610

Mechanosensitive Ion Channels, Part B, Volume 59 (Current Topics in Membranes)

Embed Size (px)

Citation preview

  • Current Topics in Membranes, Volume 59

    Mechanosensitive Ion Channels, Part B

  • Current Topics in Membranes, Volume 59

    Series Editors

    Dale J. BenosDepartment of Physiology and Biophysics

    University of Alabama

    Birmingham, Alabama

    Sidney A. SimonDepartment of Neurobiology

    Duke University Medical Centre

    Durham, North Carolina

  • Current Topics in Membranes, Volume 59

    Mechanosensitive IonChannels, Part B

    Edited by

    Owen P. HamillDepartment of Neuroscience and Cell BiologyUniversity of Texas Medical BranchGalveston, Texas

    AMSTERDAM BOSTON HEIDELBERG LONDON

    NEW YORK OXFORD PARIS SAN DIEGO

    SAN FRANCISCO SINGAPORE SYDNEY TOKYO

    Academic Press is an imprint of Elsevier

  • Academic Press is an imprint of Elsevier525 B Street, Suite 1900, San Diego, California 92101-4495, USA84 Theobalds Road, London WC1X 8RR, UK

    This book is printed on acid-free paper.

    Copyright # 2007, Elsevier Inc. All Rights Reserved.

    No part of this publication may be reproduced or transmitted in any form or by anymeans, electronic or mechanical, including photocopy, recording, or any informationstorage and retrieval system, without permission in writing from the Publisher.

    The appearance of the code at the bottom of the first page of a chapter in this bookindicates the Publishers consent that copies of the chapter may be made forpersonal or internal use of specific clients. This consent is given on the condition,however, that the copier pay the stated per copy fee through the Copyright ClearanceCenter, Inc. (www.copyright.com), for copying beyond that permitted bySections 107 or 108 of the U.S. Copyright Law. This consent does not extend toother kinds of copying, such as copying for general distribution, for advertisingor promotional purposes, for creating new collective works, or for resale.Copy fees for pre-2007 chapters are as shown on the title pages. If no fee codeappears on the title page, the copy fee is the same as for current chapters.1063-5823/2007 $35.00

    Permissions may be sought directly from Elseviers Science & Technology RightsDepartment in Oxford, UK: phone: (+44) 1865 843830, fax: (+44) 1865 853333,E-mail: [email protected]. You may also complete your request on-linevia the Elsevier homepage (http://elsevier.com), by selecting Support & Contactthen Copyright and Permission and then Obtaining Permissions.

    For information on all Elsevier Academic Press publicationsvisit our Web site at www.books.elsevier.com

    ISBN-13: 978-0-12-153359-5ISBN-10: 0-12-153359-X

    PRINTED IN THE UNITED STATES OF AMERICA07 08 0 9 10 9 8 7 6 5 4 3 2 1

  • Contents

    Contributors xiii

    Foreword xvii

    Previous Volumes in Series xix

    CHAPTER 1 Mechanosensitive Ion Channels of Spiders:

    Mechanical Coupling, Electrophysiology, and

    Synaptic ModulationAndrew S. French and Paivi H. Torkkeli

    I. Overview 1II. Introduction 2III. Types of Spider Mechanoreceptors 3IV. Mechanical Coupling 3V. Mechanotransduction in Slit Sensilla 6VI. Dynamic Properties of Mechanotransduction

    and Action Potential Encoding 13VII. Calcium Signaling During Transduction by

    Spider Mechanoreceptors 14VIII. Synaptic Modulation of

    Spider Mechanoreceptors 15IX. Conclusions 17

    References 17

    CHAPTER 2 Ion Channels for Mechanotransduction in the

    Crayfish Stretch ReceptorBo Rydqvist

    I. Overview 21II. Introduction 22III. Morphology of the SRO 23IV. Functional Properties 24V. Summary and Discussion of Future

    Research Directions 43References 45

    v

  • CHAPTER 3 Mechanosensitive Ion Channels in

    Caenorhabditis elegansDafne Bazopoulou and Nektarios Tavernarakis

    I. Overview 49II. Introduction 50III. C. elegans Mechanosensitive Behaviors 51IV. C. elegans DEG/ENaCs 55V. C. elegans TRP Ion Channels 66VI. Concluding Remarks 72

    References 73

    CHAPTER 4 Properties and Mechanism of the Mechanosensitive

    Ion Channel Inhibitor GsMTx4, a Therapeutic

    Peptide Derived from Tarantula VenomPhilip A. Gottlieb, Thomas M. Suchyna, and Frederick Sachs

    I. Overview 81II. Introduction 82III. Properties and Specificity of GsMTx4 85IV. Cellular Sites for GsMTx4 95V. Potential Therapeutic Uses for GsMTx4 97VI. Conclusions 103

    References 103

    CHAPTER 5 Mechanosensitive Channels in Neurite OutgrowthMario Pellegrino and Monica Pellegrini

    I. Overview 111II. Introduction 112III. Encoding of Guidance Cues in

    Axon Pathfinding 112IV. Requirement of TRP Channels in

    Calcium-Dependent Axon Pathfinding 114V. Physical Guidance Cues and Role of

    Mechanosensitive Ion Channels 116VI. Ion Channels as Molecular Integrators 119VII. Concluding Remarks 120

    References 122

    CHAPTER 6 ENaC Proteins in Vascular Smooth

    Muscle MechanotransductionHeather A. Drummond

    I. Overview 127II. Introduction 128

    vi Contents

  • III. DEG/ENaC/ASIC Proteins are Members of aDiverse Protein Family Involved inMechanotransduction 129

    IV. Involvement of ENaC Proteins in VascularSmooth Muscle Mechanotransduction 137

    V. Summary and Future Directions 145References 145

    CHAPTER 7 Regulation of the Mechano-Gated K2P Channel

    TREK-1 by Membrane PhospholipidsJean Chemin, Amanda Jane Patel, Patrick Delmas, Fred Sachs, Michel

    Lazdunski, and Eric Honore

    I. Overview 155II. Introduction 156III. TREK-1 Stimulation by Membrane

    Phospholipids 158IV. TREK-1 Inhibition by Membrane

    Phospholipids 161References 168

    CHAPTER 8 MechanoTRPs and TRPA1Andrew J. Castiglioni and Jaime Garca-Anoveros

    I. Overview 171II. MechanoTRP Channels 174III. Characteristics of TRPA1 Gene

    and Protein 175IV. TRPA1 Expression in

    Mechanosensory Organs 176V. Function of TRPA1 177VI. Proposed Biological Roles for TRPA1 185

    References 186

    CHAPTER 9 TRPCs as MS ChannelsOwen P. Hamill and Rosario Maroto

    I. Overview 191II. Introduction 192III. Practical Aspects of Recording

    MS Channels 193

    Contents vii

  • IV. Distinguishing Direct vs IndirectMS Channels 195

    V. Extrinsic Regulation of Stretch Sensitivity 197VI. Strategies to Identify MS Channel Proteins 197VII. General Properties of TRPCs 198VIII. Evidence for TRPC Mechanosensitivity 203IX. Conclusions 215

    References 218

    CHAPTER 10 The Cytoskeletal Connection to Ion Channels

    as a Potential Mechanosensory Mechanism:

    Lessons from Polycystin-2 (TRPP2)Horacio F. Cantiello, Nicolas Montalbetti, Qiang Li,

    and Xing-Zhen Chen

    I. Overview 234II. Introduction 235III. Role of Actin Cytoskeletal Dynamics in

    PC2-Mediated Channel Function 253IV. Identification of Actin-Binding Protein

    Interactions with Polycystin-2 261V. EVect of Hydroosmotic Pressure on PC2

    Channel Function: Role of the Cytoskeletonin Osmosensory Function 265

    VI. The ChannelCytoskeleton Interface:StructuralFunctional Correlates 272

    VII. Perspective and Future Directions 281References 282

    CHAPTER 11 Lipid Stress at Play: Mechanosensitivity of

    Voltage-Gated ChannelsCatherine E. Morris and Peter F. Juranka

    I. Overview 298II. The System Components 298III. Big Picture Issues 301IV. Reversible Stretch-Induced Changes in

    Particular VGCs 319V. Irreversible Stretch-Induced Gating

    Changes in VGCs 325VI. Technical Issues 327VII. Summary Comments 330

    References 330

    viii Contents

  • CHAPTER 12 Hair Cell Mechanotransduction: The Dynamic

    Interplay Between Structure and FunctionAnthony J. Ricci and Bechara Kachar

    I. Overview 339II. Auditory System 340III. Hair Bundle Structure 341IV. MET Involves Mechanically Gated Channels 341V. Where are These Channels? 343VI. The Gating Spring Theory 344VII. How are the Channels Activated? 347VIII. To Be or Not to Be Tethered 349IX. Characterizing Channel Properties? 351X. MET Channel Pore 352XI. Adaptation 354XII. The Dynamic Hair Bundle 361XIII. Summary and Future Directions 365

    References 366

    CHAPTER 13 Insights into the Pore of the Hair Cell Transducer

    Channel from Experiments with Permeant BlockersSietse M. van Netten and Corne J. Kros

    I. Overview 376II. Introduction 376III. Ionic Selectivity of the Transducer Channel 377IV. Permeation and Block of Mechanoreceptor

    Channels by FM1-43 378V. Permeation and Block of the Hair Cell Transducer

    Channel by Aminoglycoside Antibiotics 382VI. Transducer Channel Block by Amiloride

    and Its Derivatives 391VII. Conclusions 394

    References 396

    CHAPTER 14 Models of Hair Cell MechanotransductionSusanne Bechstedt and Jonathon Howard

    I. Overview 399II. Introduction 400III. Transduction Channel Properties 401IV. Gating 408V. Active Hair Bundle Motility 415VI. Conclusions 418

    References 418

    Contents ix

  • CHAPTER 15 TouchLiam J. Drew, Francois Rugiero, and John N. Wood

    I. Overview 426II. Introduction 426III. Structure of Skin and Touch Receptors 427IV. Physiology of Mechanoreceptive

    Nerve Fibers 432V. Quantitating Mechanical Responses in

    Animal Models 435VI. Electrophysiological Approaches to

    Mechanosensation in Rodents 436VII. Mechanosensitive Ion Channels in Cultured

    Sensory Neurons 437VIII. Gating MS Ion Channels in DRG Neurons 446IX. Candidate Ion Channels 447X. Voltage-Gated Channels and

