3 hr at 600rpm in a thermomixer (Eppendorf). Another 5 μg aliquot was added for an additional 1.5hr, the reaction was quenched with formic acid to a final concentration of 5% and microcentrifuged at 18,000 X g for 15 min; 4oC to remove particulates. The sample was loaded and analysed by MudPIT as described above.

3.5.5.2.2 Organic solvent Solubilisation: Methanol (Methanol incl. stringent washing & multi-enzyme digestion) The membrane pellet was obtained as described for the organic acid preparation and washed using ice cold 100 mM sodium carbonate (pH 11.5) followed by re-ultracentrifugation at 100,000 X g for 30 min. The pellet was rewashed in high salt solution (1.5M sodium chloride) and ultracentrifuged at 100,000 X g for 30 min. The pellet was resuspended in digestion buffer (60% methanol in 50 mM ambic) containing trypsin and chymotrypsin (1:100 enzyme:sample). Digestion occurred overnight with shaking (500rpm) at 370C. The sample was briefly spun (micro-centrifuge), acidified to

3.5.5.2.2.1 STRINGENT WASHING & ADDITION OF LIPASE The membrane pellet was obtained as described for the organic acid preparation. The pellet was washed using ice cold 100 mM sodium carbonate (pH 11.5) and re-ultracentrifuged at 100,000g X g for 30min. The pellet was rewashed in high salt solution (1.5M sodium chloride) and ultracentrifuged at 100,000 X g for 30 min. Following this, the pellet was resuspended in digestion buffer (60% methanol in 50mM ambic) containing trypsin, chymotrypsin (1:100 enzyme:sample) and non-specific lipase (2:100 enzyme:sample, Novo Nordisk). Digestion occurred overnight with shaking (500rpm) at 370C. The sample was briefly spun (micro-centrifuge) and acidified to

3.5.5.2.2.2 AFFINITY PARTITION Stock solutions of 20% Dextran DX T 500 (w/w) and 40% PEG 3350 (w/w) were prepared as previously reported. For the affinity ligand coupling step, Dextran T500 was activated by tresyl chloride as described before (Flanagan and Barondes 1975). The biphasic system (6% Dextran, 6% o PEG, 20mM H3BO4, 15mM·TrisHCL pH 7.8) was equilibrated overnight at pH 7.4 / 4 C before commencing extraction. The washed fresh sperm Percoll purified pellet was homogenised in 5 ml

95

of the uncoupled biphasic system (PEG 2 : Dextran 1) using a pestle and tube and the subsequent suspension was phase-separated by centrifugation at 800 X g for 15 minutes. The PEG upper- phase containing membrane proteins was removed and the lower Dextran phase was re-extracted twice with fresh upper phase. The combined PEG fractions were subjected to another round of extraction except in this case using lectin (WGA) coupled Dextran T500. The lower phase containing glycosylated hydrophobic proteins was diluted in 0.1 M N-acetylglucosamine, 0.25 M Saccharose, 15 mM Tris pH 7.4 and ultracentifuged (100,000 X g for 80 min). The resultant pellet was washed in 2.5 M sodium bromide and re-ultracentrifuged as before. This affinity purified and washed pellet was digested using 2μl lipase, chymotrypsin and trypsin as indicated above. MudPIT was subsequently performed.

3.5.5.2.2.3 VECTORIAL LABELLING WITH BIOTIN Only fresh PercollTM purified sperm cells were surface biotinylated by suspending the cells in PBS containing 3 mg/ml of 100 mM EZ-Link Sulfo-NHS-SS-Biotin stock solution (Pierce), to a final concentration of 50 X 106 spermatozoa per milliliter. Biotinylation proceeded for 1 hr at 4oC and was quenched by adding glycine (Sigma) to a final concentration of 100 mM. Cells were collected by centrifugation at 350 X g for 10 min, resuspended into 1 mL of ice-cold hypotonic buffer (10 mM HEPES, pH 7.5, 1.5 mM MgCl2, 10 mM KCl, Roche protease inhibitor cocktail) and broken by dounce homogenization (50 passes, homogenizer from Braun, Melsungen, Germany). The cells were subsequently sonicated for 10 s at 40 watts (Sonifier B12, Branson Sonic Power Co.). Unbroken cells and nuclei were pelleted from the cell homogenate by centrifugation at 1000 X g for 10 min at 4 °C to generate a postnuclear supernatant (PNS). The KCl concentration in the PNS fraction was adjusted to 150 mM. A 300-μL aliquot of pre-washed immobilized NeutrAvidin biotin binding beads (Pierce) was added to the PNS fraction; the resulting suspension was mixed at 4°C for 1h. The agar beads were spun down (600 X g; 5 min) and washed 5 times with hypotonic buffer. Elution of biotinylated proteins was achieved by suspending the beads in 100 mM glycine buffer (pH 2.0) for 2 min followed by brief vortexing and centrifugation (600 X g, 5 min). The eluate was removed, neutralised using 1M TRIS buffer, pH 8.0 and digested overnight as described above for the “organic solvent solubilisation”. The subsequent peptide mixture was briefly spun (microcentrifuge), acidified to

96

3.6 Results A variety of different strategies were employed in order to overcome two main hurdles - the solubilisation sperm lipid membranes and the removal of contaminating ejaculate proteins. This was based on two main strategies: In-Gel and In-Solution digestion. The following organisational chart (Figure 3.1) shows an overview of the empirical experimental design.

Sperm Membrane Proteome

Gel Based Strategy In Solution / MS Based Strategy

Stringent Washing Stringent Washing

1D-SDS PAGE Organic Acid Organic Solvent Solubilisation Solubilisation

Various detergents Chaotrophes Affinity Enrichment / (Urea/Thiourea) Biphasic System

Trypsin/ CNBr/Trypsin/ Lipase/Trypsin/ Chymotrypsin Chymotrypsin/LysC Chymotrypsin

Figure 3.1 Organisational chart showing the strategies used to identify olfactory receptors in human sperm

3.6.1 Gel Based Strategy This was based on 1D-SDS PAGE and the laemmli procedure (Laemmli 1970). The membrane pellet was solubilised in laemmli buffer which was supplemented with a selection of detergents at 1% (w/v). These are detailed in material and methods. To evaluate which detergent/laemmli buffer combination resulted in the best solubilisation, the sperm membrane preparation was compared in parallel to an olfactory epithelia membrane preparation. Some of the resolved bands were excised; digested and analysed by ESI-MS/MS i.e. a full MudPIT was not performed.

97

For all detergent combinations contaminating soluble proteins posed a significant problem, especially semenogelin I & II – the predominant proteins in human semen responsible for spontaneous coagulation of ejaculated sperm (Robert and Gagnon 1999). To demonstrate, Figure 3.2 shows sequence coverage of semenogelin I and the iron transporter, Lactotransferrin A.

MKSIILFVLSLLLILEKQAAVMGQKGGSKGQLPSGSSQFPHGQKGQHYFGQKDQQHTKSKGSFSIQHTYH VDINDHDRTRKSQQYDLNALHKATKSKQHLGGSQQLLNYKQEGRDHDKSKGHFHMIVIHHKGGQAH RGTQNPSQDQGNSPSGKGLSSQYSNTEKRLWVHGLSKEQASASGAQKGRTQGGSQSSYVLQTEELVVNK QQRETKNSHQNKGHYQNAVDVREEHSSKLQTSLHPAHQDRLQHGSKDIFTTQDELLVYNKNQHQTKN LNQDQEHGQKAHKISYQSSRTEERQLNHGEKSVQKDVSKGSISIQTEKKIHGKSQNQVTIHSQDQEHGH KENKMSHQSSSTEERHLNCGEKGIQKGVSKGSISIQTEEQIHGKSQNQVRIPSQAQEYGHKENKISYRSSS TEERRLNSGEKDVQKGVSKGSISIQTEEKIHGKSQNQVTIPSQDQEHGHKENKMSYQSSSTE

Figure 3.2 Sequence coverage of Semenogelin II (above) and Lactotransferrin precursor (below). Sequences in red indicate unique peptides, sequences in green were identified more than once.

98

A total of 462 spectra amounting to 80.1% sequence coverage were derived using the detergent/gel based strategy. Moreover, we were not convinced that this strategy satisfactorily solubilised sperm membrane proteins – a task which is evidently more challenging than with other systems. An example gel using the best detergent, CHAPSO is indicated in Figure 3.2. CHAPSO/laemmli offered the best solubilisation as determined by the number and intensity of resolved bands, and from MS results from excised bands (see below). Nevertheless, in comparing the sperm and olfactory membrane preparations, and considering that the initial protein concentration was the same, the membrane composition of human sperm clearly hampered solubilisation and therefore protein resolution. As mentioned, many of the identified sperm proteins were either secreted or sperm cytosolic proteins. This included the precursor of Lactotransferrin, of which the resolved structure was included (Figure 3.2). Clearly there was no inadequacy in enzymatic activity which could be attributed to either the digestion conditions or even product batch. Lactotransferrin was successfully cleaved at various positions in the tertiary structure - at turns, helices and β sheets – i.e. non-membrane proteins were successfully denatured, enzymatically digested and identified by MS. This was not the case for membrane proteins, however, on a positive note, this strategy (CHAPSO) afforded the identification of one olfactory receptor (hOR17-4) and associated olfactory signal transduction proteins (Table 3.1). For hOR17-4 identification a MudPIT was performed - the CHAPSO/laemmli sample was ‘run into’ the stacking gel of an 8% SDS PAGE mini gel. The electrophoretic migration was stopped before the proteins could be actually separated by the 8% polyacrylamide, the gel was coomassie stained, and a single large protein band was excised from the stacking gel. This was washed and trypically digested, the resultant complex peptide mixture was subjected to MudPIT. The following olfactory related proteins were identified – Table 3.1 (Spehr, Schwane et al. 2004).

In spite of the fact that PercollTM purified sperm cells contained fewer ejaculate proteins than using the swim-up procedure, it became apparent that stringent washing of the membrane preparation was essential. Moreover, only one olfactory receptor was identified, therefore we turned our attention to in-solution based strategies which reportedly enhance proteomic coverage of membrane proteins.

99

Table 3.1 Olfactory proteins identified using gel based strategy

Description Accession No No. of Peptides

Type VIII 1085410; 86750 7

Type IX adenylate cyclase 27151764 12

Type VI adenylate cyclase, 10181096 5 isoform a Type VII adenylate cyclase 4557255 9

Type II adenylate cyclase 13124702 7

Type I adenylate cyclase 585023 5

Type V adenylate cyclase 25014055 4

Type IV adenylate cyclase 24497587 3

Type III adenylate cyclase 4757724 1

Golf _ (olfactory type) 4504043 4

Gs _, isoform XL-_-s 479532 2

Gy _-11 3041682 4

G _-15 4504039 4

Gt _-2 (cone-type transducin _ 2119461; 20330805 1 subunit) hOR17–4 (OR1D2) 4505517 2

3.6.2 In Solution Strategy 3.6.2.1 Solubilisation using organic acid Several attempts were based on membrane solubilisation using formic acid in conjunction with chemical cleavage by CNBr and multi-enzymatic digestions (trypsin/chymotrypsin/LysC). In addition, a significant amount of urea was added to enhance solubilisation.

Contamination from highly abundant soluble proteins was still a significant problem which also hindered the identification of low abundant membrane proteins, such as olfactory receptors. To address this problem we also employed stringent carbonate and salt washes and included the

100

addition of proteinase K as described by Wu et al (Wu, MacCoss et al. 2003). Nevertheless, despite these efforts, contamination was still a problem. No additional olfactory receptors were identified although several more ion channels were found, e.g. the ubiquitous mitochondrial voltage- dependent anion channel, VDACs.

3.6.2.2 Solubilisation using organic solvent We became aware that it is likely that olfactory receptors, as with many GPCRs, are in relative low abundance in the sperm membrane. We turned our attention to digestion conditions which may also permit cleavage of more transmembrane domains. The combination of stringent washing and the addition of 60% methanol significantly improved the identification of integral membrane proteins, however, not of olfactory receptors. The following integral membrane proteins were identified (Table 3.2)-:

Table 3.2 Integral membrane proteins identified using organic solvent strategy

Description Accession No No. of Peptides

Voltage-dependent anion channel 1 4507879 2

Voltage-dependent anion channel 3 25188179 2

Seven transmembrane protein TM7SF3 7706575 2

calcium channel, voltage-dependent, alpha 1F subunit 4885103 3

potassium voltage-gated channel KQT-like protein 3 4758630 4

Chloride intracellular channel 1 14251209 2

cholinergic receptor, nicotinic, beta polypeptide 2 25014055 2

mannose receptor C type 1 precursor 4505245 2

potassium voltage-gated channel, subfamily H, member 1 isoform 1 27437001 3

neurexin I 14149613 3

Olfactory receptor OR1D2 4505517 2

101

Although one olfactory receptor and an additional seven transmembrane spanning protein was identified, this is clearly a gross under-representation given the current literature regarding sperm integral membrane proteins (IMPs). 2D-electrophoresis has yielded more promising results. Having said this, a significant number of the remaining ‘hits’ had poor sequence coverage and/or were identified using only 1 peptide, and were therefore disregarded. If all SEQUEST ‘hits’ were taken into consideration, around 43 IMPs were identified, an improvement on previous strategies. Nevertheless, from the mass spectrometry data and the physical appearance of the post-digested pellet, it was clear that further modifications were necessary.

3.6.2.3 Vectorial labelling In order to specifically enrich for plasma membrane proteins we vectorially labelled spermatozoa using biotin. This was followed by stringent washing to eliminate contaminating proteins and was subsequently digested using the organic solvent approach. Surface labelling was performed at room temperature (5 min) and also at 4oC for the duration of 1hr.

Using both strategies a total of 310 proteins were identified, of which 85% were non-surface proteins. These included calcium binding proteins (e.g. S100) and even transcription factors. No olfactory receptors were found, although one GPCR, G protein-activated inward rectifier potassium channel 4 (ACC:1352484), was recognised. It is likely that some of the frozen spermatozoa cells were damaged and thus intracellular proteins were labelled.

This was repeated using fresh sperm from young healthly donors. A greater proportion of the ‘hits’ were membrane associated e.g. outer dense fiber of sperm tails 2 isoform 2, nevertheless, a large number of ejaculate proteins were still present in the digest including an abundance of semenogelin I & II.

3.6.2.4 Lipase & Affinity Enrichment We therefore reverted to the original methanol protocol as a foundation strategy for membrane solubilisation, which facilitates cleavage of transmembrane domains. A further breakthrough was achieved by the addition of lipase which revealed an additional 10 olfactory receptors. However, it was the amalgamation of lipase treatment with an affinity enrichment technique (aqueous biphasic system) which allowed identification of a total of 222 IMPs. The results are represented as a pie- chart (Figure 3.3).

102

a Neuronal (non-GPCR) Membrane assoc. 10% Miscell. IMPs 5% 10% Mitochondrial 5% Membrane enzymes Adhesion 6% 5%

Golgi 4% Ion Channels 9%

ER 3%

Receptors 8%

Transporters GPCRs (incl.ChemoRcs) 9% 26%

Synapic Assoc. 14% Adhesion 8% b Metabotropic Receptors 5%

Ionotropic Receptors 8%

Miscellan. Chemoreceptors 10% 55%

Figure 3.3 A total of 222 Integral Membrane Proteins were identified (a), of which, a large proportion were neuronal proteins (b)

A significant proportion of identified IMPs were neuronal proteins which included GPCRs (Figure 3.3-a). GPCRs are notoriously difficult to identify, nevertheless, they accounted for around one- quarter of the total dataset. Interestingly, 63 ‘classical’ neuronal proteins were detected. Over half of these were chemoreceptors (Figure 3.3-b) comprising of 32 olfactory receptors and 3 taste receptors. These chemoreceptors are all GPCRs.

