02.12.2012 Views

Voltage-gated K+ channels in Drosophila photoreceptors ...

Voltage-gated K+ channels in Drosophila photoreceptors ...

Voltage-gated K+ channels in Drosophila photoreceptors ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

VOLTAGE-GATED K +<br />

CHANNELS IN DROSOPHILA<br />

PHOTORECEPTORS<br />

Biophysical study of neural cod<strong>in</strong>g<br />

MIKKO<br />

VÄHÄSÖYRINKI<br />

Faculty of Science,<br />

Department of Physical Sciences,<br />

Division of Biophysics,<br />

Biocenter Oulu,<br />

University of Oulu<br />

OULU 2004


MIKKO VÄHÄSÖYRINKI<br />

VOLTAGE-GATED K + CHANNELS IN<br />

DROSOPHILA PHOTORECEPTORS<br />

Biophysical study of neural cod<strong>in</strong>g<br />

Academic Dissertation to be presented with the assent of<br />

the Faculty of Science, University of Oulu, for public<br />

discussion <strong>in</strong> Oulunsali (Auditorium L5), L<strong>in</strong>nanmaa, on<br />

December 11th, 2004, at 12 noon<br />

OULUN YLIOPISTO, OULU 2004


Copyright © 2004<br />

University of Oulu, 2004<br />

Supervised by<br />

Professor Matti Weckström<br />

Reviewed by<br />

Professor Kristian Donner<br />

Professor Doekele Stavenga<br />

ISBN 951-42-7598-5 (nid.)<br />

ISBN 951-42-7599-3 (PDF) http://herkules.oulu.fi/isbn9514275993/<br />

ISSN 0355-3191 http://herkules.oulu.fi/issn03553191/<br />

OULU UNIVERSITY PRESS<br />

OULU 2004


Vähäsöyr<strong>in</strong>ki, Mikko, <strong>Voltage</strong>-<strong>gated</strong> K + <strong>channels</strong> <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong><br />

Biophysical study of neural cod<strong>in</strong>g<br />

Faculty of Science, Department of Physical Sciences, Division of Biophysics, University of Oulu,<br />

P.O.Box 3000, FIN-90014 University of Oulu, F<strong>in</strong>land, Biocenter Oulu, University of Oulu,<br />

P.O.Box 5000, FIN-90014 University of Oulu, F<strong>in</strong>land<br />

2004<br />

Oulu, F<strong>in</strong>land<br />

Abstract<br />

The activity of neurons is critically dependent upon the suite of voltage-dependent ion <strong>channels</strong><br />

expressed <strong>in</strong> their membranes. In particular, voltage-<strong>gated</strong> K + <strong>channels</strong> are extremely diverse <strong>in</strong> their<br />

function, contribut<strong>in</strong>g to the regulation of dist<strong>in</strong>ct aspects of neuronal activity by shap<strong>in</strong>g the voltage<br />

responses.<br />

In this study the role of K + <strong>channels</strong> <strong>in</strong> neural cod<strong>in</strong>g is <strong>in</strong>vesti<strong>gated</strong> <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong><br />

by us<strong>in</strong>g biophysical models with parameters derived from the electrophysiological experiments. Due<br />

to their biophysical properties, the Shaker <strong>channels</strong> attenuate the fast transients and amplify the<br />

slower signal components, enabl<strong>in</strong>g <strong>photoreceptors</strong> to use their voltage range more effectively. Slow<br />

delayed rectifier <strong>channels</strong>, shown to be encoded by the Shab gene, activate at high light <strong>in</strong>tensities,<br />

thereby attenuat<strong>in</strong>g the light-<strong>in</strong>duced depolarization and prevent<strong>in</strong>g response saturation. Activation<br />

of Shab <strong>channels</strong> also reduces the membrane time constant mak<strong>in</strong>g it possible to encode faster events.<br />

Interactions between the voltage-<strong>gated</strong> K + <strong>channels</strong> and the currents generated by the light<br />