    Mechanosensation 454XI. Indirect Signaling Between Sensory Neurons and

    Nonneuronal Cells 456XII. Conclusions 457

    References 457

    CHAPTER 16 Mechanosensitive Ion Channels in

    Dystrophic MuscleJeffry B. Lansman

    I. Overview 467II. Introduction 468III. MS Channel Expression During Myogenesis 469IV. Permeabilty Properties of MS Channels in

    Skeletal Muscle 470V. Gating 471VI. Pharmacology 478VII. Conclusions 481

    References 482

    CHAPTER 17 MscCa Regulation of Tumor Cell Migration

    and MetastasisRosario Maroto and Owen P. Hamill

    I. Overview 485II. Introduction 486

    x Contents

  • III. DiVerent Modes of Migration 487IV. Ca2 Dependence of Cell Migration 490V. The Role of MscCa in Cell Migration 499VI. Can Extrinsic Mechanical Forces Acting on MscCa

    Switch on Cell Migration? 501References 502

    CHAPTER 18 Stretch-Activated Conductances in Smooth MusclesKenton M. Sanders and Sang Don Koh

    I. Overview 511II. Introduction 512III. Mechanosensitive Conductances that Generate

    Inward Currents 514IV. Mechanosensitive Conductances that Generate

    Outward Currents 527References 535

    CHAPTER 19 Mechanosensitive Ion Channels in Blood

    Pressure-Sensing Baroreceptor NeuronsMark W. Chapleau, Yongjun Lu, and Francois M. Abboud

    I. Overview 541II. Introduction 542III. BR Sensory Transduction 544IV. Mechanosensitive Channels in BR Neurons 548V. Methodological Limitations and Challenges 558VI. Summary and Future Directions 560

    References 561

    Index 569

    Contents xi

  • This page intentionally left blank

  • Contributors

    Numbers in parentheses indicate the pages on which the authors contributions begin.

    Francois M. Abboud (541), The Cardiovascular Center,Department of Internal Medicine, and Department of MolecularPhysiology & Biophysics, The University of Iowa Carver Collegeof Medicine, Iowa City, Iowa 52242

    Dafne Bazopoulou (49), Institute of Molecular Biology andBiotechnology, Foundation for Research and Technology,Heraklion 71110, Crete, Greece

    Susanne Bechstedt (399), MaxPlanckInstitute of Molecular CellBiology and Genetics (MPICBG), 01307 Dresden, Germany

    Horacio F. Cantiello (233), Renal Unit, Massachusetts GeneralHospital East, Charlestown, Massachusetts 02129; Department ofMedicine, Harvard Medical School, Boston, Massachusetts 02115;Laboratorio de Canales Ionicos, Departamento de Fisicoqumica yQumica Analtica, Facultad de Farmacia y Bioqumica, BuenosAires 1113, Argentina

    Andrew J. Castiglioni (171), Departments of Anesthesiology,Physiology, and Neurology, Northwestern University Institute forNeuroscience, Feinberg School of Medicine, NorthwesternUniversity, Chicago, Illinois 60611

    Mark W. Chapleau (541), The Cardiovascular Center, Departmentof Internal Medicine, and Department of Molecular Physiology &Biophysics, The University of Iowa Carver College ofMedicine, Iowa City, Iowa 52242; Veterans Affairs MedicalCenter, Iowa City, Iowa 52246

    Jean Chemin (155), Institut de Genomique Fonctionnelle, UPR2580 CNRS, F34094 Montpellier cedex 05, France

    XingZhen Chen (233), Department of Physiology, University ofAlberta, Edmonton T6G2H7, Canada

    xiii

  • Patrick Delmas (155), Laboratoire de NeurophysiologieCellulaire, Faculte de Medecine, UMR 6150 CNRS, 13916Marseille Cedex 20, France

    Liam J. Drew (425), Molecular Nociception Group, BiologyDepartment, University College London, London WC1E 6BT,United Kingdom

    Heather A. Drummond (127), Department of Physiology,The Center for Excellence in CardiovascularRenal Research,University of Mississippi Medical Center, Jackson,Mississippi 39216

    Andrew S. French (1), Department of Physiology and Biophysics,Dalhousie University, Halifax, Nova Scotia B3H 1X5, Canada

    Jaime GarcaAnoveros (171), Departments of Anesthesiology,Physiology, and Neurology, Northwestern University Institute forNeuroscience, Feinberg School of Medicine, NorthwesternUniversity, Chicago, Illinois 60611

    Philip A. Gottlieb (81), The Department of Physiology andBiophysics, Center for Single Molecule Biophysics,SUNY at Buffalo, Buffalo, New York 14214

    Owen P. Hamill (191, 485), Department of Neuroscience andCell Biology, University of Texas Medical Branch, Galveston,Texas 77555

    Eric Honore (155), Institut de Pharmacologie Moleculaire etcellulaire, UMR 6097 CNRS, 06560 Valbonne, France

    Jonathon Howard (399), MaxPlanckInstitute of Molecular CellBiology and Genetics (MPICBG), 01307 Dresden, Germany

    Peter F. Juranka (297), Neuroscience, Ottawa HealthResearch Institute, Ottawa Hospital, Ottawa, Ontario K1Y 4E9,Canada

    Bechara Kachar (339), Section of Structural Biology, NationalInstitutes of Deafness and Communicative Disorders, Bethesda,Maryland 20892

    Sang Don Koh (511), Department of Physiology and CellBiology, University of Nevada School of Medicine, Reno,Nevada 89557

    xiv Contributors

  • Corne J. Kros (375), School of Life Sciences, University of Sussex,Falmer, Brighton BN1 9QG, United Kingdom

    Jeffry B. Lansman (467), Department of Cellular and MolecularPharmacology, School of Medicine, University of California,San Francisco, California 94143

    Michel Lazdunski (155), Institut de Pharmacologie Moleculaire etcellulaire, UMR 6097 CNRS, 06560 Valbonne, France

    Qiang Li (233), Department of Physiology, University of Alberta,Edmonton T6G 2H7, Canada

    Yongjun Lu (541), The Cardiovascular Center and Department ofInternal Medicine, The University of Iowa Carver College ofMedicine, Iowa City, Iowa 52242

    Rosario Maroto (191, 485), Department of Neuroscience andCell Biology, University of Texas Medical Branch, Galveston,Texas 77555

    Nicolas Montalbetti (233), Laboratorio de Canales Ionicos,Departamento de Fisicoqumica y Qumica Analtica, Facultad deFarmacia y Bioqumica, Buenos Aires 1113, Argentina

    Catherine E. Morris (297), Neuroscience, Ottawa Health ResearchInstitute, Ottawa Hospital, Ottawa, Ontario K1Y 4E9, Canada

    Amanda Jane Patel (155), Institut de Pharmacologie Moleculaire etcellulaire, UMR 6097 CNRS, 06560 Valbonne, France

    Monica Pellegrini (111), Scuola Normale Superiore, Pisa, Italy

    Mario Pellegrino (111), Dipartimento di Fisiologia UmanaG. Moruzzi, Universita di Pisa, Pisa, Italy

    Anthony J. Ricci (339), Department of Otolaryngology, StanfordUniversity, Stanford, California 94305

    Francois Rugiero (425), Molecular Nociception Group, BiologyDepartment, University College London, London WC1E 6BT,United Kingdom

    Bo Rydqvist (21), Department of Physiology and Pharmacology,Karolinska Institutet, SE171 77 Stockholm, Sweden

    Frederick Sachs (81, 155), The Department of Physiology andBiophysics, Center for Single Molecule Biophysics,SUNY at Buffalo, Buffalo, New York 14214

    Contributors xv

  • Kenton M. Sanders (511), Department of Physiology andCell Biology, University of Nevada School of Medicine, Reno,Nevada 89557

    Thomas M. Suchyna (81), The Department of Physiology andBiophysics, Center for Single Molecule Biophysics,SUNY at Buffalo, Buffalo, New York 14214

    Nektarios Tavernarakis (49), Institute of Molecular Biology andBiotechnology, Foundation for Research and Technology,Heraklion 71110, Crete, Greece

    Paivi H. Torkkeli (1), Department of Physiology and Biophysics,Dalhousie University, Halifax, Nova Scotia B3H 1X5, Canada

    Sietse M. van Netten (375), Department of Neurobiophysics,University of Groningen, 9747AG, Groningen, The Netherlands

    John N. Wood (425), Molecular Nociception Group, BiologyDepartment, University College London, London WC1E 6BT,United Kingdom

    xvi Contributors

  • Foreword

    Mechanosensitive Ion Channels, Part B

    Owen P. Hamill

    Department of Neuroscience and Cell Biology,

    University of Texas Medical Branch,

    Galveston, Texas

    One of the great challenges in studying mechanotransduction (MT)

    has been to identify the mechanisms that underlie the exquisite sensitivity

    and highfrequency response of specific animal mechanotransducersa

    spider detects substrate vibrations within thermal noise limits, whereas a bat

    generates and detects ultrasounds of frequencies up to 100 kHz in echolocat-

    ing flying prey. Part B of this volume on mechanosensitive (MS) channels

    covers the diversity of MS channels and MT mechanisms evident in diVerent

    invertebrate and vertebrate mechanotransducers. The combined chapters

    highlight the integration ofMS channels into signaling complexes that interact

    with ancillary structures and other channels that are critical in shaping the

    specific inputoutput relations of mechanotransducers. The opening chapters

    describe MT in the slit sensilla in the spiders leg, the stretch receptor organ

    in crayfish muscle, and specific touch receptors in the nematode worm,

    Caenorhabditis elegans. The studies indicate at least twomajor channel families,

    the epithelial Na channel (ENaC) and the transient receptor potential (TRP)

    channels, are involved in MT in lower invertebrates. Subsequent chapters

    review the roles for ENaC and various TRP channels, and also the MS two

    poredomain K channels andMS voltagegated channels in mediatingMT in

    mammalian cells.Oneof themajor hurdles in studyingMThasbeen the absence

    of specific agents that selectively target MS channelsthe potential for the

    tarantula spider venom peptide GsMTx4 to serve this role is discussed in one

    chapter. Perhaps one of the interesting actions of GsMTx4 is that it strongly

    potentiates neurite outgrowth presumably via block of anMS channel that acts

    as a negative regulator of neurite outgrowth first demonstrated in the leech and

    reviewed in another chapter. Several chapters highlight diVerent aspects of the

    most intensely studied of all biological mechanotransducers, namely those

    mediatingvertebratehearing and touch.These two formsofMThavepresented

    xvii

  • the greatest challenge in identifying the membrane proteins forming MS

    channels, and each chapter provides new information and diVerent approaches

    that should help in completing this goal. The last part of the volume includes

    chapters that address the properties of MS channels in cell types where

    abnormalities inMTcontribute to significant humanpathologies, including the

    elevated stretchinduced Ca2 influx that contributes to muscle fiber degener-

    ation in muscular dystrophy, abnormalities in the regulation of smooth muscle

    tone, and baroreception that lead to hypertension, and the alterations in MS

    channel functional expression that may contribute to increased tumor cell

    motility and invasion during cancer progression.