103

Upon closer inspection of the MS data it was evident that the majority of GPCRs were identified by peptides cleaved from transmembrane domains. To exemplify, the spectra and cleaved peptides which facilitated the identification of one olfactory receptor, OR9A4, are shown (Figure 3.4).

T I Y S C N Q L N K Q T L Q … … T P S K A V M L M A V A I V F … C T Y D E E I Y L V L Y H T V E H V F T P F L N V W C L I S W I I Q Q I G V L E L F L Y V F L F L F F G S A A Y L G M L I L I S F A V M F L L V V V V F H A V W F T C S F G G S V V F L T V L F V Y F T L F H T L F E V G S P V M Y F V S A F L T A F L C T L M P S L N I S N L C P F P G L Q G T T F N T A I F I V M V T I R A S H S L I K V K N R D R … M D N R I V V R C N A R R M Y N Y G I … … S I I L P S R Y N S #7844-7844 RT:80.04-80.04 NL: 2.58E3

100

95

90

85

80 y1 -7 833.2y2 -1 5 75 837.6

70

65 b1-12 1305.8 b1-8 60 859.2 55

50

45 Relative Abundance Relative 40 b2-15 824.0 35

30 y2 -1 4 809.9 y1 -8 y2 -1 6 25 945.6 911.6 y1 -9 y1b2-8 -3 b1-14 b2-13 1076.7 y1 -1 3 20 416.1430.6 b1-7 y1 -1 0 1519.9 703.0745.6 1503.9 1190.7 15 b1-4 432.3 b1-11 b1-6 b1-13 b2-12 1202.6 10 b2-9 632.4 1404.9 b2-11 653.3 b2-7 496.3 b1-10 b1-15 b1-3 b1-5602.2 b1-9y2 -1 7 5 b2-6 373.2 1648.0 y1 -2 318.2 990.6968.2 1103.8 y2 -1 y1b2-3 -1 b2-5 317.2 533.5 y1 -1 5 y1 -1 6 66.6 132.2159.7 267.3288.3 1675.1 1822.3 0 0 200 400 600 800 1000 1200 1400 1600 1800 m/z

Figure 3.4 Cleaved peptides. Example of sequence coverage (a) and spectra (b) from OR9A4. (c) schematic representation of a GPCR which indicates the origin and frequency (%) of cleaved peptides used to identify the receptor – e.g. 16% of peptides used to identify GPCRs were cleaved from TM1

Evidently the strategy facilitated access to hydrophobic transmembrane domains as exemplified by Figure 3.4. This data was statistically scrutinised to evaluate whether this enrichment strategy and/or MudPIT had any biases in addition to hydrophobicity. For example, are only basic or large proteins identified using this strategy? Is there any correlation between cleaved peptide isoelectric point and hydrophobicity?

104

The following 3D charts (Figure 3.5) plot the relationship of calculated pI, molecular weight (kDa) and hydrophobicity values (GRAVY) for the identified cleaved peptides and proteins. In both cases regression analysis reveals that no significant correlation is evident although from b) it is clear that the majority GPCRs had a molecular mass around 30 kDa. Given the fact that around half of the identified GPCRs were chemoreceptors (Table 3.3) this was unsurprising. This is further illustrated by Figure 3.6, a virtual 2D-gel displaying all of the identified IMPs. This was performed using the online tool JVirGel (http://www.jvirgel.de/). The chemoreceptors, with a relatively narrow pI and MW range, are indicated with red circles.

Figure 3.5 Scatter plots showing relationship between molecular weight (x), calculated GRAVY (y) and calculated pI (z) for a) cleaved peptides used to identify GPCRs & b) the identified GPCRs. Regression analysis including confidence (95%) and prediction intervals are also shown (grey hatched areas). [The confidence interval is 95% sure to contain the best-fit regression line]

105

Figure 3.6 Virtual 2D gel showing calculated positions of membrane proteins identified using optimised strategy. Red circles indicate identified chemoreceptors

In spite of the fact that there is no obvious correlation for the peptides or proteins as determined from their calculated physicochemical properties (pI, MW and GRAVY scores), an obvious observation remains – one-quarter of the identified membrane proteins were GPCRs, over half of which were chemoreceptors. This prompts the question – is this an accurate and proportional representation of sperm membrane receptor proteins?

The identified chemoreceptors (32 olfactory receptors, 3 taste receptors) are tabulated below (Table 3.3) using HUGO Gene Nomenclature Committee (HGNC) approved classification (Wain, Bruford et al. 2002). The nomenclature is based on a divergent evolutionary model and is outlined below (Glusman, Bahar et al. 2000)-:

106

OR 3 A 5 P

ROOT olfactory receptor Family >40% Sub-family >60% Individual gene number Pseudogene superfamily protein homology protein homology with sub-family (where needed)

Table 3.3 List of identified chemoreceptors – a total of 32 olfactory receptors and 3 bitter taste receptors were identified

Identified Chemoreceptors

Olfactory receptor 1B1 Olfactory receptor 4A4 Olfactory receptor 11G2 Olfactory receptor 1D2 Olfactory receptor 5K1 Olfactory receptor 4C15 Olfactory receptor 1S2 Olfactory receptor 4F6 Olfactory receptor 51S1 Olfactory receptor 2S2 Olfactory receptor 11A1 Olfactory receptor 4C15 Olfactory receptor 2C1 Olfactory receptor 5U1 Olfactory receptor 2C1 Olfactory receptor 2C3 Olfactory receptor 10R2 Olfactory receptor 13C2 Olfactory receptor 4A15 Olfactory receptor 6P1 Olfactory receptor 11A1 Olfactory receptor 10J6 Olfactory receptor 5AR1 Olfactory receptor 51L1 Olfactory receptor 9A4 Olfactory receptor 5AN1 Olfactory receptor 10G8 Olfactory receptor 52E1 Olfactory receptor 7A10 Olfactory receptor 9K2 Olfactory receptor 13C3 Olfactory receptor 5B17 Taste receptor type 2 member 49 (T2R49) Taste receptor type 2 member 39 Taste receptor type 2 member 7 (T2R39 (T2R7)

A wide spectrum of olfactory receptors from various families and chromosomes were identified. Multi-alignments (CLUSTAL; http://www.ebi.ac.uk/clustalw/) failed to identify any obvious motifs or trends which may explain why these receptors, and not others, were identified in sperm.

3.6.3 Validation of data It was surprising to not only identify so many neuronal receptors, but also an abundance of chemoreceptors. In addition, bitter taste receptors have never been previously identified in human spermatozoa.

107

In order to validate some of our findings, we performed immunocytochemistry on several neuronal membrane proteins, and employed calcium imaging to address whether the taste receptors could be functional.

3.6.3.1 Functional validation - calcium imaging Freshly PercollTM prepared sperm from a single healthy donor was loaded with a calcium sensitive fluorescent dye, fura-2-AM. Two bitter taste agonists, 500 μM papaverine and 1mM strychnine (Lemon and Smith 2005) were dissolved in standard Ringer solution and applied sequentially to the sperm preparation. Papaverine and strychnine are the ligands for the taste receptors, T2R7 and TR47, respectively (unpublished). This was followed by application of a positive control, 30 μM progesterone, which elicits an intracellular calcium rise in a dose-dependent manner (Kobori, Miyazaki et al. 2000). From Figure 3.7 it is evident that the spermatozoa responded to both ligands, suggesting that both taste receptors may be functional. One must accentuate that without appropriate antagonists and/or additional confirmatory experiments, e.g. applying suitable inhibitors to the PLC pathway (Zhang, Hoon et al. 2003), this data is only suggestive. Moreover, cell viability assays should also be performed in order discount any non-specific effects due to toxicity.

300

250

200

150 f340/f380 100 papaverine strychnine progesterone

50

0 1 15 29 43 57 71 85 99 113 127 141 155 169 183 197 211 225 t (sec)

Figure 3.7 Calcium imaging of human sperm. Papaverine and strychnine, the respective agonists for T2R7 and TR47 elicit a response

108

3.6.3.2 Immunocytochemistry The neuronal membrane proteins NCAM1, GluR1, P2X3 and Syntaxin 3 were identified using MudPIT. In order to validate this we performed immunocytochemistry as indicated in Figure 3.8. The nucleus was stained with DAPI (blue).

Figure 3.8 Immunocytochemistry validation of a section of neuronal proteins: NCAM1, P2X3, GluR1, and Syntaxin 3

The GluR1 and NCAM staining was mostly localised to the midpiece region whereas P2X3 was only found in the tail. Surprisingly, the synaptic protein Syntaxin 3 was evidently ubiquitous localising to the head, midpiece and tail.

The immunocytochemistry panel therefore substantiates the membrane proteins identified using MudPIT.

109

3.7 Discussion The aim of this project was to proteomically analyse the sperm membrane proteome with a view to ‘uncover’ chemo-receptors. We systematically developed a method which culminated in the successful identification of 32 olfactory receptors and 3 bitter taste receptors. In doing so we also identified an additional 187 membrane proteins, including neuronal proteins.

Two main strategies were used: ‘Gel based’ and ‘in-solution based’ approaches. The evolution and optimisation of these strategies are discussed below.

3.7.1 Strategy development From the outset it became clear that the sperm membrane preparations were rather resilient to solubilisation. As mentioned, the sperm lipid membrane content is quite unique when compared to mammalian somatic cells (Martinez and Morros 1996). For example, sperm membranes contain an abundance of unsaturated plasmogens (glycerophospholipids) which may play a role in creating non-diffusible domains and mediating certain signal transduction events. Furthermore, there are reports of cold detergent-resistant membrane microdomains (Cross 2004), the dispersal of which may terminate a set of reactions that suppress capacitation. It is likely that this rather unique lipid content compounded the solubilisation strategies. When compared to other systems, e.g. olfactory epithelium membrane preparations, solubilisation and digestion was more problematic.

3.7.1.1 Gel Based Strategy It was clear that IEF 2D-electrophoresis could not be used. As reviewed in the introduction, 1D- SDS PAGE is still a viable option for membrane proteomics. Our first trials simply performed standard 1D-SDS, as described by Laemmli (Laemmli 1970) i.e. the sample was reconstituted in laemmli buffer, reduced and subjected to heat denaturisation. This resulted in an unmanageable precipitate. When additional reducing agent, detergent (1% CHAPSO) and a lower ‘denaturisation’ temperature was used it resulted in improved sample solubilisation. This facilitated the identification one olfactory receptor and associated signal transduction proteins (Spehr, Schwane et al. 2004).

A significant benefit of in-gel based methods is the possibility to use ‘strong’ and relatively high amounts of detergents. We speculated that using a combination of detergents and laemmli buffer should improve the solubilisation of olfactory receptors.

110

3.7.1.1.1 Detergents Traditionally, the challenge of membrane proteins solubilisation has been addressed by the use of detergents. Lipid bilayers are dissolved to form lipid-water-soluble complexes and thus lipids and IMPs are maintained in soluble form. Such complexes of detergent with protein and possibly lipid are referred to as mixed micelles (Jackson, Schmidt et al. 1982). According to the three-stage hypothesis, membrane solubilisation is an all-or-none process with the characteristics of a phase transition (Kragh-Hansen, le Maire et al. 1998). Firstly the detergent is ‘associated’ with the membranous phase (stage I), following which it is incorporated forming mixed micelles. A myriad of detergents are commercially available. They come in a variety of molecular topologies and depending on the physicochemical properties of the detergent and the detergent/lipid ratio, different mixed micelle shapes and sizes are possible. Detergent charge is the largest factor effecting solubilisation; this attribute is used for classification – anionic (negatively charged), cationic (positively charged), non-ionic and zwitterionic.

A variety of detergents were employed in empirical method development. Anionic (SDS, Deoxycholic acid, N-Lauroylsarcosine, Sodium taurocholate), non-ionic (Triton X-114/100, Nonylphenyl Polyethylene Glycol NP-40) and zwitterionic (CHAPS, CHAPSO, ASB-14), detergents were used. A binary mix of SDS (in laemmli) and CHAPSO under reducing conditions (DTT) yielded the most promising results with sperm, however, this was still inadequate with only one olfactory receptor (hOR 17-04) being identified. Sodium dodecyl sulfate (SDS) has a polar anionic sulfate group at one end of the molecule and a straight chain nonpolar region at the other end. This dual polarity bestows SDS with powerful disruptive and solubilisation characteristics – the molecule ‘mimics’ protein structure inserting into and disrupting lipid bilayers, and is probably the strongest denaturing detergent known. In addition, CHAPSO was used as a co-detergent – this zwitterionic detergent possesses a rigid structure and is analogous to CHAPS. CHAPS has been commonly used to solubilise receptors such as those for colony-stimulating factor (0.5% CHAPS), opiates (0.6% CHAPS), neurotensin (0.6% CHAPS), somatostatin (0.6% CHAPS) and human prostatic sex hormone-binding globulin receptor (6.0% CHAPS) (Banerjee, Joo et al. 1995). Both CHAPS and SDS have been previously reported to successfully solubilise 5HT-1A receptors

111

(GPCR) - both detergents have a relatively high critical micelle concentration§§ (CMC ≈ 8-10 mM) and are better at solubilising lipid as compared to protein. The authors proposed that the serotonin receptor (5HT-1A) is associated with phosphatidylethanolamine (PE), phosphatidylserine (PS) and phosphatidylinositol (PI) domains and is consequently selectively enriched using CHAPS and SDS. This was not the case upon application of triton X-100/114. In a separate investigation of olfactory epithelium membrane proteins (data not shown), this strategy was relatively successful; however, in the case of the sperm membrane proteome, detergents were clearly inadequate.

Figure 3.9 The lipid composition of human sperm. Modified from Martinez and Morros (1996).

§§ Critical micelle concentration (CMC) is defined as the concentration of molecules in free solution in equilibrium with molecules in aggregated form. High CMCs indicate relatively weak hydrophobic binding strengths whereas with lower CMCs a more stable micelle is formed. Proteins are incorporated more slowly into these micelles and moreover, are also harder to remove.

112

To reiterate, it is likely that the complex lipid composition of sperm (as illustrated in Figure 3.9) effectively annulled any of our attempts at olfactory receptor solubilisation using detergent. This was further compounded by an abundance of contaminating soluble proteins and motivated development of an alternative strategy.

3.7.1.2 ‘In-solution’ based strategies In order to improve membrane solubilisation and to facilitate peptide cleavage from transmembrane domains, we focussed on ‘in-solution’ digestion. This was coupled with stringent washing. The strategies are discussed below.

3.7.1.2.1 Organic acid (formic) solubilisation In many circumstances membrane solubilisation in formic acid is a powerful and robust technique which can facilitate the identification of integral membrane proteins (IMPs). It has been employed in previous MudPIT strategies (reviewed in introduction) and recently the multi-transmembrane protein aquaporin 0 has been fully sequenced using a formic acid based strategy (Han and Schey 2004). A similar strategy examined membrane proteins from two breast cancer cell lines (Xiang, Shi et al. 2004) and in another study, this approach facilitated the identification of 242 membrane proteins from Caenorhabditis elegans (Mawuenyega, Kaji et al. 2003). ‘Shotgun proteomics’, in particular MudPIT, is evidently as powerful today as when it was first described. Admittedly the technique is evolving (Swanson and Washburn 2005), nevertheless, the principle strategy for membrane protein analysis often employs formic acid and CNBr for solubilisation and chemical cleavage of membrane proteins, respectively.