<strong>in</strong>duced conductance (LIC) were <strong>in</strong>vesti<strong>gated</strong> dur<strong>in</strong>g naturalistic stimulation <strong>in</strong> wild type and Shaker<br />

mutant <strong>photoreceptors</strong>. It is shown that <strong>in</strong> addition to elim<strong>in</strong>at<strong>in</strong>g the Shaker current, the mutation<br />

<strong>in</strong>creased the Shab current and affected the current flow<strong>in</strong>g through the LIC. Part of these changes<br />

could be attributed to direct feedback from the Shaker <strong>channels</strong> via the membrane potential.<br />

However, it is suggested that also other changes may occur <strong>in</strong> the LIC due to mutation <strong>in</strong> K + <strong>channels</strong>,<br />

possibly dur<strong>in</strong>g photoreceptor development.<br />

Comparison of the Shaker and Shab mutant <strong>photoreceptors</strong> with the wild type revealed that a<br />

concurrent decrease <strong>in</strong> the steady-state <strong>in</strong>put resistance followed from deletion of the voltage-<strong>gated</strong><br />

K + <strong>channels</strong>. This allowed partial compensation of the compression and saturation caused by the loss<br />

of Shaker <strong>channels</strong> and it ma<strong>in</strong>ta<strong>in</strong>ed the characteristics of the light-voltage relationship <strong>in</strong> Shab<br />

mutant <strong>photoreceptors</strong>. However, wild type properties were not fully restored <strong>in</strong> either mutant.<br />

Indeed, decreased <strong>in</strong>put resistance results <strong>in</strong> reduced efficiency of neural process<strong>in</strong>g, assessed by the<br />

metabolic cost of <strong>in</strong>formation.<br />

Results of this study demonstrate the importance of the voltage-<strong>gated</strong> K + <strong>channels</strong> for neural<br />

cod<strong>in</strong>g precision and highlight the robustness of neuronal <strong>in</strong>formation process<strong>in</strong>g ga<strong>in</strong>ed through<br />

regulation of the electrical properties.<br />

Keywords: Hodgk<strong>in</strong>-Huxley model, metabolic costs, naturalistic stimulus, robustness,<br />

Shab, Shaker


To Alexandra


Acknowledgements<br />

This study was carried out at the Department of Physics, Division of Biophysics <strong>in</strong> the<br />

University of Oulu. I would like to thank my supervisor Professor Matti Weckström, also<br />

the head of the Division of Biophysics and our research group, for his advices and<br />

support dur<strong>in</strong>g these early steps of m<strong>in</strong>e <strong>in</strong> the <strong>in</strong>trigu<strong>in</strong>g world of science. I admire his<br />

positive attitude towards life. I also highly appreciate the way he considers his employees<br />

as friends and co-workers. I enjoyed numerous fruitful discussions we have had.<br />

I would like to thank Dr. Jeremy Niven (Department of Zoology, University of<br />

Cambridge, Cambridge, UK) with whom I have closely collaborated right from the<br />

beg<strong>in</strong>n<strong>in</strong>g. We have had many <strong>in</strong>spir<strong>in</strong>g and productive periods of <strong>in</strong>tensive work<strong>in</strong>g,<br />

which produced the backbone for this study.<br />

I would also like to thank Dr. Mikko Juusola (Physiological Laboratory, University of<br />

Cambridge, Cambridge, UK) and Dr. Roger Hardie (Department of Anatomy, University<br />

of Cambridge, Cambridge, UK). All the experimental data used <strong>in</strong> this study was<br />

collected <strong>in</strong> their laboratories. Our enlighten<strong>in</strong>g discussions and especially their sharp<br />

criticisms revealed possible weaknesses of many hypotheses, thereby significantly<br />

improv<strong>in</strong>g the f<strong>in</strong>al outcome of this study. I am also grateful to Professor Andrew French<br />

(Department of Physiology and Biophysics, Dalhousie University, Halifax, Nova Scotia,<br />

Canada) for his fundamental role <strong>in</strong> the naturalistic stimuli project. I was very impressed<br />

by his certa<strong>in</strong>ty of how th<strong>in</strong>gs should be done and by his efficiency on gett<strong>in</strong>g these<br />

th<strong>in</strong>gs done.<br />

I am also thankful for all the people not mentioned above, especially for the staff <strong>in</strong><br />

the Division of Biophysics, who have helped me <strong>in</strong> many issues dur<strong>in</strong>g this study.<br />