    As indicated in Part A of this volume, I would like to thank Dale Benos for

    his original invitation to submit the proposal to Elsevier. I would also like to

    thank all those involved in the production of the volume and, in particular,

    Phil Carpenter for his continual and patient eVorts during the compilation

    phase. Finally, I would like to thank all the scientists for presenting their

    discoveries regarding MS channels.

    xviii Foreword

  • Previous Volumes in Series

    Current Topics in Membranes and Transport

    Volume 23 Genes and Membranes: Transport Proteins and Receptors*(1985)Edited by Edward A. Adelberg and Carolyn W. Slayman

    Volume 24 Membrane Protein Biosynthesis and Turnover (1985)Edited by Philip A. Knauf and John S. Cook

    Volume 25 Regulation of Calcium Transport across Muscle Membranes(1985)Edited by Adil E. Shamoo

    Volume 26 NaH Exchange, Intracellular pH, and Cell Function*(1986)Edited by Peter S. Aronson and Walter F. Boron

    Volume 27 The Role of Membranes in Cell Growth and Differentiation(1986)Edited by Lazaro J. Mandel and Dale J. Benos

    Volume 28 Potassium Transport: Physiology and Pathophysiology*(1987)Edited by Gerhard Giebisch

    Volume 29 Membrane Structure and Function (1987)Edited by Richard D. Klausner, Christoph Kempf, and Josvan Renswoude

    Volume 30 Cell Volume Control: Fundamental and ComparativeAspects in Animal Cells (1987)Edited by R. Gilles, Arnost Kleinzeller, and L. Bolis

    Volume 31 Molecular Neurobiology: Endocrine Approaches (1987)Edited by Jerome F. Strauss, III, and Donald W. Pfaff

    *Part of the series from the Yale Department of Cellular and Molecular Physiology.

    xix

  • Volume 32 Membrane Fusion in Fertilization, Cellular Transport, andViral Infection (1988)Edited by Nejat Duzgunes and Felix Bronner

    Volume 33 Molecular Biology of Ionic Channels* (1988)Edited by William S. Agnew, Toni Claudio, and Frederick J. Sigworth

    Volume 34 Cellular and Molecular Biology of Sodium Transport* (1989)Edited by Stanley G. Schultz

    Volume 35 Mechanisms of Leukocyte Activation (1990)Edited by Sergio Grinstein and Ori D. Rotstein

    Volume 36 ProteinMembrane Interactions* (1990)Edited by Toni Claudio

    Volume 37 Channels and Noise in Epithelial Tissues (1990)Edited by Sandy I. Helman and Willy Van Driessche

    Current Topics in Membranes

    Volume 38 Ordering the Membrane Cytoskeleton Trilayer* (1991)Edited by Mark S. Mooseker and Jon S. Morrow

    Volume 39 Developmental Biology of Membrane Transport Systems(1991)Edited by Dale J. Benos

    Volume 40 Cell Lipids (1994)Edited by Dick Hoekstra

    Volume 41 Cell Biology and Membrane Transport Processes* (1994)Edited by Michael Caplan

    Volume 42 Chloride Channels (1994)Edited by William B. Guggino

    Volume 43 Membrane ProteinCytoskeleton Interactions (1996)Edited by W. James Nelson

    Volume 44 Lipid Polymorphism and Membrane Properties (1997)Edited by Richard Epand

    Volume 45 The Eyes Aqueous Humor: From Secretion to Glaucoma(1998)Edited by Mortimer M. Civan

    xx Previous Volumes in Series

  • Volume 46 Potassium Ion Channels: Molecular Structure, Function, andDiseases (1999)Edited by Yoshihisa Kurachi, Lily Yeh Jan, and Michel Lazdunski

    Volume 47 AmilorideSensitive Sodium Channels: Physiology andFunctional Diversity (1999)Edited by Dale J. Benos

    Volume 48 Membrane Permeability: 100 Years since Ernest Overton(1999)Edited by David W. Deamer, Arnost Kleinzeller, andDouglas M. Fambrough

    Volume 49 Gap Junctions: Molecular Basis of Cell Communication inHealth and DiseaseEdited by Camillo Peracchia

    Volume 50 Gastrointestinal Transport: Molecular PhysiologyEdited by Kim E. Barrett and Mark Donowitz

    Volume 51 AquaporinsEdited by Stefan Hohmann, Sren Nielsen and Peter Agre

    Volume 52 PeptideLipid InteractionsEdited by Sidney A. Simon and Thomas J. McIntosh

    Volume 53 CalciumActivated Chloride ChannelsEdited by Catherine Mary Fuller

    Volume 54 Extracellular Nucleotides and Nucleosides: Release,Receptors, and Physiological and Pathophysiological EffectsEdited by Erik M. Schwiebert

    Volume 55 Chemokines, Chemokine Receptors, and DiseaseEdited by Lisa M. Schwiebert

    Volume 56 Basement Membrances: Cell and Molecular BiologyEdited by Nicholas A. Kefalides and Jacques P. Borel

    Volume 57 The Nociceptive MembraneEdited by Uhtaek Oh

    Volume 58 Mechanosensitive Ion Channels, Part AEdited by Owen P. Hamill

    Previous Volumes in Series xxi

  • CHAPTER 1

    Mechanosensitive Ion Channels ofSpiders: Mechanical Coupling,Electrophysiology, andSynaptic Modulation

    Andrew S. French and Paivi H. Torkkeli

    Department of Physiology and Biophysics, Dalhousie University, Halifax,

    Nova Scotia B3H 1X5, Canada

    I. Overview

    II. Introduction

    III. Types of Spider Mechanoreceptors

    IV. Mechanical Coupling

    V. Mechanotransduction in Slit Sensilla

    A. The Ionic Selectivity of Spider Mechanosensitive Channels

    B. The Location of VS3 Mechanosensitive Channels

    C. Mechanosensitive Channel Conductance, Density, and pH Sensitivity

    D. Temperature Sensitivity of Mechanosensitive Channels

    E. Molecular Characterization of Spider Mechanosensitive Channels

    VI. Dynamic Properties of Mechanotransduction and Action Potential Encoding

    VII. Calcium Signaling During Transduction by Spider Mechanoreceptors

    VIII. Synaptic Modulation of Spider Mechanoreceptors

    IX. Conclusions

    References

    I. OVERVIEW

    Arthropods have provided several important mechanoreceptor models

    because of the relatively large size and accessibility of their primary sensory

    neurons. Three types of spider receptors: tactile hairs, trichobothria, and slit

    sensilla have given important information about the coupling of external

    Current Topics in Membranes, Volume 59 1063-5823/07 $35.00Copyright 2007, Elsevier Inc. All right reserved. DOI: 10.1016/S1063-5823(06)59001-5

  • mechanical stimuli to the neuronal membrane, transduction of mechanical

    force into receptor current, encoding of aVerent action potentials, and eVerent

    modulation of peripheral sensory receptors. Slit sensilla, found only in spiders,

    have been particularly important because they allow intracellular recording from

    sensory neurons during mechanical stimulation. Experiments on slit sensilla

    have shown that their mechanosensitive ion channels are sodium selective,

    blocked by amiloride, and open more at low pH. This evidence suggests that

    the channels are members of the same molecular family as degenerins, acid

    sensitive ion channels, and epithelial sodium channels. Slit sensilla have also

    yielded evidence about the location, density, singlechannel conductance, and

    dynamic properties of the mechanosensitive channels. Spider mechanoreceptors

    are modulated in the periphery by eVerent neurons, and possibly by circulating

    chemicals. Mechanisms of modulation, intracellular signaling, and the role of

    intracellular calcium are areas of active investigation.

    II. INTRODUCTION

    Humans inhabit a sensory world dominated by vision, but we also use

    mechanotransduction to provide the senses of hearing, vestibular sensation,

    touch, and vibration, as well as chemotransduction for the senses of taste and

    smell. In contrast to our visual world, a spiders life is dominated by vibration

    and other mechanical inputs, even in those spider species that have relatively

    good vision.Waiting for prey to land on aweb, hunting along the ground or on a

    plant, and negotiating a vibratory mating ritualin all their daily activities the

    mechanical senses are vitally important. In addition, both humans and spiders

    detect a variety of internally generated mechanical signals from their musculo-

    skeletal systems and internal organs that allow feedback regulation ofmovement

    and many internal physiological processes.

    Although mechanotransduction is such an important sense for humans,

    spiders, and most other animals, its fundamental mechanisms have been

    diYcult to unravel, mainly due to the small size and complex morphology of

    most mechanoreceptor endings. Arthropods (insects, arachnids, and crus-

    taceans) not only possess large arrays of diVerent mechanoreceptors, but

    the relatively large sizes of some of their sensory neurons, and the close ass-

    ociation of many mechanosensory neurons to the external cuticle have pro-

    vided several model systems for investigating fundamental mechanisms of

    mechanotransduction.

    The most crucial step in mechanotransduction is a change in cell mem-

    brane potential, the receptor potential, produced by the application of a

    mechanical stimulus to the cell. To study this phenomenon ideally requires

    a preparation where the electrical event can be directly observed during

    2 French and Torkkeli

  • accurately controlled mechanical stimulation. This is possible in several

    spider preparations, and the information thus obtained will be the major

    subject here.

    III. TYPES OF SPIDER MECHANORECEPTORS

    The hairiness of spiders is well known, but what are the functions of the

    thousands of hairs covering a typical spider? Many provide nonsensory fun-

    ctions. These include adhesion to the substrate via surface tension, combing

    of silk threads from spinnerets, supporting the air bubbles of water spiders,

    providing attachment sites for spiderlings clinging to a female, and deterring

    predators by intense skin irritation (reviewed by Foelix, 1996). However,

    most of the surface hairs are sensory structures. Two major types of sensory

    hairs are the trichobothria, or filiform hairs, and the shorter tactile hairs

    (Fig. 1). Each of these hair structures is innervated by multiple neurons, typi-

    cally four inCupiennius salei, although it is not clear that all these neurons are

    mechanically sensitive. This situation contrasts somewhat with insects, which

    typically have only one sensory neuron per hair, but the general structures are

    otherwise similar.