In this case no olfactory receptors were found. This was most likely due to an abundance of contaminating soluble proteins. Moreover, most membrane proteins that were identified were mitochondrial in origin. Relatively few transmembrane cleaved peptides were also identified. As a consequence we decided to use an organic solvent, methanol, in order to improve solubilisation and cleavage of transmembrane regions.

3.7.1.2.2 Organic solvent (methanol) solubilisation Following a report by Goshe recommending the use of methanol for proteomic analysis of hydrophobic integral membrane proteins, there has been increased interest in this strategy (Blonder, Goshe et al. 2002; Goshe, Blonder et al. 2003). This technique was recently modified to obtain a

113

comprehensive membrane proteome of Corynebacterium glutamicum. 326 integral membrane proteins were identified including proteins with up to 24 transmembrane helices (Fischer, Wolters et al. 2005). Thus, we applied this strategy to the sperm pellet – trypsin/chymtrypsin digestion in the presence of organic solvent (methanol). A total of 43 IMPs were revealed, however, many of these were excluded when stringent SEQUESTTM search criteria was stipulated e.g. a minimum of two peptides was required for protein identification. Moreover, despite this advance, contaminating soluble proteins were still problematic.

To address this problem we combined the methanol strategy with vectorial labelling. Freshly prepared sperm were washed in standard ringer solution and vectorially labelled using biotin. Surprisingly there was no improvement in the removal of soluble proteins which even included the ejaculate proteins semenogelin I & II. With hindsight we now argue that vectorial labelling of human sperm is intrinsically inadequate, at least with regard to excluding soluble and ejaculate proteins. Due to the fact that semen proteins interact with sperm, for example, the fusion of prostrate secreted proteosomes with spermatozoa (Arienti, Carlini et al. 2004), and the ‘coating’ of sperm by seminal plasma proteins (Iborra, Morte et al. 1996), biotin will also label these proteins. It is likely that in order to remove contaminating proteins the whole sperm cells must be stringently washed, a procedure which would undoubtedly lead to cellular damage. We stress that conceptually vectorial labelling is indeed beneficial – using optimised labelling conditions one will only enrich membrane and membrane associated proteins. It has been successfully used in cell lines where mild washing prior to labelling yielded satisfactory results (Zhao, Zhang et al. 2004). Vectorial labelling was previously performed on human sperm. Naaby-Hansen (Naaby-Hansen, Flickinger et al. 1997) used a relatively high concentration of biotin at a temperature of 37oC for 10 minutes. The concentration of biotin, the temperature and duration of labelling are critical. Considering the normal testicular temperature is 35oC it is likely that Naaby-Hansen’s conditions lead to biotinylation of intracellular proteins. Furthermore, we recommend that in principle biotinylation should only be performed at 4oC, thus slowing metabolic activity and minimising any possibility of up-take e.g. via endocytosis. The vectorial labelling strategy was therefore abandoned, reverting to the methanol/chymotrypsin/trypsin approach.

A breakthrough was realised upon addition of a non-specific lipase to the membrane pellet. The digestion mixture consisted of methanol, lipase, trypsin, and chymotrypsin. Instantly, 11 olfactory

114

receptors were identified. The visible appearance of the pellet also changed. This is potentially due to lipase emulsification of the sperm membrane, hydrolyzing triglyceride to free fatty acids and partial glycerides, ameliorating the effect of methanol which denatures protein and disrupts lipid- shell-hydrophobic interaction (Anderson and Jacobson 2002; Blonder, Hale et al. 2004). Nevertheless, contaminating proteins were still a problem. The next breakthrough was the use of an aqueous biphasic system – instead of generating a membrane pellet which inevitably contains contaminating proteins, the whole cells were homogenised in a system designed to fractionate proteins according to hydrophobicity. A brief review of aqueous biphasic systems follow.

3.7.1.2.3 Biphasic system Aqueous biphasic systems are ideally suited to biological systems as the major component in both phases is water. The phenomenon was first reported in 1896 by Beijerinck, who noted that certain polymers (agar and gelatine) in an aqueous environment demonstrated a peculiar (“eigentümlich”) incompatibility (Beijerinck 1896). It later became apparent that many immiscible biphasic systems exist, the most commonly reported of which is the dextran-polyethylene glycol (PEG) system (Alan D. Diamond 1990). When both phases are solvated, the dextran forms a more hydrophilic, denser, lower phase and PEG a more hydrophobic, less dense, upper phase. The hydrophobic nature of

PEG, [HO-(CH2-CH2-O)n-H where n is the degree of polymerisation] is mainly due to the polymer’s methylene groups. Partitioning of the two phases is a complex phenomenon, involving interaction between the partitioned substance and the components of each phase. Several different physicochemical interactions are involved including hydrogen bonding, charge interaction, van der Waals’ forces, hydrophobic interaction and steric effects. Depending on the molarity of PEG (Klofutar 2003), the temperature (Flanagan and Barondes 1975; Johansson, Gysin et al. 1981; Ollero, Pascual et al. 1994; Persson and Jergil 1995), pH and salt concentration (Walter 1985), the partition properties are effected. This partitioning behaviour may be understood in terms of the

Gibbs free energy of hydration (ΔGhyd) of the salt used and of the concentration of the salt stock solution used to form the biphase. At its most elementary, this relationship may be expressed as-:

K part = Ctop /Cbottom

where, Kpart is the partition coefficient given by the ratio of the concentration, C, in the top and bottom phases. This is of course a gross simplification, a comprehensive expression is derived by

115

Diamond which is specific for the dextran-PEG system, and also takes into account electrical, hydrophobic, hydrophilic and conformation effects (Alan D. Diamond 1990).

Notably, salts have a major effect on this system effecting the cloud point (point when the interface is formed) and have preference for PEG-rich or PEG-poor phases in accordance with the Hofmeister Series*** (Hofmeister 1888)-:

- - - - 2- + + 2+ 2+ Preference for PEG-rich phase: I > Br > Cl > F > SO4 ; Cs > Na > Ba > Ca

2- - - - - Preference for PEG-poor phase: SO4 > F > Cl > Br > I

To conclude, this aqueous biphasic system adequately separates membrane proteins from their hydrophilic counterparts.

3.7.1.2.4 Affinity coupled biphasic system In order to improve the likelihood for identification of olfactory receptors an additional affinity step was included. As the vast majority ORs have N-glycosylation (NxT/S) motifs (Zozulya, Echeverri et al. 2001) a lectin affinity step was included. The term ‘lectin’ was first coined to describe erythrocyte-agglutinating proteins however, today it describes a large number of proteins which either bind to or cross-link carbohydrates (Sharon and Lis 2004). Wheat germ agglutinin (WGA) was selected due to its affinity for the entire sperm membrane (Cross and Overstreet 1987) although alternative lectins may be used in order to ‘dissect’ the sperm plasma membrane e.g. peanut agglutinin is specific for the outer acrosomal membrane (Flesch, Voorhout et al. 1998). WGA was coupled to activated dextran as indicated in Figure 3.10 below.

The resultant affinity ligand ‘pulls-down’ glycosylated membrane proteins from the upper phase forming an enriched pellet after addition of a competitive ligand and ultracentrifugation.

*** The Hofmeister series describes the ‘power’ of various ions to precipitate a mixture of hen egg white proteins. Simply put, ions with a high affinity for water remove the ‘free water’ available for protein hydration; therefore, the effective concentration of proteins increases inducing precipitation.

116

Lectin

O O NH

i ii O HO O LGO O O O HO HO RO RO O O OH OH HO O O RO OH O

Dextran activated Dextran lectin-coupled Dextran

i: activation of hydroxyls with appropriate leaving group ii: incubation with lectin in buffered soln., nucleophilic substitution

LG: leaving group, e.g. trifluoroethanesulfonic ester R: glucose(α1->3), one or two units

Man (α1−6) Fuc (β1−6) O Man (β1-4) – GlcNAc (β1-4) – GlcNAc (β1-N) – N - Asn C - Asn N Recognition region of Man (α1−3) Wheat Germ Agg. H

Figure 3.10 Coupling of affinity ligand (WGA) to dextran. The recognition motif for WGA is also shown (below)

3.7.2 The Solution: A final strategy To recap, sperm whole cells were homogenised in a biphasic system comprised of a hydrophobic upper phase (PEG) and a hydrophilic lower phase (dextran). Consequently membrane proteins are ‘forced’ in the upper phase. In order to enrich for olfactory receptors an additional affinity step was employed. The lectin-coupled dextran washing step enriched for all glycosylated hydrophobic membrane proteins containing GlcNAc-moiteies, including GPCRs. The resultant pellet was digested using a powerful digestion combination of methanol, trypsin, chymotrypsin and lipase. The lipase emulsifies any sperm membrane micelles ameliorating the effect of methanol which denatures protein and disrupts lipid-shell-hydrophobic interactions. In doing so we suggest that the enzymes, trypsin and especially chymotrypsin, can therefore gain access to transmembrane regions. This final strategy is schematically represented by Figure 3.11.

117

Figure 3.11 Strategy to identify olfactory receptors in human sperm. Hydrophobic, glycosylated proteins are enriched using a biphasic system employing an additional affinity step (lectin). The pellet is digested using a combination of methanol and an enzymatic cocktail (lipase/trypsin/chymotrypsin)

This original and optimized strategy facilitated identification of a battery of integral membrane proteins including chemoreceptors, many of which were novel findings.

3.8 Identified membrane proteins We were able to identify a total of 222 integral membrane proteins (Figure 3.3). Of particular interest are the GPCRs and neuronal proteins, many of which are novel findings. Before discussing these findings from a biological perspective, a brief technical analysis from a proteomic vantage is given.

118

3.8.1 A technical discussion on the identified proteins Interestingly, the vast majority of GPCRs were identified by transmembrane peptides as schematically represented in Figure 3.4. Contrary to these findings, it has been reported that such hydrophobic and relatively long peptide sequences usually escape detection by mass spectrometry (Eichacker, Granvogl et al. 2004). We believe it is likely that any limitation in detection is due to the digestion conditions (i.e. steric hinderence to transmembrane regions and/or lack of cleavage sites) and not to the physicochemical properties of an ionized hydrophobic peptide in the mass spectrometer. In fact, it has been suggested that the precursor ion††† from hydrophobic peptides is more intense than for hydrophilic peptides (Nadja B. Cech 2001). This is probably due to a higher affinity of hydrophobic peptides to the surface of the aqueous ESI droplet. Consequently, the hydrophobic adducts are more easily desolvated‡‡‡ and carry a greater fraction of the excess charge produced in the electrospray process. Conversely, ‘internal’ hydrophilic adducts carry less charge (Fischer, Wolters et al. 2005). Moreover, the vast majority of transmembrane peptides were cleaved by chymotrypsin. This is unsurprising considering that chymotrypsin cleaves peptide bonds selectively on the carboxylterminal side of large hydrophobic amino acids such as tryptophan, tyrosine, phenylalanine, and methionine. Furthermore, from bioinformatic analysis it is evidence that the vast majority of transmembrane regions lack trypsin cleavage sites (Eichacker, Granvogl et al. 2004).

In order to expose any potential correlation involving the pI, MW and hydrophobicity of the cleaved peptides and their ‘parent’ GPCRs, XZY charts were plotted (Figure 3.5). There was no statistically significant correlation linking the physicochemical properties of the peptides or their associated proteins. The highest correlation was seen in the receptor protein interaction chart (3.5- b). This was a reflection of a seemingly reciprocal relationship between receptor size and hydrophobicity. This correlation was poor (r2=0.69) and moreover, bioinformatic analysis of human proteins revealed a direct correlation between the length of membrane proteins and hydrophobicity. As protein length increases, so does protein hydrophobicity (Mitaku and

††† Precursor ion : An ion that reacts to form particular product ions. In the case of precursor ion analysis of peptides, the parent ion represents a peptide, this is ‘selected’ by the mass spectrometer and ‘sequenced’ by colliding the peptide with neutral molecules in a chamber - collision induced dissociation (CID). The resultant ‘product ions’ are then separated by a second stage of the mass spectrometer to yield a ‘product ion spectrum’. This spectrum represents the peptide sequence. ONLY intense or ‘strong’ precursor ions are ‘selected’ by the mass spectrometer which results in peptide ‘sequencing’. ‡‡‡ Desolvation: The removal of solvent adducts from a gas phase adduct ion.

119

Hirokawa 1999); i.e. opposite to the trend demonstrated by Figure 3.5b. This unexpected reciprocal relationship may be due to the large proportion of identified chemoreceptors which possess a relatively narrow range in terms of isoelectric point and molecular mass (Figure 3.6). The chemoreceptors, for their size, are relatively hydrophobic.

Thus, no clear correlation could be identified when one considers the pI, MW and hydrophobicity of the cleaved peptides and their associated proteins. Nevertheless, the question still remains – why were the majority of GPCRs identified from transmembrane cleaved peptides? This could be due to two reasons. Firstly, as suggested above, hydrophobic peptides have more intense parent ions and are therefore selected for MS/MS. Secondly, hydrophobic peptides elute either first or later in the MudPIT run. In doing so they do not ‘compete’ (co-elute) with more abundant hydrophilic peptides which elute during the intervening gradients from the HPLC. Early eluting hydrophobic peptides and are not retained by the cation exchange. Conversely late eluting hydrophobic peptides are retained by the reverse phase.

To summarise, we have established a technique which facilitates the identification of human sperm integral membrane proteins, a significant proportion of which were 7TM GPCRs. The optimised solubilisation/digestion condition facilitated cleavage of transmembrane peptides which were then identified by ESI-MS/MS. The evident bias in the number of transmembrane peptides which facilitated identification of GPCRs could be due to i) superior ionization of hydrophobic peptides and ii) later elution of hydrophobic peptides from the HPLC.

Undoubtedly this novel approach has proved a powerful means to identify sperm membrane proteins, nonetheless, of equal if not greater importance are the identified novel proteins.

3.8.2 The identified proteins from a biological perspective Many interesting and some surprising proteins were identified, obviously a detailed discussion on each protein is unfeasible and therefore a few of the more interesting findings were selected for elaboration. It is proposed that these proteins are good candidates for future work aimed at unravelling signal transduction.

The aim of this project was to identify novel olfactory receptors in human sperm - this was accomplished. Of particular interest was their diversity – a total of 35 chemoreceptors comprising

120

of 3 bitter taste receptors and 32 olfactory receptors. For the first time it was demonstrated that two bitter taste receptors (T2Rs) are not only expressed in human sperm, but that they may also be functional (Figure 3.7). In calcium imaging experiments human spermatozoa responded to 500μM Papaverine and 1mM Strychnine, the respective ligands for T2R7 and TR47 (unpublished). The T2Rs are a family of 30 GPCRs which are usually considered to be selectively expressed in the tongue and palate (Scott 2004). Could chemoreception (taste and olfaction) play a cooperative role in fertility? The evidence for a role for olfactory receptors is mounting, however, a series of critical experiments are necessary before taste receptors can be implicated.