F<strong>in</strong>ally, I would like to thank my family and friends for be<strong>in</strong>g there.<br />

Oulu, 2004 Mikko Vähäsöyr<strong>in</strong>ki


Abbreviations<br />

ATP Adenos<strong>in</strong>e triphosphate<br />

Ca 2+ Calcium ion<br />

Cl - Chloride ion<br />

I-V relationship Current-voltage relationship<br />

K + Potassium ion<br />

k1(u) First-order Volterra kernel<br />

k2(u,v) Second-order Volterra kernel<br />

LIC Light <strong>in</strong>duced conductance<br />

LIC-V relationship Light <strong>in</strong>duced conductance -voltage relationship<br />

MSE Mean square error<br />

Na + Sodium ion<br />

NLN-model Parametric nonl<strong>in</strong>ear-l<strong>in</strong>ear-nonl<strong>in</strong>ear cascade model<br />

LED Light emitt<strong>in</strong>g diode<br />

Q10 Temperature quotient for a 10 °C temperature change<br />

Shab Gene encod<strong>in</strong>g slowly <strong>in</strong>activat<strong>in</strong>g K + potassium current<br />

Shab 1 <strong>Drosophila</strong> mutant; missense mutation <strong>in</strong> Shab gene caus<strong>in</strong>g<br />

75% reduction <strong>in</strong> the Shab current<br />

Shab 2 <strong>Drosophila</strong> mutant; missense mutation <strong>in</strong> Shab gene caus<strong>in</strong>g<br />

71% reduction <strong>in</strong> the Shab current<br />

Shab 3 <strong>Drosophila</strong> mutant; null mutation <strong>in</strong> Shab gene<br />

Shaker Gene encod<strong>in</strong>g A-type K + current<br />

Shaker 14 <strong>Drosophila</strong> mutant; null mutation <strong>in</strong> the Shaker gene<br />

produc<strong>in</strong>g non-functional Shaker K + <strong>channels</strong><br />

Shaker 14 Shab 3 <strong>Drosophila</strong> double mutant; null mutation <strong>in</strong> the Shaker and<br />

Shab genes<br />

SNR Signal-to-noise ratio<br />

WT Wild-type <strong>Drosophila</strong>


List of orig<strong>in</strong>al papers<br />

All ideas and the f<strong>in</strong>al outcomes of these research projects <strong>in</strong> form of publications<br />

resulted from successful collaboration. This makes it impossible to strictly associate<br />

specific ideas or f<strong>in</strong>d<strong>in</strong>gs to certa<strong>in</strong> person. From purely methodological po<strong>in</strong>t of view, I<br />

developed biophysical models based on the experimental data recorded by the Dr. Jeremy<br />

Niven, Dr. Mikko Juusola and Dr. Roger Hardie. This thesis is based on the follow<strong>in</strong>g<br />

articles, which are referred to <strong>in</strong> the text by their Roman numerals I-IV:<br />

I Niven JE * , Vähäsöyr<strong>in</strong>ki M * , Kauranen M, Hardie RC, Juusola M, Weckström M<br />

(2003) The contribution of Shaker K + <strong>channels</strong> to the <strong>in</strong>formation capacity of<br />

<strong>Drosophila</strong> <strong>photoreceptors</strong>. Nature 421, 630-634. *Shared first authorship<br />

II Niven JE, Vähäsöyr<strong>in</strong>ki M, Juusola M (2003) Shaker K + -<strong>channels</strong> are predicted to<br />

reduce the metabolic cost of neural <strong>in</strong>formation <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong>. Proc R<br />

Soc Lond B Biol Sci 270, Suppl 1:S58-61.<br />

III Niven JE, Vähäsöyr<strong>in</strong>ki M, Juusola M, French AS (2004) Interaction between light<strong>in</strong>duced<br />

currents, voltage-<strong>gated</strong> currents and <strong>in</strong>put statistics <strong>in</strong> <strong>Drosophila</strong><br />