    In addition to hairs that extend beyond the cuticle, embedded in spider

    cuticle are numerous mechanoreceptors of a type that is not found in other

    arthropods, the slit sensilla (Figs. 1 and 2). These are widely distributed in

    the exoskeleton, including the legs, pedipalps, and body (Barth and Libera,

    1970; Barth, 1985, 2001; Patil et al., 2006). They detect mechanical events in

    the cuticle, primarily strains imposed by normal movements of the animal

    and vibrations due to predators, prey, and mates.

    Spiders also possess a range of mechanoreceptors deeper within the animal,

    particularly the joint receptors and muscle receptors, but spiders apparently

    lack the chordotonal structures that are widespread in insects and crusta-

    ceans, serving particularly as vibration and auditory receptors (Seyfarth,

    1985; Barth, 2001).

    IV. MECHANICAL COUPLING

    The first functional stage of any mechanoreceptor is mechanical coupl-

    ing from the initial stimulus to the mechanically sensitive membrane of

    the sensory neuron. A large contribution to overall function is suggested,

    although not yet proven, by the wide range of accessory structures found in

    mechanoreceptors of both vertebrates and invertebrates, which are assumed

    1. Mechanotransduction in Spiders 3

  • to serve a mechanical coupling role. Detailed quantitative understanding of

    this coupling function is limited by the relatively small sizes of most receptors

    and the unknown mechanical properties of the materials used to construct

    the structures surrounding the sensory endings. The dynamic properties of

    coupling structures are particularly diYcult to elucidate because it is hard to

    10 mm

    Cuticle

    Sensorydendrites

    Slit

    Lymph space

    Supportingcells

    To cellbodies

    50 mm

    FIGURE 1 Major types of spider cuticular mechanoreceptors. Top left: hair sensilla at the

    joint between the tibia (left) and the femur of a leg of Cupiennius salei. Longer, vertical hairs are

    trichobothria, typically about 1mm long, surrounded by numerous shorter tactile hairs. Top

    right: scanning electron micrograph of a lyriform organ consisting of approximately parallel slit

    sensilla from a leg of C. salei. Dark circles are the sockets of broken hair sensilla. Lower drawing

    shows the arrangement of sensory neurons and surrounding tissues at a typical slit sensillum.

    Pairs of sensory dendrites, up to 200m long terminate in a ciliary enlargement that leads to a

    tubular body surrounded by a dense dendritic sheath. Supporting cells produce a lymph space

    surrounding the terminal dendrites that has a diVerent ionic composition than the normal

    extracellular fluid. One of the two sensory dendrites proceeds further into the slit structure,

    but the functional reason for this diVerence is unknown. On the basis of data from Barth, 2001,

    2004; Widmer et al. (2005).

    4 French and Torkkeli

  • measure the individual movements of each component as the sensillum is

    mechanically stimulated.

    Barth (2001, 2004) has discussed in depth the available evidence about

    mechanical coupling of spider trichobothria, hair sensilla, and slit sensilla.

    This work also builds on a substantial base of comparable studies in insect

    cuticular sensilla. Tactile hairs, as the name implies, are thought to serve as

    touch detectors. They can bend, as well as rotate within their sockets, prov-

    iding a reduction of movement estimated to be about 1:750, so that relatively

    Stimulatorprobe

    100 mm

    Slits

    Patella cuticle

    1 mm

    500 pA

    200 ms

    FIGURE 2 Intracellular recording from VS3 neurons. The approximately tubular patella is

    split in two along its length and the muscle tissues removed to reveal the mechanosensory

    neurons lying in the hypodermal membrane. A glass microelectrode is used to penetrate the

    soma of a neuron while a mechanical probe is raised from below to indent the appropriate slit

    from the outside. Step indentations under voltage clamp produce inward receptor currents that

    saturate at a few micrometers. The receptor currents have an adapting component, but most of

    the current adapts relatively slowly and incompletely. On the basis of data from Hoger et al.

    (1997).

    1. Mechanotransduction in Spiders 5

  • large external movements can be detected without damaging the hair.

    The longer trichobothria are specialized to detect air movements, and their

    varying lengths appear to be tuned to the fluid dynamics of air flow over

    the spider surface, especially considering the boundary layer eVect. Estim-

    ates of their sensitivity indicate that they can detect movements carrying

    energy equivalent to a single photon of visible light and that they operate

    close to the level of baseline thermal noise. They seem designed optimally to

    detect turbulent air flow produced by rapidly moving prey, such as flying

    insects, and their varying lengths and diameters provide tuning to diVerent

    stimulation frequencies.

    Slit sensilla are distributed in a wide range of patterns over the spider

    body, from single, isolated slits to complex arrangements of multiple slits,

    forming lyriform structures (Fig. 1). It is clear that slit sensilla respond to

    strain in the exoskeleton, produced by the animals movements or by vibra-

    tions conducted through the substrate. Measurements in models of spider leg

    cuticle indicate that the slits are optimally positioned to detect strain at the

    locations where it is maximized by normal loading and that slit orientations

    are matched to the directions of maximum natural stress. Most compound

    lyriform organs occur near the leg joints, while individual slits are often

    found at points of muscle attachment to the cuticle (Barth, 2001). The fine

    structure of an individual slit allows cuticular stress to apply a levered com-

    pression to the tips of the sensory dendrites. This arrangement has some sim-

    ilarities to the campaniform sensilla of insects, which seem to serve a similar

    stressdetecting function but use singly innervated, circular structures.

    The varying lengths of the slits in a lyriform organ (typically 8 to 200 mlong by 1 to 2 m wide) immediately suggest tuning to diVerent temporalfrequencies, as in the eponymous lyre. There is some evidence that this

    occurs, but the varying lengths may also serve functions such as measuring

    the relative intensity of the strain by progressive recruitment of diVerent slits

    as strain increases (Barth, 2001).

    V. MECHANOTRANSDUCTION IN SLIT SENSILLA

    Spider slit sensilla have provided important experimental preparations

    for research into mechanotransduction because of the following advantages.

    (1) Their mechanical structures, while complex, are approximately two

    dimensional and relatively amenable to analysis and stimulation. (2) The

    exposed location of the sensory neurons inside the surface cuticle has

    allowed the development of preparations in which simultaneous mechani-

    cal stimulation and stable intracellular recording, including voltageclamp

    6 French and Torkkeli

  • recording can be conducted. (3) The sensory neurons are located within

    a hypodermal membrane that allows them to be removed from the animal

    intact. This has been particularly useful for studying their voltageactivated

    conductances. (4) A complex eVerent innervation of the peripheral parts

    of sensory neurons promises to shed new light into understanding how

    mechanosensation is modulated.

    The remainder of this chapter will focus on major findings about me-

    chanotransduction, sensory encoding, and eVerent modulation of these

    processes that have emerged from research on spider lyriform organs and

    trichobothria.

    A. The Ionic Selectivity of Spider Mechanosensitive Channels

    Intracellular recording during mechanical stimulation has been achieved

    in two spider leg lyriform organs, VS3 on the patella (Juusola et al., 1994)

    and HS10 on the metatarsus (Gingl et al., 2006). In each case, all neurons

    innervating the slits were found to be mechanosensitive. Voltageclamp

    recording from the neuron cell bodies of VS3 revealed an inward, depolar-

    izing receptor current with both adapting and longlasting components that

    saturated with slit indentations of about 3 m (Fig. 2). Note that the slitindentation used in these experiments does not represent a natural stimulus.

    Although the major functions of VS3 remain unclear, normal slit compres-

    sion is presumably produced by cuticle strains. However, more natural

    stimulation of HS10 was achieved by moving the tarsus and this gave very

    similar results to the VS3 slit indentation.

    The receptor current in VS3 neurons could not be reversed, even with

    strong depolarization, and was completely eliminated when external sodium

    was replaced by choline (Fig. 3). Further tests with the common monovalent

    and divalent cations showed that, other than sodium, only lithium ions had

    detectable, but much lower, permeation (Hoger et al., 1997). These experi-

    ments indicate that spider mechanosensitive channels are highly selective for

    sodium ions.

    Further support for this selectivity comes from measurements of the ionic

    composition of the solution in the lymph space that surrounds the dendrite

    tips (Fig. 1). Comparable insect mechanoreceptors have a high concentration

    of potassium ions in this region, as well as a potential that is positive

    compared to the normal extracellular space (Thurm and Kuppers, 1980;

    Grunert and Gnatzy, 1987), but in spiders this region not only lacks the high

    potassium and positive potential but also has a relatively high concentration

    of sodium ions (Rick et al., 1976).

    1. Mechanotransduction in Spiders 7

  • B. The Location of VS3 Mechanosensitive Channels

    The bipolar structure of arthropod cuticular mechanoreceptor neurons

    (Fig. 2) has led to a long history of attempts to find the location of the mechan-

    osensitive channels, as well as the location of the action potentialinitiating

    region. Although the obvious location for transduction would seem to be at

    the distal tips of the dendrites because of the close apposition to the initial

    mechanical stimulus and the specialized electrochemical gradient of the lymph

    space (Fig.1), there have also been theories that transduction occurs near the

    ciliarybasalbodyand thatactionpotentialsmightarise in theaxosomatic region

    (reviewed by French, 1988).

    A direct test of the location of mechanotransduction was performed by

    applying small punctate stimuli to diVerent locations along the dendrites of

    VS3 neurons (Hoger and Seyfarth, 2001). Only stimuli applied to the distal

    dendrites, close to the inner surface of the slits, produced electrical activity in

    the neurons, suggesting a distal location.

    The general direction of signal flow in a sensory receptor from distal to

    proximal implies that transduction should occur either at the site of action

    potential initiation or possibly distal to it. Gingl and French (2003) used

    several techniques to locate the site of action potential initiation in VS3

    neurons, including the voltage jumpmethod that measures collisions between

    50 ms

    50 pA

    Control

    Choline100 pA

    100 pA

    200 pA

    300 pA

    100 mV 100 mV

    FIGURE 3 Receptor current is carried by sodium ions in VS3 neurons. Graph shows

    typical peak receptor currents produced by step slit indentations of 3 m while the neuronalmembrane was held at diVerent potentials. Note the failure to reverse, even at strong positive

    potentials. Replacement of the sodium ions in spider saline with the large choline cation

    completely eliminated the receptor current, but it returned when the normal saline solution

    was restored (control). On the basis of data from Hoger et al. (1997).

    8 French and Torkkeli

  • voltage waves started by the receptor potential and an artificially created

    potential step at the soma. These measurements all indicated that transduc-

    tion and action potential initiation both start at the distal end of the dendrite.

    More recent work has directly observed action potentials flowing along the

    dendrite from the distal tips (Gingl et al., 2004).