Using homology searches and low-stringency polymerase chain reaction, Parmentier (Parmentier, Libert et al. 1992) previously reported approximately 50 mammalian olfactory receptor transcripts in testis. Surprisingly only one of these (OR5K1) was accounted for in our proteomic dataset, however, the Parmentier receptors were screened from many testicular cells types. We proteomically confirmed the existence of hOR17-4 (Spehr, Schwane et al. 2004) in human sperm – this receptor was also unreported by Parmentier. In fact, there is no doubt that hOR17-4 is functional in human sperm. The lack of agreement between the Parmentier receptor list and our dataset is, to say the least, surprising. Could even more olfactory receptors be expressed in human sperm? Notwithstanding this, the current proteomic dataset already raises the interesting question why such a wide receptor spectrum with a broad receptive field is expressed in mature spermatozoa. Could some receptors play a role in spermatid development and are redundant in the mature cell? Perhaps only smaller receptor subsets actually facilitate chemotaxis? It is also possible that there are heterogeneous sperm subpopulations having varying degrees of fecundity or different functions in fertility. Several sperm subsets have been suggested – the ‘egg getter’ (interacts with oocyte), ‘killer’ sperm (interacts with cellular obstructions e.g. leucocytes and bacteria) and ‘blocker’ sperm (block spermatozoa from other males) (Tatsura, Nagao et al. 2001). It has already been shown that only a small fraction of mammalian sperm (approx. 10%) is capacitated and chemotactically responsive (Eisenbach 1999). Moreover, using an alternative aqueous two-phase partition system, Cartwright reported that bovine spermatozoa have heterogeneous cell populations (Cartwright, Cowin et al. 1991). Whether these subpopulations have a functional role in mammalian fertilization is uncertain, nevertheless, only a fraction of ejaculated sperm is actually capable of fertilization (Holt and Van Look 2004). Finally, we must also take into account the fact that for proteomic experiments up to 15 donors were required for each experiment. It is feasible

121

that an individual donor may only express a subset of the identified receptors. Admittedly these hypotheses are highly speculative but even so, they merit discussion and future investigation.

In addition to chemoreceptors we also identified several other GPCRs including four probable G- protein coupled receptors, three of which belong to the adhesion (formely known as LN-TM7) subfamily (Krasnoperov, Lu et al. 2002). This relatively recently discovered subfamily possesses distinctively long N-termini with multiple binding domains. The N-termini are thought to be highly glycosylated due to an abundance of Ser and Thr residues and presumably have a function in cell-to-cell and cell-to-matrix interaction. Moreover, most adhesion GPCRs contain GPCR Proteolytic Site (GPS) domains – a cysteine-rich site of around 75 amino acids which may be proteolytically cleaved. Virtually all proteins that contain GPS domains are heptahelical receptors with the interesting exception of the sea urchin sperm receptor, suREJ (Hughes, Ward et al. 1999). suREJ binds to glycoproteins on the sea urchin egg surface (egg jelly) triggering the acrosome reaction, a calcium mediated process which enables the sperm to penetrate the egg. The human homologue is termed ‘polycystic kidney disease and receptor for egg jelly related gene’ (PKDREJ) which is suggested to have a similar role in fertility (Castellano, Trevino et al. 2003). We did not detect PKDREJ in mature human sperm although arguably our candidate secretin receptors may also play an integral role in sperm-egg fusion and perhaps even in the acrosome reaction.

‘Classical neuronal’ proteins were also identified; some of which have been validated using immunocytochemistry (Figure 3.8). These not only included ionotropic (e.g. GluR1, Figure 3.8) and metabotropic receptors, but also neuronal adhesion molecules such as neuronal cell adhesion molecule 1 (NCAM) and beta-neurexin. NCAM is expressed widely in embryonic development, however, in adulthood it is predominantly found in neuronal tissues where it is implicated in neuron-neuron adhesion, neurite outgrowth and fasciculation. It has been suggested that NCAM facilitates adherence of germ stem cells (gonocytes) to sertoli cells which are critically important in spermatozoa development. From Figure 3.8 it is clear that most NCAM staining is localized to the mid-piece. This region is rich in mitochondria providing energy for flagellar beating. It is currently unknown whether NCAM has a function in mature human sperm. Unlike NCAM, beta-neurexin has never been associated with reproduction and is considered to be exclusively neuronal. Beta- neurexin is thought to play an important role in synapse formation and maturation. Furthermore, we also identified additional synaptic machinery proteins including syntaxin 3, synaptogyrin 4 and

122

secretory carrier-associated membrane protein 5 (SCAMP5). Syntaxin 3 localised to the entire sperm cell body (Figure 3.8). This was a surprising finding considering that in mouse sperm syntaxin 2 colocalizes with synaptotagmin to the acrosomal compartment - both proteins are involved in neuronal secretion (Hutt, Baltz et al. 2005). It has been suggested that the sperm acrosome is essentially a secretory granule derived from golgi and is analogous to synaptic vesicles in neurons (Abraham L. Kierszenbaum 2000; Hutt, Baltz et al. 2005). Indeed, increasing evidence suggests that the synaptic exocytosis machinery such as the SNARE complex (Soluble NSF Attachment protein REceptor) is involved in the ‘all-or-nothing’ acrosome reaction. It is therefore unclear why syntaxin 3 is localised to the entire sperm cell.

In addition, the ATP purinergic receptors P2X3, P2X4 and P2Y11 were identified. P2X receptors function as ligand-gated, transmembrane cation channels that allow influx of extracellular cations, 2+ including calcium. Consequently, the receptors mediate an intercellular elevation in calcium [Ca ]i and also instigate depolarisation. They have consensus sequences for N-linked glycosylation (NxS/T) and display different desensitisation kinetics - P2X3 is fast (milliseconds) whereas P2X4 is markedly slower (seconds) (North 2002). The proteomic identification of P2X3 was validated using immunocytochemistry (Figure 3.8) where it localised to the tail and midpiece. There has been no previous reports of P2X3 in mature human spermatozoa, although it has been located in rat germ cells (Glass, Bardini et al. 2001), the rat vas deferens and prostate (Lee, Bardini et al. 2000). It seems likely that P2X receptors may play a neuronal role in the vas deferens in facilitating the sympathetic nervous ejaculatory reflex (Mulryan, Gitterman et al. 2000). P2X3 may also have a function in the development of spermatids, however, its role in the mature spermatozoa is less clear. ATP has been shown to activate a P2Y receptor, and not P2X receptor, resulting in an increase of intracellular calcium which culminates in the acrosome reaction (Shariatmadari, Sipila et al. 2003).

Finally, an array of ion channels were also identified. These included - voltage gated potassium/ calcium/sodium channels; potassium inwardly-rectifying channels; chloride channels; cyclic nucleotide-gated channels and even transient receptor potential (TRP) channels. With respect to the latter, TRPC channels have already been implicated in the mouse acrosome reaction (AR) and TRPC1, 3, 6 and 7 have also been immunolocalised on mature human sperm (Castellano, Trevino

123

et al. 2003). A specific blocker of TRPC channels, SKF96365, has also been shown to retard human spermatozoa motility.

In conclusion, from the results it is clear that this strategy is a promising technique in order to identify human sperm membrane proteins. In doing so it is axiomatic that the ‘simple sperm cell’ is much more complex as often attributed. These findings shed new light on potential molecular mechanisms underlying these processes and raise pertinent questions for current biological research.

4 Final Conclusion This project analysed the olfactory receptor associated proteome. It comprised of two main themes: A proteomic examination of olfactory receptor plasticity in the olfactory epithelium and the identification of olfactory receptors expressed in human spermatozoa.

4.1.1 Olfactory receptor plasticity A strategy was devised in order to uncover odorant triggered changes in protein regulation of the olfactory sensory neuron. Using Fluorescent 2-D Difference Gel Electrophoresis (DIGE) technology the olfactory epithelium from continuously and pulsed odorant (octanal) treated mice were compared to untreated control mice. In spite of the fact that treated mice displayed a similar phenotype, peripheral desensitization to octanal, different proteins were regulated in both groups suggesting discrete underlying molecular mechanisms. For example, in the continuously octanal treated group, the major peripheral desensitization factor was probably down-regulation of odorant binding proteins. In contrast, the pulsed group demonstrated more complex protein plasticity; no single dominant factor was evident. Desensitization may have been effectuated by modulation of synaptic plasticity via up-regulation of calcium binding proteins and cytoskeletal rearrangements. In addition, there is good experimental evidence that in both groups octanal was subjected to xenobiotic metabolism in sustentacular cells of the olfactory epithelium.

Using this powerful strategy several promising candidate proteins were identified including a potential odorant binding protein, a potential receptor chaperone and calcium binding proteins.

124

4.1.2 Spermatozoa membrane proteome Analysis of the sperm membrane proteome proved to be a formidable challenge. A refined strategy was empirically optimised, namely, the use of an affinity biphasic partition system in conjunction with powerful solubulisation and digestion conditions. This consisted of a potent enzymatic cocktail of lipase, trypsin and chymotrypsin in 60% methanol.

222 integral membrane proteins were identified, of which 57 were G-protein coupled receptors. A significant fraction of the GPCRs were chemoreceptors - 32 olfactory receptors and 3 taste receptors. Notably, most of the cleaved GPCR peptides were from transmembrane domains, attesting the effectiveness of this strategy at not only identifying chemoreceptors, but also highly hydrophobic receptors such as GPCRs.

4.1.3 The olfactory receptor proteome Clearly the functional roles of olfactory sensory neurons and spermatozoa are widely divergent. Nevertheless, with regard to the receptor proteome, one can no longer argue a distinct demarcation between both cell types. A staggering array of olfactory receptors are expressed in mature human sperm, not to mention a significant number of neuronal proteins. A clear relationship between the membrane proteome from both cell types is therefore evident and furthermore, it is likely that these similarities also include membrane associated proteins. For example, it is plausible that the sperm acrosomal reaction is based on similar mechanisms as described for synaptic exocytosis. It is also feasible that chemoreceptors play a role in sperm chemotaxis. What other molecular mechanisms may be shared between these two very different cells types?

In conclusion, using a proteomic approach interesting candidates were identified from the olfactory epithelium – these may shed light onto molecular events underpinning stimulus induced plasticity. Using a separate proteomic strategy, many previously unreported sperm membrane proteins were identified – these included an array of olfactory receptors.

It is suggested that these findings pave the way for functional investigations which could not only have immediate relevance to neuroscience, but also potentially to reproductive biology.

125

REFERENCES

A. Privat (2003). "Astrocytes as support for axonal regeneration in the central nervous system of mammals." Glia 43(1): 91-93. A. Stojanoff, H. B., A. G. Andrews, R. V. Hyne, (1987). "Isolation of a stable apical segment of the guinea pig sperm acrosome." Gamete Research 17(4): 321-332. Abraham L. Kierszenbaum (2000). "Fusion of membranes during the acrosome reaction: A tale of two SNAREs." Molecular Reproduction and Development 57(4): 309-310. Agopyan, N., J. Head, et al. (2004). "TRPV1 receptors mediate particulate matter-induced apoptosis." Am J Physiol Lung Cell Mol Physiol 286(3): L563-572. Alan D. Diamond, J. T. H. (1990). "Protein partitioning in PEG/dextran aqueous two-phase systems." AIChE Journal 36(7): 1017-1024. Alberti, S., S. M. Krause, et al. (2005). "Neuronal migration in the murine rostral migratory stream requires serum response factor." PNAS 102(17): 6148-6153. Alexander-Kaufman, K., G. James, et al. (2005). "Differential protein expression in the prefrontal white matter of human alcoholics: a proteomics study." Mol Psychiatry. Andersen, J. S., C. J. Wilkinson, et al. (2003). "Proteomic characterization of the human centrosome by protein correlation profiling." 426(6966): 570-574. Anderson, R. G. W. and K. Jacobson (2002). "A Role for Lipid Shells in Targeting Proteins to Caveolae, Rafts, and Other Lipid Domains." Science 296(5574): 1821-1825. Apfelbach, R. and E. Weiler (1985). "Olfactory deprivation enhances normal spine loss in the olfactory bulb of developing ferrets." Neurosci Lett 62(2): 169-73. Araneda, R. C., Z. Peterlin, et al. (2004). "A pharmacological profile of the aldehyde receptor repertoire in rat olfactory epithelium." J Physiol (Lond) 555(3): 743-756. Archer, S. K., C. A. Behm, et al. (2004). "The flightless I protein and the gelsolin family in nuclear hormone receptor- mediated signalling." Biochem Soc Trans 32(Pt 6): 940-2. Argon, Y. and B. B. Simen (1999). "GRP94, an ER chaperone with protein and peptide binding properties." Seminars in Cell and Developmental Biology 10(5): 495-505. Arienti, G., E. Carlini, et al. (2004). "Role of human prostasomes in the activation of spermatozoa." J Cell Mol Med 8(1): 77-84. Armbruster, B. N. and B. L. Roth (2005). "Mining the receptorome." J Biol Chem 280(7): 5129-32. Arnold, S. E., G. S. Smutzer, et al. (1998). "Cellular and Molecular Neuropathology of the Olfactory Epithelium and Central Olfactory Pathways in Alzheimer's Disease and Schizophrenia." Ann NY Acad Sci 855(1): 762-775. Atkins, C. M., J. C. Selcher, et al. (1998). "The MAPK cascade is required for mammalian associative learning." Nat Neurosci 1(7): 602-9. Aungst, J. L., P. M. Heyward, et al. (2003). "Centre-surround inhibition among olfactory bulb glomeruli." 426(6967): 623-629. Baba, K., M. Ikeda, et al. (1997). "Odor exposure reveals non-uniform expression profiles of c-Jun protein in rat olfactory bulb neurons." Brain Research 774(1-2): 142-148. Bailey, J. L., S. Tardif, et al. (2005). "Use of phosphoproteomics to study tyrosine kinase activity in capacitating boar sperm. Kinase activity and capacitation." Theriogenology 63(2): 599-614. Baker, M. A., R. Witherdin, et al. (2005). "Identification of post-translational modifications that occur during sperm maturation using difference in two-dimensional gel electrophoresis." Proteomics 5(4): 1003-12. Bando, Y., T. Katayama, et al. (2003). "GRP94 (94 kDa glucose-regulated protein) suppresses ischemic neuronal cell death against ischemia/reperfusion injury." Eur J Neurosci 18(4): 829-40. Banerjee, P., J. B. Joo, et al. (1995). "Differential solubilization of lipids along with membrane proteins by different classes of detergents." Chem Phys Lipids 77(1): 65-78. Bastianelli, E. and R. Pochet (1994). "Distribution of calmodulin, calbindin-D28k and calretinin among rat olfactory nerve bundles." Neurosci Lett 169(1-2): 223-6. Becamel, C., G. Alonso, et al. (2002). "Synaptic multiprotein complexes associated with 5-HT(2C) receptors: a proteomic approach." Embo J 21(10): 2332-42. Beijerinck, M. W. (1896). "Ueber une Eigentumlichkeit der loslichen Starke." Zbl. Bakt. 2: 627.