<strong>photoreceptors</strong>. J Neurophysiol 91, 2696-2706.<br />

IV Vähäsöyr<strong>in</strong>ki M * ., Niven, J.E * , Hardie, R.C., Weckström, M., Juusola, M. Role of<br />

Shab K + <strong>channels</strong> <strong>in</strong> the <strong>in</strong>formation process<strong>in</strong>g of <strong>Drosophila</strong> <strong>photoreceptors</strong>.<br />

Submitted. *Shared first authorship


Contents<br />

Abstract<br />

Acknowledgements<br />

Abbreviations<br />

List of orig<strong>in</strong>al papers<br />

Contents<br />

1 Introduction ................................................................................................................... 15<br />

2 Review of the literature ................................................................................................. 16<br />

2.1 Basis of neural <strong>in</strong>formation process<strong>in</strong>g ..................................................................16<br />

2.2 Hodgk<strong>in</strong>-Huxley model of the excitable membrane...............................................17<br />

2.3 K + <strong>channels</strong>.............................................................................................................19<br />

2.4 Efficiency of neural cod<strong>in</strong>g ....................................................................................20<br />

2.5 <strong>Drosophila</strong> <strong>photoreceptors</strong> .....................................................................................20<br />

3 Aims of the research ...................................................................................................... 22<br />

4 Material and methods .................................................................................................... 24<br />

4.1 Fly stocks................................................................................................................24<br />

4.2 Optical neutralisation of the cornea........................................................................24<br />

4.3 Electrophysiological experiments...........................................................................24<br />

4.3.1 In vitro patch-clamp record<strong>in</strong>gs from isolated <strong>photoreceptors</strong> ........................24<br />

4.3.2 In vivo <strong>in</strong>tracellular experiments.....................................................................25<br />

4.4 L<strong>in</strong>ear system analysis............................................................................................26<br />

4.5 Nonl<strong>in</strong>ear system analysis ......................................................................................27<br />

4.6 Hodgk<strong>in</strong>-Huxley type photoreceptor models..........................................................29<br />

4.6.1 <strong>Voltage</strong>-dependent properties of Shaker, slow delayed rectifier and novel<br />

non<strong>in</strong>activat<strong>in</strong>g K + <strong>channels</strong>.............................................................................29<br />

4.6.2 Derivation of the other model parameters .......................................................32<br />

4.6.3 MATLAB© implementation of the Hodgk<strong>in</strong>-Huxley type of photoreceptor<br />

model................................................................................................................33<br />

4.6.4 Reconstruction of the light <strong>in</strong>duced conductance ............................................33<br />

4.6.5 Model simulations ...........................................................................................34<br />

4.6.6 Estimation of the metabolic costs....................................................................35


5 Summary of the results <strong>in</strong> the orig<strong>in</strong>al papers ............................................................... 37<br />

5.1 The contribution of Shaker K + <strong>channels</strong> to the <strong>in</strong>formation capacity<br />

of <strong>Drosophila</strong> <strong>photoreceptors</strong> (I) ...........................................................................37<br />

5.2 Metabolic cost of Shaker K + <strong>channels</strong> <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong> (II) ..............38<br />

5.3 Interactions between light-<strong>in</strong>duced currents, voltage-<strong>gated</strong> currents, and <strong>in</strong>put<br />

signal properties <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong> (III) .............................................39<br />

5.4 Role of Shab K + <strong>channels</strong> <strong>in</strong> the <strong>in</strong>formation process<strong>in</strong>g of <strong>Drosophila</strong><br />

<strong>photoreceptors</strong> (IV) ...............................................................................................40<br />

6 Discussion ..................................................................................................................... 43<br />

6.1 Modell<strong>in</strong>g of <strong>Drosophila</strong> <strong>photoreceptors</strong>................................................................43<br />

6.2 K + <strong>channels</strong> <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong> .............................................................44<br />

6.3 The robustness of neural cod<strong>in</strong>g <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong>..............................44<br />

6.4 Naturalistic stimulus...............................................................................................46<br />