    Although all these experiments support a distal location for the mechan-

    osensitive channels, they cannot provide a more accurate position than

    somewhere within about 50 m from the end of the dendrite. The basalbody occurs at the distal end of the dendritic enlargement in VS3 neurons

    (Fig. 1), which is close to the lymph space. More accurate localization will

    probably have to wait for better anatomical evidence such as antibodies to

    the mechanosensitive channels.

    C. Mechanosensitive Channel Conductance, Density, and pH Sensitivity

    Singlechannel recordings of the mechanosensitive channels have not yet

    been achieved. Patch clamp recording from VS3 neurons is complicated

    by their locationwithin a hypodermalmembrane and extensive glial wrappings.

    The probable location of the channels near the tip of the sensory dendrite adds

    further diYculty. An alternative approach is to measure the variance, or noise,

    of the total receptor current to estimate the singlechannel conductance and

    number of channels (Traynelis and Jaramillo, 1998). This approach requires

    current variance measurements over a range of diVerent current amplitudes,

    which can be achieved by varying the stimulus used to open the channels being

    investigated. In VS3 neurons the receptor current adapts slowly after a step

    indentation of the slit, and this natural change in current was used to estimate

    the mechanosensitive channel properties.

    For a single group of identical ion channels, the total variance, s2, of the

    current flowing through a membrane is given by:

    s2 s20 IV Eg I2=N 1

    where s02 is the background variance due to other sources, I is total mem-

    brane current, V is the voltage across the membrane, E is the equilibrium

    potential of the ions flowing through the channel, is the singlechannelconductance, and N is the number of channels in the membrane. Given the

    singlechannel conductance and number of channels, the open probability of

    the channels can be calculated from:

    Po I

    NV Eg2

    1. Mechanotransduction in Spiders 9

  • Hoger and French (1999a) showed that the mechanosensitive channels

    were almost completely open at the start of a step indentation, but then

    closed with several time constants over a period of several minutes (Fig. 4).

    Their singlechannel conductance estimate was about 7 pS and the number

    of channels per neuron was about 470. Neither of these parameters was

    sensitive to pH (Hoger and French, 2002). However, acid conditions signifi-

    cantly raised the open probability of the channels, and hence the overall

    receptor current.

    From the estimated singlechannel conductance and number of channels,

    total mechanosensitive conductance was calculated to be about 3.5 nS in a

    single VS3 neuron. However, independent estimates of total charge flowing

    during a step indentation gave a significantly higher estimate of about 15 nS

    (Gingl and French, 2003). A possible cause of this diVerence lies in the cable

    properties of the sensory dendrite. The measured length constant of the

    sensory dendrites is about 200 m, which is comparable to the physicallength of the dendrites (Gingl and French, 2003). Although the noise mea-

    surements were made at the neuronal resting potential to minimize the

    current requirements of the voltage clamp, it is possible that the current

    flowing through the mechanosensitive channels at the dendrite tip could

    depolarize the membrane beyond the control of the voltage clamp in the

    soma. This would reduce the estimated receptor current and its variance.

    *

    1.0

    0.0

    1.0

    0.0

    Popen

    0 40Time (s)

    pH 8 pH 5

    n = 9 n = 23

    FIGURE 4 Noise analysis and pH sensitivity of VS3 receptor current. Step indentations of

    the slits lasting 40 s produced a slowly adapting receptor current. Noise analysis was used to

    estimate the number of mechanosensitive ion channels, singlechannel conductance, and

    channel open probability (Popen) during the step. Traces show Popen for a typical neuron at

    pH 8 (approximately normal conditions) and at pH 5. Inset shows mean values of Popen at 36

    s after the step under normal and acid conditions.Asterisk indicates p < 0.05. On the basis of data

    from Hoger and French (2002).

    10 French and Torkkeli

  • Therefore, the singlechannel conductance of the mechanosensitive channels

    could be 20 pS or more. This would be in better agreement with estimates

    from mammalian auditory hair cells based on singlechannel recordings,

    which are as high as 100 pS (Fettiplace et al., 1992).

    D. Temperature Sensitivity of Mechanosensitive Channels

    Mechanotransduction has been found to be more thermally sensitive than

    would be predicted from simple ion channel conductance in a range of

    vertebrate and invertebrate sensory receptors (reviewed in Hoger and

    French, 1999b). Most of these measurements were made on the action

    potential signals from sensory receptors so that the location of temperature

    sensitivity could not be clearly established. The VS3 organ provided the first

    direct measure of temperature sensitivity in the receptor current (Hoger and

    French, 1999b). These data were wellfitted by the Arrhenius rate equation

    to give a mean activation energy of 23 kcal/mol (97 kJ/mol or Q10 3.2 at20C). This is the highest activation energy measured for mechanotransduc-

    tion, although close to measurements in other systems (Hoger and French,

    1999b). It confirms the general finding that mechanotransduction involves a

    significant energy barrier, comparable to the energy required to break a

    covalent chemical bond. The reason for this relatively high activation energy

    is not clear but is probably associated with the mechanism that links mecha-

    nical stimulus to channel opening. It is much higher than the activation

    energy required for ionic movement through a waterfilled channel or for the

    production of action potentials by voltageactivated ion channels.

    E. Molecular Characterization of Spider Mechanosensitive Channels

    Two major groups of ion channel molecules have been associated with

    sensory mechanotransduction. Members of the transient receptor poten-

    tial (TRP) family of channels have been implicated in a range of sensory

    functions of both vertebrates and invertebrates, including phototransduc-

    tion, thermal transduction, mechanotransduction, pain, and osmosensation

    (Minke and Cook, 2002; Corey, 2003; Maroto et al., 2005; Montell, 2005;

    Dhaka et al., 2006; Kwan et al., 2006). TRP channels have been strongly

    linked to hearing and touch in Drosophila (Kim et al., 2003; Gong et al.,

    2004) and to touch in Caenorhabditis elegans (Goodman and Schwarz, 2003;

    Li et al., 2006). TRP1 channels have been found in vertebrate pain receptors

    (Kwan et al., 2006), as well as mouse, bullfrog, and zebrafish inner ear hair

    receptors (Corey, 2003), appearing at the same embryonic stage as sound

    1. Mechanotransduction in Spiders 11

  • sensitivity in mice (Lewin and Moshourab, 2004). However, a knockout

    mouse lacking TRP1 had an impaired response to painful stimuli but its

    hair cell transduction was not aVected (Kwan et al., 2006). None of the

    evidence yet gives clear proof that these channels are the primary source of

    the receptor current.

    The other channel family associated with mechanotransduction are the

    degenerin/acidsensitive/epithelial sodium channels (DEG/ASIC/ENaC),

    best known for the amilorideblockable epithelial sodium channels that con-

    duct sodium flux through a wide range of epithelia (Bianchi and Driscoll,

    2002). In C. elegans, two of the four proteins found only in mechanoreceptor

    cells are DEG molecules that have been proposed to form the core of the

    mechanotransduction channel, and the receptor current was carried by

    sodium ions (Goodman and Schwarz, 2003; Syntichaki and Tavernarakis,

    2004). A DEG gene family was also associated with mechanosensitivity in

    Drosophila larvae (Adams et al., 1998). In rodents, several members of the

    DEG family have been found in dorsal root ganglia and in fine nerve endings

    surrounding tactile hairs (Price et al., 2000). Knockout animals for one

    channel, BNC1, showed reductions, but not elimination, of mechanosensa-

    tion (Price et al., 2000), and none of these molecules have yet been identified

    in well known skin mechanoreceptors, such as Pacinian corpuscles or RuYni

    endings.

    Although the molecular evidence favors TRP channels in Drosophila

    mechanosensation (Kim et al., 2003), all the data from spider slit sensilla is

    more supportive of ASIC channels. The receptor current is highly selective

    for sodium and blocked by amiloride (Hoger et al., 1997). Mechanosensitive

    channel open probability is strongly increased at low pH (Hoger and French,

    2002). These are all characteristic properties of ASIC channels. In contrast,

    TRP channels are quite strongly associated with calcium signaling, and at

    least some sensory TRP channels are calcium permeable (Montell, 2005),

    whereas spider mechanosensitive channels are probably not permeable to

    calcium (Hoger et al., 2005).

    Two other commonly proposed features of sensory mechanically acti-

    vated channels are heteromeric construction and connections to extracellular

    and intracellular structural proteins. Evidence from several preparations

    indicates that multiple proteins are required to form functioning eukary-

    otic mechanically activated channels, and this may explain the diYculty of

    demonstrating mechanosensitivity from proteins expressed in oocytes or

    other systems (Hamill and McBride, 1996; Emtage et al., 2004; Syntichaki

    and Tavernarakis, 2004). Mechanical connections to cytoskeletal and extra-

    cellular matrix structures have been proposed by several lines of evidence,

    including the amino acid sequences of proposed channel molecules (Emtage

    et al., 2004). It has also been argued that lipid membrane alone could not

    12 French and Torkkeli

  • provide enough force to open a protein channel (Sachs, 1997). Microtubules

    are often prominent in mechanoreceptor endings, and in some cases have

    been suggested to form a cytoskeletal anchor (Gillespie and Walker, 2001).

    Spider slit sensilla, like other arthropod cuticular mechanoreceptors, contain

    prominent arrangements of microtubules in the sensory dendrites that ex-

    tend to the distal tips, but mechanotransduction in VS3 neurons and some

    insect cuticular mechanoreceptors persists after pharmacological destruction

    of microtubules (French, 1988; Hoger and Seyfarth, 2001).

    VI. DYNAMIC PROPERTIES OF MECHANOTRANSDUCTION AND

    ACTION POTENTIAL ENCODING

    Recordings of action potentials from spider tactile hairs and trichobothria

    show neurons that are normally silent, signaling brief touching or vibration

    (Barth, 2004). Slit sensilla neurons are also silent in their resting condition and

    respond preferentially to rapid changes. Each slit is innervated by two neu-

    rons that have diVerent dynamic properties. Type A neurons are very rapidly

    adapting, giving only one or two action potentials at the start of a step

    indentation, while Type B neurons give a longer burst of action potentials

    (Fig. 5). This pattern has been observed in both VS3 and HS10 lyriform

    organs (Seyfarth and French, 1994; Gingl et al., 2006) so it probably

    generalizes to most or all of the slit sensilla.

    50 mV

    10 mV

    100 msType A Type B

    FIGURE 5 Spider slit sensilla are innervated by pairs of functionally diVerent neurons.

    Intracellular recordings are shown from the two neuron types in a VS3 preparation receiving

    step indentations of 150ms duration. Upper traces show normal action potential responses

    from Types A (left) and B (right) neurons. Lower traces show receptor potentials produced by

    similar steps after action potentials were blocked by treatment with tetrodotoxin. On the basis of

    data from Juusola and French (1998).