Beites, C. L., S. Kawauchi, et al. (2005). "Identification and molecular regulation of neural stem cells in the olfactory epithelium." Experimental Cell Research 306(2): 309-316. Bende, M. and S. Nordin (1997). "Perceptual Learning in Olfaction: Professional Wine Tasters versus Controls." Physiology & Behavior 62(5): 1065-1070. Beranova-Giorgianni, S., M. J. Pabst, et al. (2002). "Preliminary analysis of the mouse cerebellum proteome." Brain Res Mol Brain Res 98(1-2): 135-40. Bhandawat, V., J. Reisert, et al. (2005). "Elementary Response of Olfactory Receptor Neurons to Odorants." Science 308(5730): 1931-1934. Billing-Marczak, K. and J. Kuznicki (1999). "Calretinin--sensor or buffer--function still unclear." Pol J Pharmacol 51(2): 173-8. Blonder, J., M. B. Goshe, et al. (2002). "Enrichment of integral membrane proteins for proteomic analysis using liquid chromatography-tandem mass spectrometry." J Proteome Res 1(4): 351-60. Blonder, J., M. L. Hale, et al. (2004). "Proteomic analysis of detergent-resistant membrane rafts." Electrophoresis 25(9): 1307-18. Bohren, K. M., J. P. von Wartburg, et al. (1987). "Kinetics of carbonyl reductase from human brain." Biochem J 244(1): 165-71. Bohring, C., E. Krause, et al. (2001). "Isolation and identification of sperm membrane antigens recognized by antisperm antibodies, and their possible role in immunological infertility disease." Mol. Hum. Reprod. 7(2): 113-118. Bohring, C. and W. Krause (1999). "The characterization of human spermatozoa membrane proteins--surface antigens and immunological infertility." Electrophoresis 20(4-5): 971-6. Brachvogel, B., J. Dikschas, et al. (2003). "Annexin A5 is not essential for skeletal development." Mol. Cell. Biol. 23(8): 2907-2913. Bradley, J., J. Reisert, et al. (2005). "Regulation of cyclic nucleotide-gated channels." Current Opinion in Neurobiology 15(3): 343-349. Breer, H. (2003). "Olfactory receptors: molecular basis for recognition and discrimination of odors." Analytical and Bioanalytical Chemistry 377(3): 427-433. Breer, H. (2003). ": recognition and transduction of olfactory signals." Biochem Soc Trans 31(Pt 1): 113-6. Brunkhorst, A., M. Karlen, et al. (2005). "A specific role for the TFIID subunit TAF4 and RanBPM in neural progenitor differentiation." Molecular and Cellular Neuroscience 29(2): 250-258. Brunkhorst, A., T. Neuman, et al. (2004). "Novel isoforms of the TFIID subunit TAF4 modulate nuclear receptor- mediated transcriptional activity." Biochemical and Biophysical Research Communications 325(2): 574-579. Bruno P. Meloni, D. V. D., Rebecca Cole, Neville W. Knuckey, (2005). "Proteome analysis of cortical neuronal cultures following cycloheximide, heat stress and MK801 preconditioning." PROTEOMICS 9999(9999): NA. Buck, L. and R. Axel (1991). "A novel multigene family may encode odorant receptors: A molecular basis for odor recognition." Cell 65(1): 175-187. Bunai, K. and K. Yamane (2005). "Effectiveness and limitation of two-dimensional gel electrophoresis in bacterial membrane protein proteomics and perspectives." J Chromatogr B Analyt Technol Biomed Life Sci 815(1-2): 227-36. Buonviso, N. and M. Chaput (1999). "Olfactory experience decreases responsiveness of the olfactory bulb in the adult rat." Neuroscience 95(2): 325-332. Buonviso, N., R. Gervais, et al. (1998). Short lasting exposure to one odour decreases general reactivity in the olfactory bulb of adult rats. European Journal of Neuroscience. 10: 2472-2475. Cartwright, E. J., A. Cowin, et al. (1991). "Surface heterogeneity of bovine sperm revealed by aqueous two-phase partition." Biosci Rep 11(5): 265-73. Castellano, L. E., C. L. Trevino, et al. (2003). "Transient receptor potential (TRPC) channels in human sperm: expression, cellular localization and involvement in the regulation of flagellar motility." FEBS Lett 541(1-3): 69-74. Cavaggioni, A. and C. Mucignat-Caretta (2000). "Major urinary proteins, [alpha]2U-globulins and aphrodisin." Biochimica et Biophysica Acta (BBA) - Protein Structure and Molecular Enzymology 1482(1-2): 218-228. Chan, W. Y., C. L. Xia, et al. (2003). "Differential expression of S100 proteins in the developing human hippocampus and temporal cortex." Microsc Res Tech 60(6): 600-13. Chen, C. J. and J. G. Vandenbergh (1993). "Cocaine attenuates puberty acceleration in female house mice." Pharmacol Biochem Behav 44(2): 281-5. Chung-Liang Chien, T.-H. L., Kuo-Shyan Lu, (1998). "Distribution of neuronal intermediate filament proteins in the developing mouse olfactory system." Journal of Neuroscience Research 54(3): 353-363.

Clarke, E. J. and V. Allan (2002). "Intermediate Filaments: Vimentin Moves in." Current Biology 12(17): R596-R598. Colucci-Guyon, E., M. M. Portier, et al. (1994). "Mice lacking vimentin develop and reproduce without an obvious phenotype." Cell 79(4): 679-94. Coppens, A. G., R. Kiss, et al. (2001). "Immunolocalization of the calcium binding S100A1, S100A5 and S100A6 proteins in the dog cochlea during postnatal development." Brain Res Dev Brain Res 126(2): 191-9. Cross, N. L. (2004). "Reorganization of Lipid Rafts During Capacitation of Human Sperm." Biol Reprod 71(4): 1367- 1373. Cross, N. L. and J. W. Overstreet (1987). "Glycoconjugates of the human sperm surface: distribution and alterations that accompany capacitation in vitro." Gamete Res 16(1): 23-35. Dalton, P. (2000). "Psychophysical and Behavioral Characteristics of Olfactory Adaptation." Chem. Senses 25(4): 487- 492. Darszon, A., P. Labarca, et al. (1999). "Ion channels in sperm physiology." Physiol Rev 79(2): 481-510. Davis, R. L. (2004). "Olfactory Learning." Neuron 44(1): 31-48. Dawson, P. A., S. E. Steane, et al. (2005). "Impaired memory and olfactory performance in NaSi-1 sulphate transporter deficient mice." Behavioural Brain Research 159(1): 15-20. Dear, T., T. Boehm, et al. (1991). "Novel genes for potential ligand-binding proteins in subregions of the olfactory mucosa." EMBO J. 10(10): 2813-2819. Demange, P., D. Voges, et al. (1994). "Annexin V: the key to understanding ion selectivity and voltage regulation?" Trends Biochem Sci 19(7): 272-6. Descoteaux, A., H. A. Avila, et al. (2002). "LeishmaniaLPG3 encodes a GRP94 homolog required for phosphoglycan synthesis implicated in parasite virulence but not viability." EMBO J. 21(17): 4458-4469. Dewi W. Owens, E. B. L. (2003). "The quest for the function of simple epithelial keratins." BioEssays 25(8): 748-758. Dietz, S. B. and V. N. Murthy (2005). "Contrasting short-term plasticity at two sides of the mitral-granule reciprocal synapse in the mammalian olfactory bulb." J Physiol (Lond) 569(2): 475-488. Dirk Claeys, K. G., Beat J. Meyer, (2005). "Two-dimensional Blue Native/sodium dodecyl sulfate gel electrophoresis for analysis of multimeric proteins in platelets." Electrophoresis 26(6): 1189-1199. Dollins, D. E., R. M. Immormino, et al. (2005). "Structure of Unliganded GRP94, the Endoplasmic Reticulum Hsp90: Basis for nucleotide-induced conformational change." J. Biol. Chem. 280(34): 30438-30447. Domagala, A. and M. Kurpisz (2004). "Identification of sperm immunoreactive antigens for immunocontraceptive purposes: a review." Reproductive Biology and Endocrinology 2(1): 11. Dorothy G. Flood, J. P. F., Kevin M. Walton, Craig A. Dionne, Patricia C. Contreras, Matthew S. Miller, Ratan V. Bhat, (1998). "Immunolocalization of the mitogen-activated protein kinases p42 and JNK1, and their regulatory kinases MEK1 and MEK4, in adult rat central nervous system." The Journal of Comparative Neurology 398(3): 373-392. Doving, K. B. and A. J. Pinching (1973). "Selective degeneration of neurons in the olfactory bulb following prolonged odour exposure." Brain Res 52: 115-29. Edwards, R. G., B. D. Bavister, et al. (1969). "Early stages of fertilization in vitro of human oocytes matured in vitro." Nature 221(181): 632-5. Eichacker, L. A., B. Granvogl, et al. (2004). "Hiding behind Hydrophobicity: Transmembrane Segments In Mass Spectrometry." J. Biol. Chem. 279(49): 50915-50922. Eisenbach, M. (1999). "Mammalian sperm chemotaxis and its association with capacitation." Dev Genet 25(2): 87-94. Eisenbach, M. (2004). "Towards understanding the molecular mechanism of sperm chemotaxis." J Gen Physiol 124(2): 105-8. Emsley, J. G., B. D. Mitchell, et al. (2005). "Adult neurogenesis and repair of the adult CNS with neural progenitors, precursors, and stem cells." Progress in Neurobiology 75(5): 321-341. Eng, J. K., McCormack, A. L., Yates, J. R., III, (1994). J. Am. Soc. Mass Spectrom 5: 976-989. Farr, C. D., P. R. Gafken, et al. (2004). "Proteomic analysis of native metabotropic glutamate receptor 5 protein complexes reveals novel molecular constituents." Journal of Neurochemistry 91(2): 438-450. Fausther, M., L. Villeneuve, et al. (2004). "Heat shock protein 70 expression, keratin phosphorylation and Mallory body formation in hepatocytes from griseofulvin-intoxicated mice." Comp Hepatol 3(1): 5. Felinski, E. A. and P. G. Quinn (1999). "The CREB Constitutive Activation Domain Interacts with TATA-binding Protein-associated Factor 110 (TAF110) through Specific Hydrophobic Residues in One of the Three Subdomains Required for Both Activation and TAF110 Binding." J. Biol. Chem. 274(17): 11672-11678. Felmy, F. and R. Schneggenburger (2004). "Developmental expression of the Ca2+-binding proteins calretinin and parvalbumin at the calyx of Held of rats and mice." European Journal of Neuroscience 20(6): 1473-1482.

Feng, G., X.-c. Xu, et al. (2001). "Diminished Expression of S100A2, a Putative Tumor Suppressor, at Early Stage of Human Lung Carcinogenesis." Cancer Res 61(21): 7999-8004. Fenn, J. B., M. Mann, et al. (1989). "Electrospray ionization for mass spectrometry of large biomolecules." Science 246(4926): 64-71. Ficarro, S., O. Chertihin, et al. (2003). "Phosphoproteome analysis of capacitated human sperm. Evidence of tyrosine phosphorylation of a kinase-anchoring protein 3 and valosin-containing protein/p97 during capacitation." J Biol Chem 278(13): 11579-89. Ficarro, S. B., M. L. McCleland, et al. (2002). "Phosphoproteome analysis by mass spectrometry and its application to Saccharomyces cerevisiae." Nat Biotechnol 20(3): 301-5. Fischer, F., D. Wolters, et al. (2005). "Towards the complete membrane proteome: High coverage of integral membrane proteins through transmembrane peptide detection." Mol Cell Proteomics: M500234-MCP200. Flanagan, S. D. and S. H. Barondes (1975). "Affinity partitioning. A method for purification of proteins using specific polymer-ligands in aqueous polymer two-phase systems." J Biol Chem 250(4): 1484-9. Flesch, F. M., W. F. Voorhout, et al. (1998). "Use of Lectins to Characterize Plasma Membrane Preparations from Boar Spermatozoa: A Novel Technique for Monitoring Membrane Purity and Quantity." Biol Reprod 59(6): 1530- 1539. Fletcher, M. L. and D. A. Wilson (2003). "Olfactory Bulb Mitral- Plasticity: Odorant-Specific Tuning Reflects Previous Odorant Exposure." J. Neurosci. 23(17): 6946-6955. Flower, D. R. (1996). "The lipocalin protein family: structure and function." Biochem J 318 ( Pt 1): 1-14. Forrest, G. L. and B. Gonzalez (2000). "Carbonyl reductase." Chemico-Biological Interactions 129(1-2): 21-40. Franziska F. Wiebel, V. R., Kristina Vintersten, Alfred Nordheim, (2002). "Generation of mice carrying conditional knockout alleles for the transcription factor SRF." genesis 32(2): 124-126. Frazier, L. L. and P. C. Brunjes (1988). "Unilateral odor deprivation: early postnatal changes in olfactory bulb cell density and number." J Comp Neurol 269(3): 355-70. Friso, G. and L. Wikstrom (1999). "Analysis of proteins from membrane-enriched cerebellar preparations by two- dimensional gel electrophoresis and mass spectrometry." Electrophoresis 20(4-5): 917-27. Fujii, J., T. Myint, et al. (1995). "Characterization of Wild-Type and Amyotrophic Lateral Sclerosis-Related Mutant Cu,Zn-Superoxide Dismutases Overproduced in Baculovirus-Infected Insect Cells." Journal of Neurochemistry 64(4): 1456-1461. Gall, W. E., C. F. Millette, et al. (1975). "Chemical dissection of mammalian spermatozoa." Acta Endocrinol Suppl (Copenh) 194: 154-72. Gatlin, C. L., G. R. Kleemann, et al. (1998). "Protein Identification at the Low Femtomole Level from Silver-Stained Gels Using a New Fritless Electrospray Interface for Liquid Chromatography-Microspray and Nanospray Mass Spectrometry,." Analytical Biochemistry 263(1): 93-101. Gerke, V., C. E. Creutz, et al. (2005). "Annexins: Linking Ca2+ Signalling To Membrane Dynamics." Nature Reviews Molecular Cell Biology 6(6): 449-461. Gharahdaghi, F., C. R. Weinberg, et al. (1999). "Mass spectrometric identification of proteins from silver-stained polyacrylamide gel: a method for the removal of silver ions to enhance sensitivity." Electrophoresis 20(3): 601- 5. Ghosh, D., M. Sawicki, et al. (2001). "Porcine Carbonyl Reductase. Structural basis for a functional monomer in short chain dehydrogenases/reductases." J. Biol. Chem. 276(21): 18457-18463. Glass, R., M. Bardini, et al. (2001). "Expression of nucleotide P2X receptor subtypes during spermatogenesis in the adult rat testis." Cells Tissues Organs 169(4): 377-87. Glusman, G., A. Bahar, et al. (2000). "The olfactory receptor gene superfamily: data mining, classification, and nomenclature." Mamm Genome 11(11): 1016-23. Goldman, A. L., W. Van der Goes van Naters, et al. (2005). "Coexpression of Two Functional Odor Receptors in One Neuron." Neuron 45(5): 661-666. Gomez, G., F. W. Lischka, et al. (2005). "Evidence for Multiple Calcium Response Mechanisms in Mammalian Olfactory Receptor Neurons." Chem. Senses 30(4): 317-326. Gorg, A., C. Obermaier, et al. (1997). "Very alkaline immobilized pH gradients for two-dimensional electrophoresis of ribosomal and nuclear proteins." Electrophoresis 18(3-4): 328-37. Goshe, M. B., J. Blonder, et al. (2003). "Affinity labeling of highly hydrophobic integral membrane proteins for proteome-wide analysis." J Proteome Res 2(2): 153-61. Graumann, J., L. A. Dunipace, et al. (2004). "Applicability of tandem affinity purification MudPIT to pathway proteomics in yeast." Mol Cell Proteomics 3(3): 226-37.