6.5 Efficiency of neural cod<strong>in</strong>g ....................................................................................47<br />

6.6 Relevance of the f<strong>in</strong>d<strong>in</strong>gs .......................................................................................48<br />

7 Conclusions ...................................................................................................................49<br />

References<br />

Appendix


1 Introduction<br />

In nature the success of an <strong>in</strong>dividual depends upon accurately determ<strong>in</strong><strong>in</strong>g and execut<strong>in</strong>g<br />

the appropriate behaviour based on the sensory <strong>in</strong>formation it gathers from its<br />

surround<strong>in</strong>gs. The most important cellular components of neural <strong>in</strong>formation process<strong>in</strong>g<br />

are ion <strong>channels</strong> found <strong>in</strong> the neural cell membrane. To understand why particular ion<br />

<strong>channels</strong> are expressed <strong>in</strong> specific neurons requires detailed knowledge of their<br />

contribution to <strong>in</strong>formation process<strong>in</strong>g. Insect <strong>photoreceptors</strong> have provided a model<br />

system for exam<strong>in</strong><strong>in</strong>g specific molecular mechanisms <strong>in</strong>volved <strong>in</strong> <strong>in</strong>formation process<strong>in</strong>g<br />

with graded voltage signals, <strong>in</strong>clud<strong>in</strong>g signal transduction (the phototransduction<br />

cascade) (Hardie & Raghu 2001, Hardie 2003) and membrane filter<strong>in</strong>g (the photo<strong>in</strong>sensitive<br />

membrane) (Weckström & Laughl<strong>in</strong> 1995). In <strong>Drosophila</strong> these mechanisms<br />

can be studied <strong>in</strong> relative isolation by comb<strong>in</strong><strong>in</strong>g electrophysiological experiments with<br />

genetic methods and mathematical modell<strong>in</strong>g. Focus of this study is on the widely<br />

expressed voltage-activated K + <strong>channels</strong> that are known to confer specific electrical<br />

properties upon the membranes enabl<strong>in</strong>g neurons to process biologically relevant<br />

<strong>in</strong>formation (Rudy 1988, Weckström & Laughl<strong>in</strong> 1995, Coetzee et al. 1999, Hille 2001,<br />

Johnston et al. 2003).<br />

<strong>Drosophila</strong> <strong>photoreceptors</strong> express voltage-activated K + <strong>channels</strong> <strong>in</strong> their photo<strong>in</strong>sensitive<br />

membrane: Shaker and slow delayed rectifier <strong>channels</strong> be<strong>in</strong>g characterised as<br />

the predom<strong>in</strong>ant ones (Hardie 1991a, Hardie et al. 1991, Hevers & Hardie 1995). What is<br />

the contribution of these voltage-activated <strong>channels</strong> to the <strong>in</strong>formation process<strong>in</strong>g <strong>in</strong><br />

<strong>Drosophila</strong> <strong>photoreceptors</strong>? How do the currents generated by these <strong>channels</strong> <strong>in</strong>teract<br />

with each other and with the currents generated by the light transduction mach<strong>in</strong>ery?<br />

How do these <strong>channels</strong> affect the energy efficiency of neural code? How robust are<br />

<strong>Drosophila</strong> <strong>photoreceptors</strong> to imposed changes <strong>in</strong> their ion channel composition? These<br />

questions allow a systematic approach for estimat<strong>in</strong>g the contribution of voltage-<strong>gated</strong> K +<br />

<strong>channels</strong> to the performance of <strong>Drosophila</strong> <strong>photoreceptors</strong>. Results of this study may<br />

provide new <strong>in</strong>sight to graded voltage signal process<strong>in</strong>g by K + <strong>channels</strong> (e.g., <strong>in</strong> sensory<br />

neurons and <strong>in</strong> dendrites) and to possible mechanisms of robustness <strong>in</strong> neural cod<strong>in</strong>g.