    1. Mechanotransduction in Spiders 13

  • Recordings of the receptor current (Fig. 2) or the receptor potential

    (Fig. 5) do not show such strong adaptation or such a diVerence between

    the two neuron types (Juusola and French, 1998). Receptor potential in the

    Type A neurons does adapt more rapidly than in Type B neurons, but the

    diVerence is less dramatic than the firing behavior. This diVerence in action

    potential encoding can also be seen with direct electrical stimulation of the

    neurons, and can be explained by diVerences in the inactivation properties of

    the voltageactivated sodium channels that cause the initial phase of the

    action potentials (Torkkeli and French, 2002).

    The time course of the receptor current and potential must be controlled

    by the combination of mechanical coupling components and mechanosensi-

    tive ion channels. However, little is known about the dynamic properties of

    either. Somatic measurements indicate that the receptor current decays with

    at least two time constants (Fig. 2), and voltage jump experiments indicated

    that there are larger, very transient components occurring in the distal

    dendrites (Gingl and French, 2003). It is possible that these diVerent time

    constants represent separate filtering by the mechanical components and the

    mechanosensitive ion channels. The existing evidence is compatible with

    the most parsimonious model of transduction, that is, that a single type of

    mechanosensitive channel is present in both Types A and B neurons.

    VII. CALCIUM SIGNALING DURING TRANSDUCTION BY

    SPIDER MECHANORECEPTORS

    The membranes of VS3 neurons contain lowvoltageactivated calcium

    selective ion channels (Sekizawa et al., 2000). Measurements of intracellular

    calcium concentration during mechanical stimulation of the slits showed

    that calcium rises from a resting level of about 400 nM to a maximum level

    of about 2 M during rapid action potential firing (Hoger et al., 2005).These experiments failed to show any change in calcium concentration with-

    out action potentials, even when there was a receptor potential of 10 mV

    amplitude or more, confirming that the mechanosensitive ion channels are

    not significantly permeable to calcium. They also failed to show any release

    of calcium from internal stores. The amount of calcium entering during

    action potential firing was compatible with the estimated conductance via

    voltageactivated calcium channels.

    These data raise the question of what role the elevation of calcium plays

    during normal sensory transduction. There are no known calciumsensitive

    ion channels in VS3 neurons, and blockade of calcium entry does not reliably

    aVect action potential firing. Calcium rose by similar amounts throughout

    the VS3 neurons, but with diVerent time courses in diVerent regions

    14 French and Torkkeli

  • (Fig. 6), suggesting that calcium channels are distributed throughout the cells.

    One possible role for calcium would be regulation of the mechanosensitive

    channels. Calcium ions play major roles in controlling the dynamic properties

    of auditory hair cells, and at least some of the time constants involved seem to

    depend on intracellular actions of calcium on mechanosensitive ion channels

    (Ricci et al., 2005).

    VIII. SYNAPTIC MODULATION OF SPIDER MECHANORECEPTORS

    An interesting feature of arachnid mechanoreceptors is that even their

    most peripherally located parts receive extensive and complex eVerent inner-

    vation (Foelix, 1975; FabianFine et al., 2002), allowing an early modulation

    of the neuronal responses to mechanical stimuli. Several fine eVerent fibers

    in the legs of C. salei extend along the sensory nerves all the way to the tips

    of the sensory dendrites. They form many types of synaptic contacts with

    the sensory neurons, the glial cells that enwrap the sensory neurons, and they

    also synapse with other eVerents (FabianFine et al., 2002). The eVerent

    fibers have been shown to contain a variety of transmitters, including

    aminobutyric acid (GABA), glutamate, acetylcholine (ACh), and octopa-mine (Fig. 7; FabianFine et al., 2002; Widmer et al., 2005), and the mechan-

    osensory neurons respond to these transmitters (Panek et al., 2002; Panek

    500 nM

    50 s

    Distaldendrite

    Middendrite

    Soma Soma

    FIGURE 6 Calcium concentration rises significantly in VS3 neurons when they are firing.

    Traces show calcium elevations in diVerent regions during stimulation at 10 action potentials per

    second. Resting calcium concentration was about 400 nM in all regions and the increases in

    diVerent regions were not significantly diVerent. However, the time course of elevation was

    significantly slower in the soma, as shown by the traces. On the basis of data from Hoger et al.

    (2005).

    1. Mechanotransduction in Spiders 15

  • and Torkkeli, 2005; Widmer et al., 2005, 2006). In addition, antibodies

    against transmitter receptors labeled specific sites on the sensory neurons

    (Panek et al., 2003, 2005;Widmer et al., 2005, 2006; Fig. 7).

    GABA and glutamate both act on inhibitory ionotropic receptors that are

    Clgated ion channels. Although both transmitters blocked VS3 neurons

    responses to mechanical stimuli, GABA had a significantly stronger eVect

    than glutamate (Panek and Torkkeli, 2005). However, GABA only inhibited

    axonal action potentials while the glutamate eVect involved both dendritic

    and axonal action potentials and it also reduced the receptor current ampli-

    tude (Gingl et al., 2004; Panek and Torkkeli, 2005). Thus, glutamatergic

    eVerents may control the cellular response to mechanical stimuli at earlier

    stages than GABAergic eVerents. The VS3 neurons also have metabotropic

    GABAB receptors, concentrated on the most distal parts of the cell bodies

    and on the dendrites (Panek et al., 2003). Agonists of these receptors

    modulated voltageactivated calcium and potassium currents, allowing a

    longer lasting modulatory eVect.

    Inhibitory glutamate receptor

    GABAB receptor

    Octopamine receptor

    Inhibitory GABA receptor

    DendriteAxon

    GABA

    Octopamine

    Glutamate

    Soma

    ACh

    mACh receptor

    Inhibitory ACh receptor

    ?

    ?

    ?

    ??

    ?

    ?

    ?

    AChE

    +

    +

    Inhibitory+ Excitatory

    ?

    +

    ????

    FIGURE 7 Schematic illustration of the arrangement of eVerent neurons and transmitter

    receptors on a Type A spider VS3 neuron based on immunocytochemical and electrophysio-

    logical evidence. The eVerent fibers contain GABA, glutamate, octopamine, and ACh. The

    sensory neurons have inhibitory ionotropic GABA and glutamate receptors and excitatory

    octopamine receptors. Type A neurons also have inhibitory ionotropic ACh receptors and they

    express acetylcholine esterase (AChE) activity. In addition, metabotropic GABAB and musca-

    rinic ACh receptors are found in all VS3 neurons, but their physiological functions are

    unknown. Glutamate and mACh receptors are also present in the eVerent fibers. On the basis

    of data from FabianFine et al. (2002), Panek et al. (2002, 2003, 2005), Gingl et al. (2004), Panek

    and Torkkeli (2005), Widmer et al. (2005, 2006).

    16 French and Torkkeli

  • Application of octopamine, the invertebrate analogue of noradrenaline,

    enhanced trichobothria neuron sensitivity to mechanical stimuli (Widmer

    et al., 2005). Immunocytochemical evidence indicated that one eVerent fiber

    containing octopamine innervated each mechanosensory neuron in the spider

    leg and that octopamine receptors were concentrated at and close to the axon

    hillock (Widmer et al., 2005). These findings suggest that octopamine acts as

    a transmitter rather than a neurohormone on spider mechanoreceptors,

    controlling each sensory neuron individually.

    These recent findings only unravel a small part of the complex synaptic

    mechanisms that control the sensitivity and gain of spider mechanosensory

    neurons. For example, we still know very little about the cholinergic inner-

    vation that involves both muscarinic ACh receptors and ionotropic inhibi-

    tory receptors and is distinctly diVerent in the two diVerent types of VS3

    neurons.

    IX. CONCLUSIONS

    Spider mechanoreceptors have yielded a great deal of information about

    their mechanosensitive ion channels and their mechanisms of activation and

    modulation. However, much remains to be discovered. The electrophysio-

    logical data from slit sensilla suggest that the channel molecules are related

    to ASIC channels and they are probably located near the tips of the sensory

    dendrites. The relatively low numbers of channel molecules per cell are one

    reason why molecular characterization has so far proved elusive as it has in

    other mechanoreceptor systems. However, the spider preparations should

    continue to provide useful models for identifying the molecular basis of

    mechanosensation and this knowledge can be expected to assist the broader

    investigation of this crucial sense in animals and humans.

    AcknowledgmentsWe thank Ewald Gingl, Ulli Hoger, Mikko Juusola, Izabela Panek, ShannonMeisner, Ernst

    August Seyfarth, and Alexandre Widmer for all their contributions to work described here.

    Research in our laboratories has been funded by the Canadian Institutes of Health Research,

    the Natural Sciences and Engineering Council of Canada, NATO, the Canadian Foundation

    for Innovation, the Nova Scotia Research and Innovation Trust, and the Dalhousie Medical

    Research Foundation.

    ReferencesAdams, C. M., Anderson, M. G., Motto, D. G., Price, M. P., Johnson, W. A., and Welsh, M. J.

    (1998). Ripped pocket and pickpocket, novel DrosophilaDEG/ENaC subunits expressed in

    early development and in mechanosensory neurons. J. Cell Biol. 140, 143152.

    Barth, F. G. (1985). Slit sensilla and the measurement of cuticular strains. In Neurobiology of

    Arachnids (F. G. Barth, ed.), pp. 162188. SpringerVerlag, Berlin.

    1. Mechanotransduction in Spiders 17

  • Barth, F. G. (2001). A Spiders World: Senses and Behavior. SpringerVerlag, Berlin.

    Barth, F. G. (2004). Spider mechanoreceptors. Curr. Opin. Neurobiol. 14, 415422.

    Barth, F. G., and Libera, W. (1970). Ein Atlas der Spaltsinnesorgane von Cupiennius saleiKeys.

    Chelicerata (Araneae). Z. Morph. Tiere. 68, 343369.

    Bianchi, L., and Driscoll, M. (2002). Protons at the gate: DEG/ENaC ion channels help us feel

    and remember. Neuron 34, 337340.

    Corey, D. P. (2003). New TRP channels in hearing and mechanosensation. Neuron 39, 585588.

    Dhaka, A., Viswanath, V., and Patapoutian, A. (2006). TRP ion channels and temperature

    sensation. Annu. Rev. Neurosci. 29, 135161.

    Emtage, L., Gu, G., Hartwieg, E., and Chalfie, M. (2004). Extracellular proteins organize the

    mechanosensory channel complex inC. elegans touch receptor neurons.Neuron 44, 795807.