Gudermann, T. (2001). "Multiple pathways of ERK activation by G protein-coupled receptors." Novartis Found Symp 239: 68-79; discussion 80-4, 150-9. Guthrie, K., J. Rayhanabad, et al. (2000). "Odors regulate Arc expression in neuronal ensembles engaged in odor processing." Neuroreport 11(9): 1809-13. Guthrie, K. M. and C. M. Gall (1995). "Odors increase Fos in olfactory bulb neurons including dopaminergic cells." Neuroreport 6(16): 2145-9; discussion 2103. Gye Sun Jeon, S. W. P., Dong Woon Kim, Je Hoon Seo, Jaeyoung Cho, So Young Lim, Seong Deok Kim, Sa Sun Cho, (2004). "Glial expression of the 90-kDa heat shock protein (HSP90) and the 94-kDa glucose-regulated protein (GRP94) following an excitotoxic lesion in the mouse hippocampus." Glia 48(3): 250-258. Hamil, K., Q. Liu, et al. (2003). "LCN6, a novel human epididymal lipocalin." Reproductive Biology and Endocrinology 1(1): 112. Han, D. K., J. Eng, et al. (2001). "Quantitative profiling of differentiation-induced microsomal proteins using isotope- coded affinity tags and mass spectrometry." 19(10): 946-951. Han, J. and K. L. Schey (2004). "Proteolysis and mass spectrometric analysis of an integral membrane: aquaporin 0." J Proteome Res 3(4): 807-12. Han, S., K. H. Zhang, et al. (2004). "Effects of annexins II and V on survival of neurons and astrocytes in vitro." Acta Pharmacol Sin 25(5): 602-10. Hao, Z., M. J. Wolkowicz, et al. (2002). "SAMP32, a Testis-Specific, Isoantigenic Sperm Acrosomal Membrane- Associated Protein." Biol Reprod 66(3): 735-744. Harris, B. Z. and W. A. Lim (2001). "Mechanism and role of PDZ domains in signaling complex assembly." J Cell Sci 114(Pt 18): 3219-31. Hartinger, J., K. Stenius, et al. (1996). "16-BAC/SDS-PAGE: a two-dimensional gel electrophoresis system suitable for the separation of integral membrane proteins." Anal Biochem 240(1): 126-33. Hayes, M. J., U. Rescher, et al. (2004). "Annexin-Actin Interactions." Traffic 5(8): 571-576. He, X. B., J. H. Hu, et al. (2001). "Identification of GABA(B) receptor in rat testis and sperm." Biochem Biophys Res Commun 283(1): 243-7. Heike Schaefer, J.-P. C., Christian Bunse, Cornelia Joppich, Helmut E. Meyer, Katrin Marcus, (2004). "A peptide preconcentration approach for nano-high-performance liquid chromatography to diminish memory effects." Proteomics 4(9): 2541-2544. Hildebrand, J. G. and G. M. Shepherd (1997). "Mechanisms of olfactory discrimination: converging evidence for common principles across phyla." Annu Rev Neurosci 20: 595-631. Hisatake, K., S. Hasegawa, et al. (1993). "The p250 subunit of native TATA box-binding factor TFIID is the cell-cycle regulatory protein CCG1." 362(6416): 179-181. Hofmeister, F. (1888). "Zur Lehre von der Wirkung der Salze." Arch. Exp. Pathol. Pharmakol. (Leipzig) 24: 247-260. Holle, G. E. (2004). "The influence of neural signal transduction on EEC gene expression under consideration of chromatin, following myenteric ablation (review)." Int J Mol Med 14(4): 497-504. Holt, W. V. and K. J. W. Van Look (2004). "Concepts in sperm heterogeneity, sperm selection and sperm competition as biological foundations for laboratory tests of semen quality." Reproduction 127(5): 527-535. Hoyer, D., J. P. Hannon, et al. (2002). "Molecular, pharmacological and functional diversity of 5-HT receptors." Pharmacol Biochem Behav 71(4): 533-54. Hsu, C.-L., Y.-L. Chen, et al. (2005). "Androgen Receptor (AR) NH2- and COOH-Terminal Interactions Result in the Differential Influences on the AR-Mediated Transactivation and Cell Growth." Mol Endocrinol 19(2): 350- 361. Hughes, J., C. J. Ward, et al. (1999). "Identification of a human homologue of the sea urchin receptor for egg jelly: a polycystic kidney disease-like protein." Hum Mol Genet 8(3): 543-9. Hummel, T., H. Guel, et al. (2004). "Olfactory Sensitivity of Subjects Working in Odorous Environments." Chem. Senses 29(6): 533-536. Husi, H., M. A. Ward, et al. (2000). "Proteomic analysis of NMDA receptor-adhesion protein signaling complexes." 3(7): 661-669. Hutt, D. M., J. M. Baltz, et al. (2005). "Synaptotagmin VI and VIII and Syntaxin 2 Are Essential for the Mouse Sperm Acrosome Reaction." J Biol Chem 280(21): 20197-203. Hutt, D. M., J. M. Baltz, et al. (2005). "Synaptotagmin VI and VIII and Syntaxin 2 Are Essential for the Mouse Sperm Acrosome Reaction." J. Biol. Chem. 280(21): 20197-20203. Iborra, A., C. Morte, et al. (1996). "Human sperm coating antigen from seminal plasma origin." Am J Reprod Immunol 36(2): 118-25.

Inouye, S., M. Seo, et al. (1998). "Increase of Adenylate Kinase Isozyme 1 Protein During Neuronal Differentiation in Mouse Embryonal Carcinoma P19 Cells and in Rat Brain Primary Cultured Cells." Journal of Neurochemistry 71(1): 125-133. Isam Rais, M. K., Hermann Schägger, (2004). "Two-dimensional electrophoresis for the isolation of integral membrane proteins and mass spectrometric identification." Proteomics 4(9): 2567-2571. Jackson, M. L., C. F. Schmidt, et al. (1982). "Solubilization of phosphatidylcholine bilayers by octyl glucoside." Biochemistry 21(19): 4576-82. Janssen, E., P. P. Dzeja, et al. (2000). "Adenylate kinase 1 gene deletion disrupts muscle energetic economy despite metabolic rearrangement." EMBO J. 19(23): 6371-6381. Jin, K., X. O. Mao, et al. (2004). "Proteomic and immunochemical characterization of a role for stathmin in adult neurogenesis." Faseb J 18(2): 287-99. Jiuhong Kang, Y. S., Bin Xiang, Bin Qu, Wenjuan Su, Min Zhu, Min Zhang, Guobin Bao, Feifei Wang, Xiaoqing Zhang, Rongxi Yang, Fengjuan Fan, Xiaoqing Chen, Gang Pei, and Lan Ma (2005). "A Nuclear Function of ß-Arrestin1 in GPCR Signaling: Regulation of Histone Acetylation and Gene Transcription." Cell 123: 833- 847. Johansson, G., R. Gysin, et al. (1981). "Affinity partitioning of membranes. Evidence for discrete membrane domains containing cholinergic receptor." J Biol Chem 256(17): 9126-35. Johnson, B. A., H. Farahbod, et al. (2004). "Local and global chemotopic organization: general features of the glomerular representations of aliphatic odorants differing in carbon number." J Comp Neurol 480(2): 234-49. Kaluza, J. and H. Breer (2000). "Responsiveness of olfactory neurons to distinct aliphatic aldehydes." J Exp Biol 203(5): 927-933. Karas, M. and F. Hillenkamp (1988). "Laser desorption ionization of proteins with molecular masses exceeding 10,000 daltons." Anal Chem 60(20): 2299-301. Katada, S., T. Hirokawa, et al. (2005). "Structural Basis for a Broad But Selective Ligand Spectrum of a Mouse Olfactory Receptor: Mapping the Odorant-Binding Site." J. Neurosci. 25(7): 1806-1815. Kaul, S. C., M. Matsui, et al. (1997). "Expression Analysis of Mortalin, a Unique Member of the Hsp70 Family of Proteins, in Rat Tissues." Experimental Cell Research 232(1): 56-63. Kaul, S. C., K. Taira, et al. (2002). "Mortalin: present and prospective." Experimental Gerontology 37(10-11): 1157- 1164. Kaupp, U. B., J. Solzin, et al. (2003). "The signal flow and motor response controling chemotaxis of sea urchin sperm." Nat Cell Biol 5(2): 109-17. Kearse, K. and G. Hart (1991). "Lymphocyte Activation Induces Rapid Changes in Nuclear and Cytoplasmic Glycoproteins." PNAS 88(5): 1701-1705. Kharen L. Doyle, M. K., Anne M. Cunningham, (2001). "Expression of the intermediate filament protein nestin by sustentacular cells in mature olfactory neuroepithelium." The Journal of Comparative Neurology 437(2): 186- 195. Kietselaer, B. L., L. Hofstra, et al. (2003). "The role of labeled Annexin A5 in imaging of programmed cell death. From animal to clinical imaging." Q J Nucl Med 47(4): 349-61. Kim, M., L. H. Jiang, et al. (2001). "Proteomic and functional evidence for a P2X7 receptor signalling complex." Embo J 20(22): 6347-58. Kishimoto, J., E. B. Keverne, et al. (1993). "Calretinin, calbindin-D28k and parvalbumin-like immunoreactivity in mouse chemoreceptor neurons." Brain Res 610(2): 325-9. Klofutar, D. R.-T. a. C. (2003). "Osmotic coefficients and solvation thermodynamics of aqueous solutions of some lower poly(ethylene glycol)s at different temperatures." J. Mol. Liquids 103-104: 187-200. Klose, J. (1975). "Protein mapping by combined isoelectric focusing and electrophoresis of mouse tissues. A novel approach to testing for induced point mutations in mammals." Humangenetik 26(3): 231-43. Klose, J. and U. Kobalz (1995). "Two-dimensional electrophoresis of proteins: an updated protocol and implications for a functional analysis of the genome." Electrophoresis 16(6): 1034-59. Knecht, M. and T. Hummel (2004). "Recording of the human electro-olfactogram." Physiol Behav 83(1): 13-9. Knowlton, M. N., B. M. C. Chan, et al. (2003). "The zebrafish band 4.1 member Mir is involved in cell movements associated with gastrulation." Developmental Biology 264(2): 407-429. Kobori, H., S. Miyazaki, et al. (2000). "Characterization of Intracellular Ca2+ Increase in Response to Progesterone and Cyclic Nucleotides in Mouse Spermatozoa." Biol Reprod 63(1): 113-120. Kopsidas, G., S. A. Kovalenko, et al. (2000). "Tissue mitochondrial DNA changes. A stochastic system." Ann N Y Acad Sci 908: 226-43.

Kragh-Hansen, U., M. le Maire, et al. (1998). "The Mechanism of Detergent Solubilization of Liposomes and Protein- Containing Membranes." Biophys. J. 75(6): 2932-2946. Krasnoperov, V., Y. Lu, et al. (2002). "Post-translational Proteolytic Processing of the Calcium-independent Receptor of alpha -Latrotoxin (CIRL), a Natural Chimera of the Cell Adhesion Protein and the G Protein-coupled Receptor. Role of the G protein-coupled receptor proteolysis site(GPS) motif." J. Biol. Chem. 277(48): 46518- 46526. Krautwurst, D., K. W. Yau, et al. (1998). "Identification of ligands for olfactory receptors by functional expression of a receptor library." Cell 95(7): 917-26. Kroeze, W. K., D. J. Sheffler, et al. (2003). "G-protein-coupled receptors at a glance." J Cell Sci 116(24): 4867-4869. Ku, N.-O. and M. B. Omary (1994). "Expression, Glycosylation, and Phosphorylation of Human Keratins 8 and 18 in Insect Cells." Experimental Cell Research 211(1): 24-35. Ku, N.-O. and M. B. Omary (1995). "Identification and Mutational Analysis of the Glycosylation Sites of Human Keratin 18." J. Biol. Chem. 270(20): 11820-11827. Kulesh, D. A., G. Cecena, et al. (1989). "Posttranslational regulation of keratins: degradation of mouse and human keratins 18 and 8." Mol Cell Biol 9(4): 1553-65. Kyte, J. and R. F. Doolittle (1982). "A simple method for displaying the hydropathic character of a protein." J Mol Biol 157(1): 105-32. Lacazette, E., A.-M. Gachon, et al. (2000). "A novel human odorant-binding protein gene family resulting from genomic duplicons at 9q34: differential expression in the oral and genital spheres." Hum. Mol. Genet. 9(2): 289-301. Laemmli, U. K. (1970). "Cleavage of structural proteins during the assembly of the head of bacteriophage T4." Nature 227(5259): 680-5. Lagier, S., A. Carleton, et al. (2004). "Interplay between Local GABAergic Interneurons and Relay Neurons Generates {gamma} Oscillations in the Rat Olfactory Bulb." J. Neurosci. 24(18): 4382-4392. Laurent, G. (2002). "Olfactory network dynamics and the coding of multidimensional signals." Nature Reviews Neuroscience 3(11): 884-895. Laurent, G., M. Stopfer, et al. (2001). "Odor encoding as an active, dynamical process: Experiments, Computation, and Theory." Annual Review of Neuroscience 24(1): 263-297. Lee, H. Y., M. Bardini, et al. (2000). "P2X receptor immunoreactivity in the male genital organs of the rat." Cell and Tissue Research 300(2): 321-330. Lefkowitz, R. J. (2000). "The superfamily of heptahelical receptors." Nat Cell Biol 2(7): E133-6. Legrand, C., C. Ferraz, et al. (1991). "Distribution of gelsolin in the retina of the developing rabbit." Cell Tissue Res 264(2): 335-8. Leinders-Zufall, T., A. P. Lane, et al. (2000). "Ultrasensitive pheromone detection by mammalian vomeronasal neurons." Nature 405(6788): 792-796. Lemon, C. H. and D. V. Smith (2005). "Neural Representation of Bitter Taste in the Nucleus of the Solitary Tract." J Neurophysiol 94(6): 3719-3729. Ling, G., J. Gu, et al. (2004). "Regulation of cytochrome P450 gene expression in the olfactory mucosa." Chemico- Biological Interactions 147(3): 247-258. Lledo, P. M., G. Gheusi, et al. (2005). "Information processing in the mammalian olfactory system." Physiol Rev 85(1): 281-317. Lobel, D., J. Strotmann, et al. (2001). "Identification of a Third Rat Odorant-binding Protein (OBP3)." Chem. Senses 26(6): 673-680. Lu, H., M. Hesse, et al. (2005). "Type II keratins precede type I keratins during early embryonic development." European Journal of Cell Biology 84(8): 709-718. Lu, M., F. Echeverri, et al. (2003). "Endoplasmic Reticulum Retention, Degradation, and Aggregation of Olfactory G- Protein Coupled Receptors." Traffic 4(6): 416-433. Lucas, P., K. Ukhanov, et al. (2003). "A Diacylglycerol-Gated Cation Channel in Vomeronasal Neuron Dendrites Is Impaired in TRPC2 Mutant Mice: Mechanism of Pheromone Transduction." Neuron 40(3): 551-561. M. Montag-Sallaz, N. B. (2002). "Altered odor-induced expression of c-fos and arg 3.1 immediate early genes in the olfactory system after familiarization with an odor." Journal of Neurobiology 52(1): 61-72. Macfarlane, D. E. (1989). "Two dimensional benzyldimethyl-n-hexadecylammonium chloride----sodium dodecyl sulfate preparative polyacrylamide gel electrophoresis: a high capacity high resolution technique for the purification of proteins from complex mixtures." Anal Biochem 176(2): 457-63.

Makio Utsumi, K. O., Yoshiaki Kawasaki, Manabu Tamura, Takeshi Kubo, Masaya Tohyama, (1999). "Expression of major urinary protein genes in the nasal glands associated with general olfaction." Journal of Neurobiology 39(2): 227-236. Marouga, R., S. David, et al. (2005). "The development of the DIGE system: 2D fluorescence difference gel analysis technology." Anal Bioanal Chem 382(3): 669-78. Martinez, P. and A. Morros (1996). "Membrane lipid dynamics during human sperm capacitation." Front Biosci 1: d103- 17. Mathews, C. K. v. H., K.E. (1990). Biochemistry. Redwood City, The Benjamin/Cummings Publishing Company, Inc. Mawuenyega, K. G., H. Kaji, et al. (2003). "Large-Scale Identification of Caenorhabditis elegans proteins by Multidimensional Liquid Chromatography-Tandem Mass Spectrometry." J. Proteome Res. 2(1): 23-35. McDonald, P. H., C.-W. Chow, et al. (2000). "beta -Arrestin 2: A Receptor-Regulated MAPK Scaffold for the Activation of JNK3." Science 290(5496): 1574-1577. McDonald, P. H. and R. J. Lefkowitz (2001). "[beta]Arrestins: New roles in regulating heptahelical receptors' functions." Cellular Signalling 13(10): 683-689. McGough, A. M., C. J. Staiger, et al. (2003). "The gelsolin family of actin regulatory proteins: modular structures, versatile functions." FEBS Letters 552(2-3): 75-81. Mashukova M A, Hatt H and Neuhaus EM. "ß-arrestin2-mediated desensitization of olfactory receptors." Submitted Mechref, Y., W. Ma, et al. (1999). "N-Linked Oligosaccharides of Vomeromodulin, a Putative Pheromone Transporter in Rat." Biochemical and Biophysical Research Communications 255(2): 451-455. Meizel, S. (2004). "The sperm, a neuron with a tail: 'neuronal' receptors in mammalian sperm." Biol Rev Camb Philos Soc 79(4): 713-32. Menini, A. (1999). "Calcium signalling and regulation in olfactory neurons." Curr Opin Neurobiol 9(4): 419-26. Meur, S. K., R. L. Dhoble, et al. (1988). "Acrosome separation from buffalo spermatozoa by chemical dissection." Indian J Exp Biol 26(8): 586-8. Michael Hamacher, K. M., Christian Stephan, Andre van Hall, Helmut E. Meyer, (2005). "HUPO BPP Workshop on Mouse Models for Neurodegeneration - Choosing the right models." Proteomics 5(14): 3558-3559. Millette, C. F., P. G. Spear, et al. (1973). "Chemical dissection of mammalian spermatozoa." J. Cell Biol. 58(3): 662-675. Mitaku, S. and T. Hirokawa (1999). "Physicochemical factors for discriminating between soluble and membrane proteins: hydrophobicity of helical segments and protein length." Protein Eng. 12(11): 953-957. Molloy, M. P. (2000). "Two-dimensional electrophoresis of membrane proteins using immobilized pH gradients." Anal Biochem 280(1): 1-10. Mombaerts, P. (2004). "Odorant receptor gene choice in olfactory sensory neurons: the one receptor-one neuron hypothesis revisited." Current Opinion in Neurobiology 14(1): 31-36. Moohan, J. M. and K. S. Lindsay (1995). "Spermatozoa selected by a discontinuous Percoll density gradient exhibit better motion characteristics, more hyperactivation, and longer survival than direct swim-up." Fertil Steril 64(1): 160-5. Morgan, R. O., S. Martin-Almedina, et al. (2004). "Evolutionary perspective on annexin calcium-binding domains." Biochimica et Biophysica Acta (BBA) - Molecular Cell Research 1742(1-3): 133-140. Mori, K., H. Nagao, et al. (1999). "The olfactory bulb: coding and processing of odor molecule information." Science 286(5440): 711-5. Mori, K., H. von Campenhaus, et al. (2000). "Zonal organization of the mammalian main and accessory olfactory systems." Philos Trans R Soc Lond B Biol Sci 355(1404): 1801-12. Morrison, R. S., Y. Kinoshita, et al. (2002). "Proteomic analysis in the neurosciences." Mol Cell Proteomics 1(8): 553-60. Mulryan, K., D. P. Gitterman, et al. (2000). "Reduced vas deferens contraction and male infertility in mice lacking P2X1 receptors." Nature 403(6765): 86-89. Naaby-Hansen, S., C. J. Flickinger, et al. (1997). "Two-dimensional gel electrophoretic analysis of vectorially labeled surface proteins of human spermatozoa." Biol Reprod 56(3): 771-87. Nadja B. Cech, C. G. E. (2001). "Practical implications of some recent studies in electrospray ionization fundamentals." Mass Spectrometry Reviews 20(6): 362-387. Nakamura, T. (2000). "Cellular and molecular constituents of olfactory sensation in vertebrates." Comparative Biochemistry and Physiology - Part A: Molecular & Integrative Physiology 126(1): 17-32. Nan Li, A. R. E. S., Nan Zhang, Allan Mak, Liang Li, (2004). "Lipid raft proteomics: Analysis of in-solution digest of sodium dodecyl sulfate-solubilized lipid raft proteins by liquid chromatography-matrix-assisted laser desorption/ionization tandem mass spectrometry." Proteomics 4(10): 3156-3166.

Nelson, M. R., E. Thulin, et al. (2002). "The EF-hand domain: A globally cooperative structural unit." Protein Sci 11(2): 198-205. Nesterenko, M. V., M. Tilley, et al. (1994). "A simple modification of Blum's silver stain method allows for 30 minute detection of proteins in polyacrylamide gels." J Biochem Biophys Methods 28(3): 239-42. Neuhaus EM, Mashukova A , Barbour DJ, Walters D and Hatt H. "Novel function ß-arrestin2 in human sperm". Submitted Neuhaus EM, Mashukova A, Zhang W, Barbour DJ and Hatt H. "A specific heat shock protein enhances expression of odorant receptor proteins." Submitted Nicchitta, C. V., D. M. Carrick, et al. (2004). "The messenger and the message: gp96 (GRP94)-peptide interactions in cellular immunity." Cell Stress Chaperones 9(4): 325-31. Nicholson-Dykstra, S., H. N. Higgs, et al. (2005). "Actin Dynamics: Growth from Dendritic Branches." Current Biology 15(9): R346-R357. Norman, C., M. Runswick, et al. (1988). "Isolation and properties of cDNA clones encoding SRF, a transcription factor that binds to the c-fos serum response element." Cell 55(6): 989-1003. North, R. A. (2002). "Molecular physiology of P2X receptors." Physiol Rev 82(4): 1013-67. Novotny, M. V. (2003). "Pheromones, binding proteins and receptor responses in rodents." Biochem Soc Trans 31(Pt 1): 117-22. Novotny, M. V., W. Ma, et al. (1999). "Positive identification of the puberty-accelerating pheromone of the house mouse: the volatile ligands associating with the major urinary protein." Proc Biol Sci 266(1432): 2017-22. O'Farrell, P. H. (1975). "High resolution two-dimensional electrophoresis of proteins." J. Biol. Chem. 250(10): 4007- 4021. Oh, P., Y. Li, et al. (2004). "Subtractive proteomic mapping of the endothelial surface in lung and solid tumours for tissue-specific therapy." 429(6992): 629-635. Ohno, K., Y. Kawasaki, et al. (1996). "Differential expression of odorant-binding protein genes in rat nasal glands: Implications for odorant-binding proteinII as a possible pheromone transporter." Neuroscience 71(2): 355- 366. Ollero, M., M. L. Pascual, et al. (1994). "Revealing surface changes associated with maturation of ram spermatozoa by centrifugal counter-current distribution in an aqueous two-phase system." J Chromatogr A 668(1): 173-8. Oritani, H., Y. Deyashiki, et al. (1992). "Purification and characterization of pig lung carbonyl reductase." Arch Biochem Biophys 292(2): 539-47. Osheroff, J. E., P. E. Visconti, et al. (1999). "Regulation of human sperm capacitation by a cholesterol efflux-stimulated signal transduction pathway leading to protein kinase A-mediated up-regulation of protein tyrosine phosphorylation." Mol Hum Reprod 5(11): 1017-26. Othman, M. M., K. M. Klueber, et al. (2003). "Identification and culture of olfactory neural progenitors from GFP mice." Biotech Histochem 78(2): 57-70. Padmanabhan, B., T. Kuzuhara, et al. (2004). "The Crystal Structure of CCG1/TAFII250-interacting Factor B (CIB)." J. Biol. Chem. 279(10): 9615-9624. Padmanabhan, B., T. Kuzuhara, et al. (2000). "Purification, crystallization and preliminary X-ray crystallographic analysis of human CCG1-interacting factor B." Acta Crystallogr D Biol Crystallogr 56 ( Pt 11): 1479-81. Panhuber, H., A. Mackay-Sim, et al. (1987). "Prolonged odor exposure causes severe cell shrinkage in the adult rat olfactory bulb." Brain Res 428(2): 307-11. Park, S.-K., S. R. Shanbhag, et al. (2000). "Expression patterns of two putative odorant-binding proteins in the olfactory organs of Drosophila melanogaster have different implications for their functions." Cell and Tissue Research 300(1): 181-192. Parmentier, M., F. Libert, et al. (1992). "Expression of members of the putative olfactory receptor gene family in mammalian germ cells." 355(6359): 453-455. Persson, A. and B. Jergil (1995). "The purification of membranes by affinity partitioning." Faseb J 9(13): 1304-10. Pes, D., M. Mameli, et al. (1998). "Cloning and expression of odorant-binding proteins Ia and Ib from mouse nasal tissue." Gene 212(1): 49-55. Piast, M., I. Kustrzeba-Wojcicka, et al. (2005). "Molecular evolution of enolase." Acta Biochim Pol 52(2): 507-13. Pixton, K. L., E. D. Deeks, et al. (2004). "Sperm proteome mapping of a patient who experienced failed fertilization at IVF reveals altered expression of at least 20 proteins compared with fertile donors: Case report." Hum. Reprod. 19(6): 1438-1447. Poon, H. F., R. A. Vaishnav, et al. (2005). "Proteomic identification of differentially expressed proteins in the aging murine olfactory system and transcriptional analysis of the associated genes." J Neurochem 94(2): 380-92.

Porteros, A., R. Arevalo, et al. (1997). "Calretinin immunoreactivity in the developing olfactory system of the rainbow trout." Brain Res Dev Brain Res 100(1): 101-9. Pouyssegur, J. (2000). "Signal Treansduction: An Arresting Start for MAPK." Science 290(5496): 1515-1518. Rajeev, S. K. and K. V. R. Reddy (2004). "Sperm membrane protein profiles of fertile and infertile men: identification and characterization of fertility-associated sperm antigen." Hum. Reprod. 19(2): 234-242. Rao, J., J. C. Herr, et al. (2003). "Cloning and Characterization of a Novel Sperm-Associated Isoantigen (E-3) with Defensin- and Lectin-Like Motifs Expressed in Rat Epididymis." Biol Reprod 68(1): 290-301. Rebois, R. V. and T. E. Hebert (2003). "Protein complexes involved in heptahelical receptor-mediated signal transduction." Receptors Channels 9(3): 169-94. Rehn, B., H. Panhuber, et al. (1988). "Spine density on olfactory granule cell dendrites is reduced in rats reared in a restricted olfactory environment." Brain Res 468(1): 143-7. René-Peiman Zahedi, C. M., Albert Sickmann, (2005). "Two-dimensional benzyldimethyl-n-hexadecylammonium chloride/SDS-PAGE for membrane proteomics." Proteomics 5(14): 3581-3588. Reutelingsperger, C. P. M., E. Dumont, et al. (2002). "Visualization of cell death in vivo with the annexin A5 imaging protocol." Journal of Immunological Methods 265(1-2): 123-132. Richter, R. P., J. Lai Kee Him, et al. (2005). "On the Kinetics of Adsorption and Two-Dimensional Self-Assembly of Annexin A5 on Supported Lipid Bilayers." Biophys. J. 89(5): 3372-3385. Robert, M. and C. Gagnon (1999). "Semenogelin I: a coagulum forming, multifunctional seminal vesicle protein." Cell Mol Life Sci 55(6-7): 944-60. Rochefort, C., G. Gheusi, et al. (2002). "Enriched Odor Exposure Increases the Number of Newborn Neurons in the Adult Olfactory Bulb and Improves Odor Memory." J. Neurosci. 22(7): 2679-2689. Roesch Ely, M., M. Nees, et al. (2005). "Transcript and proteome analysis reveals reduced expression of calgranulins in head and neck squamous cell carcinoma." European Journal of Cell Biology 84(2-3): 431-444. Rosario Donato (2003). "Intracellular and extracellular roles of S100 proteins." Microscopy Research and Technique 60(6): 540-551. Rosselli-Austin, L. and J. Williams (1990). "Enriched neonatal odor exposure leads to increased numbers of olfactory bulb mitral and granule cells." Brain Res Dev Brain Res 51(1): 135-7. Sallaz, M. and F. Jourdan (1993). "C-fos expression and 2-deoxyglucose uptake in the olfactory bulb of odour- stimulated awake rats." Neuroreport 4(1): 55-8. Santoni, V., M. Molloy, et al. (2000). "Membrane proteins and proteomics: un amour impossible?" Electrophoresis 21(6): 1054-70. Sarria, A. J., S. R. Panini, et al. (1992). "A functional role for vimentin intermediate filaments in the metabolism of lipoprotein-derived cholesterol in human SW-13 cells." J Biol Chem 267(27): 19455-63. Satoh, M. and M. Takeuchi (1995). "Induction of NCAM expression in mouse olfactory keratin-positive basal cells in vitro." Developmental Brain Research 87(2): 111-119. Schafer, B. W., J. M. Fritschy, et al. (2000). "Brain S100A5 is a novel calcium-, zinc-, and copper ion-binding protein of the EF-hand superfamily." J Biol Chem 275(39): 30623-30. Schagger, H. and G. von Jagow (1991). "Blue native electrophoresis for isolation of membrane protein complexes in enzymatically active form." Anal Biochem 199(2): 223-31. Schlehuber, S. and A. Skerra (2005). "Lipocalins in drug discovery: from natural ligand-binding proteins to "anticalins"." Drug Discov Today 10(1): 23-33. Schleicher, S., I. Boekhoff, et al. (1993). "A {beta}-Adrenergic Receptor Kinase-Like Enzyme is Involved in Olfactory Signal Termination." PNAS 90(4): 1420-1424. Schmachtenberg, O., J. Diaz, et al. (2003). "NO activates the olfactory cyclic nucleotide-gated conductance independent from cGMP in isolated rat olfactory receptor neurons." Brain Res 980(1): 146-50. Schmidt, U. and M. Eckert (1988). "The influence of early odour experience on the neural response of the olfactory bulb in laboratory mice." J Comp Physiol [A] 163(6): 771-6. Schuel, H. and L. J. Burkman (2005). "A Tale of Two Cells: Endocannabinoid-Signaling Regulates Functions of Neurons and Sperm." Biol Reprod 73(6): 1078-1086. Schwaller, B., M. Meyer, et al. (2002). "'New' functions for 'old' proteins: the role of the calcium-binding proteins calbindin D-28k, calretinin and parvalbumin, in cerebellar physiology. Studies with knockout mice." Cerebellum 1(4): 241-58. Schwob, J., N. Farber, et al. (1986). "Neurons of the olfactory epithelium in adult rats contain vimentin." J. Neurosci. 6(1): 208-217. Scott, K. (2004). "The sweet and the bitter of mammalian taste." Current Opinion in Neurobiology 14(4): 423-427.