    FabianFine, R., Seyfarth, E.A., andMeinertzhagen, I. A. (2002). Peripheral synaptic contacts at

    mechanoreceptors in arachnids and crustaceans: Morphological and immunocytochemical

    characteristics. Microsc. Res. Tech. 58, 283298.

    Fettiplace, R., Crawford, A. C., and Evans, M. G. (1992). The hair cells mechanoelectrical

    transducer channel. Ann. NY Acad. Sci. 656, 111.

    Foelix, R. F. (1975). Occurrence of synapses in peripheral sensory nerves of arachnids. Nature

    254, 146148.

    Foelix, R. F. (1996). Biology of Spiders. Oxford University Press, New York.

    French, A. S. (1988). Transduction mechanisms of mechanosensilla. Annu. Rev. Entomol. 33,

    3958.

    Gillespie, P. G., and Walker, R. G. (2001). Molecular basis of mechanosensory transduction.

    Nature 413, 194202.

    Gingl, E., and French, A. S. (2003). Active signal conduction through the sensory dendrite of a

    spider mechanoreceptor neuron. J. Neurosci. 23, 60966101.

    Gingl, E., French, A. S., Panek, I., Meisner, S., and Torkkeli, P. H. (2004). Dendritic

    excitability and localization of GABAmediated inhibition in spider mechanoreceptor

    neurons. Eur. J. Neurosci. 20, 5965.

    Gingl, E., Burger, A. M., and Barth, F. G. (2006). Intracellular recording from a spider

    vibration receptor. J. Comp. Physiol. A 192, 551558.

    Gong, Z., Son, W., Chung, Y. D., Kim, J., Shin, D. W., McClung, C. A., Lee, Y., Lee, H. W.,

    Chang, D. J., Kaang, B. K., Cho, H., Oh, U., et al. (2004). Two interdependent TRPV

    channel subunits, inactive and Nanchung, mediate hearing in Drosophila. J. Neurosci. 24,

    90599066.

    Goodman, M. B., and Schwarz, E. M. (2003). Transducing touch in Caenorhabditis elegans.

    Annu. Rev. Physiol. 65, 429452.

    Grunert, U., and Gnatzy, W. (1987). K and Ca in the receptor lymph of arthropod cuticular

    receptors. J. Comp. Physiol. A 161, 329333.

    Hamill, O. P., and McBride, D. W. (1996). A supramolecular complex underlying touch

    sensitivity. Trends Neurosci. 19, 258261.

    Hoger, U., and French, A. S. (1999a). Estimated singlechannel conductance of mechanically

    activated channels in a spider mechanoreceptor. Brain Res. 826, 230235.

    Hoger, U., and French, A. S. (1999b). Temperature sensitivity of transduction and action

    potential conduction in a spider mechanoreceptor. Pflugers Arch. 438, 837842.

    Hoger, U., and French, A. S. (2002). Extracellular acid increases the open probability of

    transduction channels in spider mechanoreceptors. Eur. J. Neurosci. 16, 23112316.

    Hoger, U., and Seyfarth, E.A. (2001). Structural correlates of mechanosensory transduction

    and adaptation in identified neurons of spider slit sensilla. J. Comp. Physiol. A 187,

    727736.

    18 French and Torkkeli

  • Hoger, U., Torkkeli, P. H., Seyfarth, E.A., and French, A. S. (1997). Ionic selectivity of

    mechanically activated channels in spider mechanoreceptor neurons. J. Neurophysiol. 78,

    20792085.

    Hoger, U., Torkkeli, P. H., and French, A. S. (2005). Calcium concentration changes during

    sensory transduction in spider mechanoreceptor neurons. Eur. J. Neurosci. 22, 31713178.

    Juusola, M., and French, A. S. (1998). Adaptation properties of two types of sensory neurons in

    a spider mechanoreceptor organ. J. Neurophysiol. 80, 27812784.

    Juusola, M., Seyfarth, E.A., and French, A. S. (1994). Sodiumdependent receptor current in a

    new mechanoreceptor preparation. J. Neurophysiol. 72, 30263028.

    Kim, J., Chung, Y. D., Park, D. Y., Choi, S., Shin, D. W., Soh, H., Lee, H. W., Son, W., Yim, J.,

    Park, C. S., Kernan, M. J., and Kim, C. (2003). A TRPV family ion channel required for

    hearing in Drosophila. Nature 424, 2829.

    Kwan, K. Y., Allchorne, A. J., Vollrath, M. A., Christensen, A. P., Zhang, D. S., Woolf, C. J.,

    and Corey, D. P. (2006). TRPA1 contributes to cold, mechanical, and chemical nociception

    but is not essential for haircell transduction. Neuron 50, 277289.

    Lewin, G. R., and Moshourab, R. (2004). Mechanosensation and pain. J. Neurobiol. 61, 3044.

    Li, W., Feng, Z., Sternberg, P. W., and Xu, X. Y. (2006). A C. elegans stretch receptor neuron

    revealed by a mechanosensitive TRP channel homologue. Nature 440, 684687.

    Maroto, R., Raso, A.,Wood, T. G., Kurosky, A.,Martinac, B., andHamill, O. P. (2005). TRPC1

    forms the stretchactivated cation channel in vertebrate cells. Nat. Cell Biol. 7, 179185.

    Minke, B., and Cook, B. (2002). TRP channel proteins and signal transduction. Physiol. Rev.

    82, 429472.

    Montell, C. (2005). Drosophila TRP channels. Pflugers Arch. 451, 1928.

    Panek, I., and Torkkeli, P. H. (2005). Inhibitory glutamate receptors in spider peripheral

    mechanosensory neurons. Eur. J. Neurosci. 22, 636646.

    Panek, I., French, A. S., Seyfarth, E.A., Sekizawa, S.I., and Torkkeli, P. H. (2002). Peripheral

    GABAergic inhibition of spider mechanosensory aVerents. Eur. J. Neurosci. 16, 96104.

    Panek, I., Meisner, S., and Torkkeli, P. H. (2003). The distribution and function of GABABreceptors in spider peripheral mechanosensilla. J. Neurophysiol. 90, 25712580.

    Panek, I., Meisner, S., and Torkkeli, P. H. (2005). Glutamate acts on inhibitory receptors on

    spider peripheral mechanoreceptors. Soc. Neurosci. Abstr. 31, 296.7.

    Patil, B., Prabhu, S., and Rajashekhar, K. P. (2006). Lyriform slit sense organs on the pedipalps

    and spinnerets of spiders. J. Biosci. 31, 7584.

    Price, M. P., Lewin, G. R.,McIlwrath, S. L., Cheng, C., Xie, J., Heppenstall, P. A., Stucky, C. L.,

    Mannsfeldt, A. G., Brennan, T. J., Drummond, H. A., Qiao, J., Benson, C. J., et al. (2000).

    The mammalian sodium channel BNC1 is required for normal touch sensation.Nature 407,

    10071011.

    Ricci, A. J., Kennedy, H. J., Crawford, A. C., and Fettiplace, R. (2005). The transduction

    channel filter in auditory hair cells. J. Neurosci. 25, 78317839.

    Rick, R., Barth, F. G., and Pawel, A. V. (1976). Xray microanalysis of receptor lymph in a

    cuticular arthropod sensillum. J. Comp. Physiol. 110, 8995.

    Sachs, F. (1997). Mechanical transduction by ion channels: How forces reach the channel.

    In Cytoskeletal Regulation of Membrane Function (S. C. Froehner, ed.), pp. 209218.

    Rockefeller University Press, New York.

    Sekizawa, S., French, A. S., and Torkkeli, P. H. (2000). Lowvoltageactivated calcium current

    does not regulate the firing behavior in paired mechanosensory neurons with diVerent

    adaptation properties. J. Neurophysiol. 83, 746753.

    Seyfarth, E.A. (1985). Spider proprioception: Receptors, reflexes, and control of locomotion.

    In Neurobiology of Arachnids (F. G. Barth, ed.), pp. 230248. SpringerVerlag, Berlin.

    1. Mechanotransduction in Spiders 19

  • Seyfarth, E.A., and French, A. S. (1994). Intracellular characterization of identified sensory

    cells in a new spider mechanoreceptor preparation. J. Neurophysiol. 71, 14221427.

    Syntichaki, P., and Tavernarakis, N. (2004). Genetic models of mechanotransduction: The

    nematode Caenorhabditis elegans. Physiol. Rev. 84, 10971153.

    Thurm, U., and Kuppers, J. (1980). Epithelial physiology of insect sensilla. In Insect Biology in

    the Future (M. Locke and D. Smith, eds.), pp. 735763. Academic Press, New York.

    Torkkeli, P. H., and French, A. S. (2002). Simulation of diVerent firing patterns in paired spider

    mechanoreceptor neurons: The role of Na channel inactivation. J. Neurophysiol. 87,

    13631368.

    Traynelis, S. F., and Jaramillo, F. (1998). Getting the most out of noise in the central nervous

    system. Trends Neurosci. 21, 137145.

    Widmer, A., Hoger, U., Meisner, S., French, A. S., and Torkkeli, P. H. (2005). Spider peripheral

    mechanosensory neurons are directly innervated and modulated by octopaminergic

    eVerents. J. Neurosci. 25, 15881598.

    Widmer, A., Panek, I., Hoger, U., Meisner, S., French, A. S., and Torkkeli, P. H. (2006).

    Acetylcholine receptors in spider peripheral mechanosensilla. J. Comp. Physiol. A 192,

    8595.

    20 French and Torkkeli

  • CHAPTER 2

    Ion Channels for Mechanotransductionin the Crayfish Stretch Receptor

    Bo Rydqvist

    Department of Physiology and Pharmacology, Karolinska Institutet,

    SE171 77 Stockholm, Sweden

    I. Overview

    II. Introduction

    III. Morphology of the SRO

    IV. Functional Properties

    A. General Behavior

    B. Viscoelastic Properties of the Receptor Muscles

    C. MSCs in the Receptor Neurons

    D. Macroscopic Receptor Currents in the Stretch Receptor Neurons

    E. Pharmacology of the Crayfish MSCs

    F. VoltageGated Ion Channels and the Generation of Impulse Response

    G. Adaptation: A Multifactor Property

    V. Summary and Discussion of Future Research Directions

    References

    I. OVERVIEW

    Mechanosensitivity is found in almost every cell in all organisms from

    bacteria to vertebrates and covers a wide spectrum of function from osmo-

    sensing to mechanical sensing in the specialized receptors like the hair cells of

    the cochlea. The molecular substrate for such mechanosensitivity is thought

    to be mechanosensitive ion channels (MSCs). Since most development

    regarding the molecular aspects of the MSC has been made in nonsensory

    or sensory systems which have not been accessible to recordings from ion

    channels, it is important to focus on mechanosensitivity of sensory organs

    where their functional importance is undisputed. The stretch receptor organ

    Current Topics in Membranes, Volume 59 1063-5823/07 $35.00Copyright 2007, Elsevier Inc. All right reserved. DOI: 10.1016/S1063-5823(06)59002-7

  • (SRO) of the crustaceans is a suitable preparation for such studies. Each

    organ contains two receptors: one slowly and one rapidly adapting receptor

    neurons. The primary mechanosensitivity is generated by two types of MSC

    of hitherto unknown molecular type located in the neuronal dendrites, which

    are inserted into a receptor muscle fiber. In addition to the MSCs, the

    neurons contain voltagegated Na channels which seem to be diVerently

    located in the slowly and rapidly adapting neurons. Finally, at least three

    types of voltagegated K channels are present in the sensory neurons, the

    location of which is not known. The spatial distribution of ion channels

    and the kinetics of the channels, together with the viscoelastic properties

    of the receptor muscles, determine the overall transducer properties and

    impulse firing of the two receptor neurons including their typical adaptive

    characteristics.

    II. INTRODUCTION

    The crustacean SRO has been amajor preparation for the study ofmechan-

    otransduction both on the macroscopic and on the ion channel levels. The

    SRO is considered to be an organ analogous to themammalianmuscle spindle

    organ that is instrumental for proper skeletal muscle function. The receptor

    organ was first described in the lobster by the PolishBritish zoologist

    Alexandrowicz (1951, 1967). Later, Florey and Florey (1955) described the

    same type of organ in the crayfish (Astacus fluviatilis presently namedAstacus

    astacus). Identical and similar muscle receptor organs can also be found in

    a number of other invertebrate phyla such as Mollusca, Chelicerata, and

    Uniramia (for a review see Rydqvist, 1992). The importance of this organ,

    and its accessibility relative to the human muscle spindle, and mechanotrans-

    duction in general, was soon acknowledged and triggered a number of electro-

    physiological studies in several laboratories (Wiersma et al., 1953; KuZer,

    1954; Eyzaguirre and KuZer, 1955a,b; Edwards and Ottoson, 1958).

    The mechanosensory neurons of the SRO of the crayfish are of the

    nonciliated type and are diVerent from the ciliated type represented by the

    classical hair cells in the hearing organs. Most investigations regarding

    the molecular aspects of the MSC have been made in nonsensory systems,

    and it is thus important to focus on mechanosensitivity of sensory organs

    where the functional importance of these channels is undisputed. The SRO

    of the crustaceans is a suitable preparation for such studies. The SRO is

    experimentally accessible to mechanical manipulation and electrophysiolog-

    ical recordings using intracellular microelectrodes or patch clamp techniques

    for ion channel analysis, although the latter technique is not without prob-

    lems since the sensory neuron is covered by supporting glial cells. It is,

    22 Bo Rydqvist

  • however, relatively easy to inject substances into the neuron, which makes

    the neuron accessible to measurements using fluorescent probes.

    In the present chapter, I have focused on the overall function of the

    SRO in the crayfish stretch receptor including results obtained with struc-

    tural techniques, classical electrophysiology, and patch clamp techniques.

    The main emphasis will be on the mechanotransduction processes and the

    ion channels involved in the SRO of the species A. astacus, Pacifastacus

    leniusculus, Procambarus clarkii, and Orconectes limosus.

    III. MORPHOLOGY OF THE SRO

    Since the crayfish stretch receptor is a genuine mechanosensory organ

    with several ion channels involved in the overall mechanotransduction, a

    brief description of the SRO seems relevant. The SRO has two sensory

    neurons, each connected to a receptor muscle, located in the extensor

    muscles of the abdomen (Florey and Florey, 1955; Purali, 2005). The sensory

    neuron is of the multipolar type (Fig. 1) with its dendrites inserted into the

    central (intercalated) part of the receptor muscle which consists of only one

    muscle cell (TaoCheng et al., 1981). The receptor muscles insert on consec-

    utive segments and the aVerent axons from the neurons join the dorsal

    segmental nerve to the ventral ganglion. The SROs also receive eVerent

    innervations: (1) one or two motor axons to the receptor muscle cell and

    (2) two or three accessory axons conveying inhibitory signals to the receptor

    muscle and the sensory neurons (Alexandrowicz, 1951, 1967; Elekes and

    Florey, 1987a,b). Functionally, the receptors are activated when stretched by

    flexion of the abdomen or contraction of the receptor muscles (KuZer, 1954)

    and the SROs are involved in the control of the extensor muscles.

    In the crayfish, both the slowly and the rapidly adapting receptor muscles

    consist of a single muscle fiber that is divided by invagination of the cell

    membrane into numerous cytoplasmic processes in the central region of the

    muscle, the intercalated tendon, which is mainly made up of collagen. Some

    of the myofibrils insert in the intercalated tendon but some pass this region.

    The slowly adapting muscle is in the order of 3080 mm in the central region

    but considerably thinner in the distal ends. The rapidly adapting muscle has

    a more even diameter and is thicker 70150 mm (Komuro, 1981).

    The sensory neurons are large (30100 mm) multiterminal cells of mainly

    pyramidal or fusiform shape. They contain a nucleus (ca. 10 mm) with a clear

    nucleolus. The dendrites branch about four to five times and intermingle

    with the connective tissue and muscle strands in the intercalated tendon. The

    fine terminal branches are about 2mm long and about 0.l mm in diameter and

    are devoid of mitochondria (TaoCheng et al., 1981). The axon is in the

    2. Stretch Receptor Mechanotransduction 23

  • order of 30 mm in diameter. The receptor neurons of the crayfish have several

    layers of sheet cells that surround them except for the dendritic tips.

    The fine structure of the inhibitory synapses has been investigated by

    several authors (Elekes and Florey, 1987a,b) using serial sectioning and

    immunohistochemical technique, which have revealed a complex array of

    GABAergic inhibitory synapses on the axon, neuron, and muscle fibers and

    also reciprocal synapses on the inhibitory axon.

    IV. FUNCTIONAL PROPERTIES

    A. General Behavior

    The receptors are activated (stretched) by flexion of the abdomen or

    contraction of the receptor muscle. The receptors are involved in the motor

    control of the abdominal muscles and the physiological range is up to 40%

    of resting length (Alexandrowicz, 1951). The first measurements from the

    FIGURE 1 (A) Abdomen and thorax of crayfish. (B) Slowly (top) and rapidly (bottom)

    adapting receptor with typical action potential firing pattern as a result of a ramp and hold

    extension of the receptor muscle. (C) Confocal microscopic image of a slowly adapting neuron

    injected with Fluo4 (Bruton and Rydqvist, unpublished data), a recurrent fiber is present at left.

    (D) Drawing of stretch receptor neuron with proposed channel distribution. (A, B) Adapted

    with kind permission from Springer Science and Business Media (Rydqvist, 1992).

    24 Bo Rydqvist

  • receptor neuron were done by Wiersma et al. (1953) and KuZer (1954)

    and subsequent studies by Eyzaguirre and KuZer (1955a,b) and Edwards

    and Ottoson (1958) on lobster and crayfish showed that stretching the re-

    ceptor organs gave rise to a distinctive pattern of impulse discharge from the

    neurons. It was found that the firing properties of the two neurons were

    clearly diVerent, one neuron maintained firing as long as the stretch was

    applied (slowly adapting) whereas the other neuron generated a short high

    frequency discharge (rapidly adapting) at the onset of the stretch (Fig. 1B).

    The chain of events that leads from extension of the receptor muscle to

    action potential generation in the stretch receptor is represented by the steps

    outlined in Fig. 2. In a first step, the extension of the receptor muscle is

    Actionpotential

    50 mV

    100 nA

    100 kPa

    Na+

    K+

    cm gL

    100 ms

    ReceptorpotentialTTX

    Receptorcurrent

    SA-channels

    ViscoelasticReceptormuscle

    Voltage-gatedchannels

    Passivemembrane

    DendritesMuscletension

    Extension

    ReceptorpotentialTTX, 4AP,TEA

    FIGURE 2 Transduction processes in a stretch receptor neuron. Left: recorded responses of

    muscle tension, receptor current, receptor potential, and action potentials in response to a ramp

    and hold extension of the muscle. The receptor potential is seen both after block of Na

    channels with tetrodotoxin (TTX) and after additional block of K channels with tetraethyl-

    ammonium chloride (TEA) and 4aminopyridine (4AP). Right: functional blocks in transduc-

    tion. Stretchactivated channels; SA channels, MSC. Adapted from Swerup and Rydqvist, 1992

    with permission from Elsevier.

    2. Stretch Receptor Mechanotransduction 25

  • converted to tension in the muscle, which leads to deformation of the

    dendritic membrane of the sensory neuron. This opens nonselective mechan-

    osensitive (gated) ion channels (MSCs) permeable to Na, K, and Ca2

    ions producing an inward generator current (Erxleben, 1989). The transfor-

    mation from generator current to impulse response is a complex process

    determined by the passive membrane properties, that is capacitance (cm) and

    membrane resistance (rm or leak conductance gL), and the voltagegated ion

    conductances (gion) present in the neuron. At present only voltagegated K

    and Na channels and a Ca2activated K channel have been observed.

    Figure 2 shows the receptor potential after block by tetrodotoxin (TTX),

    4aminopyridine (4AP), and tetraethylammonium chloride (TEA), and the

    receptor potential after block with TTX only. In addition, the geometry of

    the cell and the spatial distribution of the diVerent ion channels will contrib-

    ute to the type of impulse response seen in the cell. The diVerence in respon-

    ses reflects the relativity of the concept of receptor potential (see discussion

    in Swerup and Rydqvist, 1992).

    B. Viscoelastic Properties of the Receptor Muscles

    In most mechanosensory cells, the accessory structures contribute to the

    overall behavior of the transducer function. In particular, it has been dis-

    cussed to what extent the accessory structures contribute to the adaptation

    in sensory cells. This is obvious in, for example, the Pacinian corpuscle. The

    two neurons in the SRO have diVerent adaptive properties and this could

    arise solely from possible diVerences in passive mechanical properties in the

    two receptor muscles. Earlier studies (KuZer, 1954) observed that the con-

    tractile properties indeed diVered; the rapidly adapting muscle had proper-

    ties resembling a fast twitch fiber, whereas the slowly adapting muscle

    behaved as a slo