Shariatmadari, R., P. Sipila, et al. (2003). "Adenosine triphosphate induces Ca2+ signal in epithelial cells of the mouse caput epididymis through activation of P2X and P2Y purinergic receptors." Biol Reprod 68(4): 1185-92. Sharon, N. and H. Lis (2004). "History of lectins: from hemagglutinins to biological recognition molecules." Glycobiology 14(11): 53R-62. Shepherd, G. M. (2005). "Perception without a thalamus how does olfaction do it?" Neuron 46(2): 166-8. Shetty, J., S. Naaby-Hansen, et al. (1999). "Human sperm proteome: immunodominant sperm surface antigens identified with sera from infertile men and women." Biol Reprod 61(1): 61-9. Shibahara, H., I. Sato, et al. (2002). "Two-dimensional electrophoretic analysis of sperm antigens recognized by sperm immobilizing antibodies detected in infertile women." Journal of Reproductive Immunology 53(1-2): 1-12. Shimshek, D. R., T. Bus, et al. (2005). "Impaired Reproductive Behavior by Lack of GluR-B Containing AMPA Receptors But Not of NMDA Receptors in Hypothalamic and Septal Neurons." Mol Endocrinol: me.2005- 0262. Silacci, P., L. Mazzolai, et al. (2004). "Gelsolin superfamily proteins: key regulators of cellular functions." Cell Mol Life Sci 61(19-20): 2614-23. Singer, M. S. (2000). "Analysis of the Molecular Basis for Octanal Interactions in the Expressed Rat I7 Olfactory Receptor." Chem. Senses 25(2): 155-165. Sitek, B., O. Apostolov, et al. (2005). "Identification of dynamic proteome changes upon ligand activation of Trk- receptors using two-dimensional fluorescence difference gel electrophoresis and mass spectrometry." Mol Cell Proteomics 4(3): 291-9. Skarda, C. A. and W. J. Freeman (1987). "How make chaos in order to make sense of the world." Behavioral & Brain Sciences 10(2): 161-195. Spehr, M., G. Gisselmann, et al. (2003). "Identification of a testicular odorant receptor mediating human sperm chemotaxis." Science 299(5615): 2054-8. Spehr, M., K. Schwane, et al. (2004). "Particulate adenylate cyclase plays a key role in human sperm olfactory receptor- mediated chemotaxis." J Biol Chem 279(38): 40194-203. Spehr, M., C. H. Wetzel, et al. (2002). "3-Phosphoinositides Modulate Cyclic Nucleotide Signaling in Olfactory Receptor Neurons." Neuron 33(5): 731-739. Stuart K. Archer, C. C., Hugh D. Campbell, (2005). "Evolution of the gelsolin family of actin-binding proteins as novel transcriptional coactivators." BioEssays 27(4): 388-396. Suzuki, K., J.-J. Lareyre, et al. (2004). "Molecular evolution of epididymal lipocalin genes localized on mouse chromosome 2." Gene 339: 49-59. Suzuki, Y. and M. Takeda (1991). "Keratins in the developing olfactory epithelia." Brain Res Dev Brain Res 59(2): 171- 8. Swain, M. and N. W. Ross (1995). "A silver stain protocol for proteins yielding high resolution and transparent background in sodium dodecyl sulfate-polyacrylamide gels." Electrophoresis 16(6): 948-51. Swanson, S. K. and M. P. Washburn (2005). "The continuing evolution of shotgun proteomics." Drug Discovery Today 10(10): 719-725. Takei, N., K. Ohsawa, et al. (1994). "Neurotrophic effects of annexin V on cultured neurons from embryonic rat brain." Neurosci Lett 171(1-2): 59-62. Taoka, M., A. Wakamiya, et al. (2000). "Protein profiling of rat cerebella during development." Electrophoresis 21(9): 1872-9. Tatsura, H., H. Nagao, et al. (2001). "Developing germ cells in mouse testis express pheromone receptors." FEBS Letters 488(3): 139-144. Tegoni, M., P. Pelosi, et al. (2000). "Mammalian odorant binding proteins." Biochimica et Biophysica Acta (BBA) - Protein Structure and Molecular Enzymology 1482(1-2): 229-240. Teratani, T., T. Watanabe, et al. (2002). "Restricted expression of calcium-binding protein S100A5 in human kidney." Biochem Biophys Res Commun 291(3): 623-7. Thompson, J., T. Gibson, et al. (1997). "The CLUSTAL_X windows interface: flexible strategies for multiple sequence alignment aided by quality analysis tools." Nucl. Acids Res. 25(24): 4876-4882. Timm, D. E., L. J. Baker, et al. (2001). "Structural basis of pheromone binding to mouse major urinary protein (MUP- I)." Protein Sci 10(5): 997-1004. Treloar, H. B., P. Feinstein, et al. (2002). "Specificity of Glomerular Targeting by Olfactory Sensory Axons." J. Neurosci. 22(7): 2469-2477. Tremblay, R., B. Chakravarthy, et al. (2000). "Transient NMDA Receptor Inactivation Provides Long-Term Protection to Cultured Cortical Neurons from a Variety of Death Signals." J. Neurosci. 20(19): 7183-7192.

Tun-Tzu Yu, J. C. M., Soma C. Bose, Debra Hardin, Michael C. Owen, Timothy S. McClintock, (2005). "Differentially expressed transcripts from phenotypically identified olfactory sensory neurons." The Journal of Comparative Neurology 483(3): 251-262. Vanderhaeghen, P., S. Schurmans, et al. (1993). "Olfactory receptors are displayed on dog mature sperm cells." J. Cell Biol. 123(6): 1441-1452. Vanderwolf, C. H. and E. M. Zibrowski (2001). "Pyriform cortex [beta]-waves: odor-specific sensitization following repeated olfactory stimulation." Brain Research 892(2): 301-308. Vener, A. V. and P. Stralfors (2005). "Vectorial proteomics." IUBMB Life 57(6): 433-40. Wain, H. M., E. A. Bruford, et al. (2002). "Guidelines for Human Gene Nomenclature." Genomics 79(4): 464-470. Walensky, L. D., A. J. Roskams, et al. (1995). "Odorant receptors and desensitization proteins colocalize in mammalian sperm." Mol Med 1(2): 130-41. Walter, H., Brooks, D. E. and Fisher, D. (1985). Partitioning in Aqueous Two-Phase Systems: Theory, Methods, Uses and Applications to Biotechnology, Academic Press, London. Walton, M., E. Sirimanne, et al. (1997). "Annexin V labels apoptotic neurons following hypoxia-ischemia." Neuroreport 8(18): 3871-5. Wang, H. and D. R. Storm (2003). "Calmodulin-regulated adenylyl cyclases: cross-talk and plasticity in the central nervous system." Mol Pharmacol 63(3): 463-8. Wang, H. W., C. J. Wysocki, et al. (1993). "Induction of olfactory receptor sensitivity in mice." Science 260(5110): 998- 1000. Wang, J., Z. A. Luthey-Schulten, et al. (2003). "Is the olfactory receptor a metalloprotein?" PNAS 100(6): 3035-3039. Washburn, M. P., D. Wolters, et al. (2001). "Large-scale analysis of the yeast proteome by multidimensional protein identification technology." Nat Biotechnol 19(3): 242-7. Watt, W. C. and D. R. Storm (2001). "Odorants Stimulate the ERK/Mitogen-activated Protein Kinase Pathway and Activate cAMP-response Element-mediated Transcription in Olfactory Sensory Neurons." J. Biol. Chem. 276(3): 2047-2052. Weinhold, B., G. Schratt, et al. (2000). "Srf-/- ES cells display non-cell-autonomous impairment in mesodermal differentiation." EMBO J. 19(21): 5835-5844. Wetzel, C. H., D. Brunert, et al. (2005). "Cellular mechanisms of olfactory signal transduction." Chem Senses 30 Suppl 1: i321-i322. Weyand, I., M. Godde, et al. (1994). "Cloning and functional expression of a cyclic-nucleotide-gated channel from mammalian sperm." 368(6474): 859-863. Whitehouse, C. M., R. N. Dreyer, et al. (1985). "Electrospray interface for liquid chromatographs and mass spectrometers." Anal Chem 57(3): 675-9. Wilkins, M. R., J. C. Sanchez, et al. (1996). "Progress with proteome projects: why all proteins expressed by a genome should be identified and how to do it." Biotechnol Genet Eng Rev 13: 19-50. Willis, D., K. W. Li, et al. (2005). "Differential Transport and Local Translation of Cytoskeletal, Injury-Response, and Neurodegeneration Protein mRNAs in Axons." J. Neurosci. 25(4): 778-791. Wilson, D. A. (2000). "Comparison of Odor Receptive Field Plasticity in the Rat Olfactory Bulb and Anterior Piriform Cortex." J Neurophysiol 84(6): 3036-3042. Wilson, D. A., A. R. Best, et al. (2004). "Plasticity in the olfactory system: lessons for the neurobiology of memory." Neuroscientist 10(6): 513-24. Wilson, D. A. and M. Leon (1988). "Spatial patterns of olfactory bulb single-unit responses to learned olfactory cues in young rats." J Neurophysiol 59(6): 1770-1782. Wolters, D. A., M. P. Washburn, et al. (2001). "An automated multidimensional protein identification technology for shotgun proteomics." Anal Chem 73(23): 5683-90. Wu, C. C., M. J. MacCoss, et al. (2003). "A method for the comprehensive proteomic analysis of membrane proteins." Nat Biotechnol 21(5): 532-8. Wu, C. C. and J. R. Yates, 3rd (2003). "The application of mass spectrometry to membrane proteomics." Nat Biotechnol 21(3): 262-7. Xiang, R., Y. Shi, et al. (2004). "2D LC/MS Analysis of Membrane Proteins from Breast Cancer Cell Lines MCF7 and BT474." J. Proteome Res. 3(6): 1278-1283. Yamada, K., E. Wada, et al. (2001). "Female gastrin-releasing peptide receptor (GRP-R)-deficient mice exhibit altered social preference for male conspecifics: implications for GRP/GRP-R modulation of GABAergic function." Brain Research 894(2): 281-287.

Yamada, S., D. Wirtz, et al. (2003). "The mechanical properties of simple epithelial keratins 8 and 18: discriminating between interfacial and bulk elasticities." Journal of Structural Biology 143(1): 45-55. Yang, Y. and Z. Li (2005). "Roles of heat shock protein gp96 in the ER quality control: redundant or unique function?" Mol Cells 20(2): 173-82. Yates III, J. R., A. Gilchrist, et al. (2005). "Proteomics of organelles and large cellular structures." Nature Reviews Molecular Cell Biology 6(9): 702-714. Yin, H. L. and T. P. Stossel (1979). "Control of cytoplasmic actin gel-sol transformation by gelsolin, a calcium- dependent regulatory protein." 281(5732): 583-586. Yoshihara, Y., K. Katoh, et al. (1993). "Odor stimulation causes disappearance of R4B12 epitope on axonal surface molecule of olfactory sensory neurons." Neuroscience 53(1): 101-110. Yue Ge, M. P. M., Jeffrey S. Chamberlain, Philip C. Andrews, (2003). "Proteomic analysis of mdx skeletal muscle: Great reduction of adenylate kinase 1 expression and enzymatic activity." Proteomics 3(10): 1895-1903. Zhang, J.-J., F. Okutani, et al. (2003). "Activation of the mitogen-activated protein kinase/extracellular signal-regulated kinase signaling pathway leading to cyclic AMP response element-binding protein phosphorylation is required for the long-term facilitation process of aversive olfactory learning in young rats." Neuroscience 121(1): 9-16. Zhang, W., G. Zhou, et al. (2003). "Affinity enrichment of plasma membrane for proteomics analysis." Electrophoresis 24(16): 2855-63. Zhang, X., I. Rodriguez, et al. (2004). "Odorant and vomeronasal receptor genes in two mouse genome assemblies." Genomics 83(5): 802-811. Zhang, Y., M. A. Hoon, et al. (2003). "Coding of Sweet, Bitter, and Umami Tastes: Different Receptor Cells Sharing Similar Signaling Pathways." Cell 112(3): 293-301. Zhao, Y., W. Zhang, et al. (2004). "Proteomic analysis of integral plasma membrane proteins." Anal Chem 76(7): 1817- 23. Zhong, B. and M. B. Omary (2004). "Actin overexpression parallels severity of pancreatic injury." Experimental Cell Research 299(2): 404-414. Zhong, H., S. M. Wade, et al. (2003). "A Spatial Focusing Model for G Protein Signals. Regulator of G protein signaling(RGS) protein-mediated kinetic scaffolding." J. Biol. Chem. 278(9): 7278-7284. Zozulya, S., F. Echeverri, et al. (2001). "The human olfactory receptor repertoire." Genome Biol 2(6): 1-12. Zufall, F. and H. Hatt (1991). "Dual Activation of a Sex Pheromone-Dependent Ion Channel from Insect Olfactory Dendrites by Protein Kinase C Activators and Cyclic GMP." PNAS 88(19): 8520-8524. Zufall, F. and T. Leinders-Zufall (1997). "Identification of a Long-Lasting Form of Odor Adaptation that Depends on the Carbon Monoxide/cGMP SecondMessenger System." J. Neurosci. 17(8): 2703-2712.

Jon Barbour

CONTACT DETAILS Dept. Cell Physiology, Ruhr University Bochum, Universitätsstrasse 150, Bochum, GermanyD-44780 Bochum

EDUCATION 2002–present Doctorate student Ruhr-Universität Bochum,Germany

„ Reading for PhD in Neuroscience with the International Graduate School of Neuroscience (IGSN).

„ Full collegiate member of the Graduiertenkolleg for “Development and Plasticity of the Nervous System: Molecular, synaptic and cellular mechanisms”.

„ School member of the International Max Plank Research School in Chemical Biology, Dortmund, Germany.

1997–1998 Masters of Research (MRes) University College London „ MRes in Advanced Instrumentation Systems, Faculty of Physics and Astronomy, UCL.

1994–1997 BSc (Hons) Biomedical Sc. Univeristy of Ulster (UU) „ Awarded First Class degree with commendation in diploma in industrial studies.

„ Received presidents prize from Inst of Biomedical Sciences for outstanding achievement.

1992–1994 Higher National Diploma in Applied Biology. UU „ Pass with distinction.

PUBLICATIONS/CONFERENCES „ Spehr M, Schwane K, Riffell JA, Barbour J, Zimmer RK, Neuhaus EM, Hatt H. Particulate adenylate cyclase plays a key role in human sperm olfactory receptor- mediated chemotaxis. J Biol Chem. 2004. 279(38):40194-203

„ Neuhaus EM, Mashukova A , Barbour DJ, Walters D and Hatt H. Novel function ß-arrestin2 in human sperm. Submitted

„ Neuhaus EM, Mashukova A, Zhang W, Barbour DJ and Hatt H. A specific heat shock protein enhances expression of odorant receptor proteins. Submitted

„ Barbour J, Warschield B, Stoepel N, Hatt H, Meyer H, Neuhaus E. Unravelling olfactory stimulus induced plasticity: Differential Analysis of Murine Olfactory Epithelium. In preparation

„ Proceedings of the European Life Sciences Organisation (ELSO), 2003, Dresden, Germany.

„ Barbour J, Warscheid B, Meyer H, Wolters D, Hatt H and Neuhaus E. Differential Proteomic Analysis of Murine Olfactory Epithelium. Proceedings of the International Symposium for Neuroscience Dynamics and Plasticity of Sensory Systems Function, 2004, Ruhr-University Bochum, Germany.

„ Barbour J, Warscheid B, Meyer H, Hatt H and Neuhaus E. Analysis of the mouse olfactory receptor associated micro-proteome. Proceedings of the 6th Meeting of the German Neuroscience Society / 30th Göttingen Neurobiology Conference, 2005, Göttingen, Germany.

„ Barbour J, Warscheid B, Kai Stühler, Meyer H, Hatt H and Neuhaus E. Stimulus induced plasticity of the olfactory receptor neuron. Proceedings of the 1st Hellenic Proteomics Society, in Cancer Genomics & Proteomics – Oral Presentation.

„ Barbour J, Wolters D, Hatt H, and Neuhaus M. Identification of Novel Neuronal Receptors in Human Sperm by Proteomic Analysis. Proceedings of the 4th Human Proteome Organisation, Munich 2005.

REFERENCES Prof. Dr. Dr. Dr. Hanns Hatt (department head) Dept. Cell Physiology, Ruhr University Bochum, Universitätsstrasse 150, Bochum, Germany D-44780 Bochum

Dr Eva Neuhaus (post-doctorate researcher) Dept. Cell Physiology, Ruhr University Bochum, Universitätsstrasse 150, Bochum, Germany D-44780 Bochum

Dr Dirk Wolters Dept. Analytical Chemistry, Ruhr-Universität Bochum, Universitätsstrasse 150, Bochum, Germany D-44780 Bochum

Junior Professor Bettina Warscheid Medical Proteome Centre, Ruhr-Universität Bochum, Universitätsstrasse 150, Bochum, Germany D-44780 Bochum