2 Review of the literature<br />

2.1 Basis of neural <strong>in</strong>formation process<strong>in</strong>g<br />

Information is carried with<strong>in</strong> and between neurons by complex electrical and chemical<br />

signall<strong>in</strong>g pathways. Every neuron has a separation of charges across its cell membrane<br />

consist<strong>in</strong>g of a th<strong>in</strong> cloud of positive and negative ions spread over the <strong>in</strong>ner and outer<br />

surfaces of the lipid bilayer membrane. The charge separation gives rise to a difference <strong>in</strong><br />

electrical potential across the membrane called the membrane potential, Vm:<br />

Vm= V<strong>in</strong>−Vout The membrane potential of a cell at rest is called the rest<strong>in</strong>g potential. A reduction of<br />

charge separation, lead<strong>in</strong>g to less negative membrane potential, is called depolarisation.<br />

An <strong>in</strong>crease <strong>in</strong> charge separation, lead<strong>in</strong>g to a more negative membrane potential, is<br />

called hyperpolarisation.<br />

Electrical signals – receptor potentials, synaptic potentials, and action potentials – are<br />

produced by the movements of ions <strong>in</strong>to and out of the cells through ion <strong>channels</strong><br />

embedded <strong>in</strong> the cell membrane. The membrane’s overall selectivity for <strong>in</strong>dividual ion<br />

species, ma<strong>in</strong>ly Na + , K + , Cl - and Ca 2+ , is determ<strong>in</strong>ed by the relative proportions of the<br />

various types of ion <strong>channels</strong> <strong>in</strong> the cell that are open. The ions are subject to two forces<br />

driv<strong>in</strong>g them across the membrane: (1) a chemical driv<strong>in</strong>g force that depends on the<br />

concentration gradient across the membrane and (2) an electrical driv<strong>in</strong>g force that<br />

depends on the electrical potential differences across the membrane. The equilibrium<br />

potential, Ei, for any ion can be calculated from the Nernst equation:<br />

RT c<strong>in</strong><br />

Ei=− ln<br />

(2)<br />

zF cout<br />

where R is the gas constant, T is the temperature, z is the valence of the ion, c<strong>in</strong> and cout<br />

are the <strong>in</strong>tra- and extracellular concentrations for the ions.<br />

(1)


17<br />

The equivalent electrical circuit of a neuron provides an <strong>in</strong>tuitive understand<strong>in</strong>g as<br />

well as a quantitative description of how the current flows due to the movements of ions<br />

generate voltage signals <strong>in</strong> neurons. The circuit model of a patch of the squid axon<br />

(Hodgk<strong>in</strong> & Huxley 1952) is shown <strong>in</strong> figure 1.<br />

Fig. 1. Equivalent electrical circuit of the squid axon. C m is the membrane capacitance, I i are<br />

the currents and gi are the conductances.<br />

ss<br />

The steady-state potential, V m , for the equivalent circuit <strong>in</strong> Fig. 1 can be calculated with<br />

the equation:<br />

V<br />

g E + g E + g E<br />

=<br />

ss Na Na K K L L<br />

m<br />

gNa + g<strong>K+</strong> gL<br />

where gi are the conductances and Ei are the Nernst potentials for sodium, potassium and<br />

the leak, respectively. In the limit, if the conductance for one ion is much greater than that<br />

for the other ions, Vm will approach the value of the Nernst potential of the correspond<strong>in</strong>g<br />

ion. Dur<strong>in</strong>g signall<strong>in</strong>g the membrane voltage deviates from the steady-state and changes<br />

constantly. The dynamics of a passive membrane can be def<strong>in</strong>ed by the time constant, τ:<br />

τ= RC <strong>in</strong> m<br />

where R<strong>in</strong> is the <strong>in</strong>put resistance and Cm is the capacitance of the membrane.<br />

2.2 Hodgk<strong>in</strong>-Huxley model of the excitable membrane<br />

In 1952 A.L. Hodgk<strong>in</strong> and A.F. Huxley published a series of papers <strong>in</strong> which they<br />

<strong>in</strong>vesti<strong>gated</strong> the flow of current through the membrane of the giant nerve fibre of the<br />

(3)<br />

(4)


18<br />

squid (Hodgk<strong>in</strong> & Huxley 1952). Currents were recorded by us<strong>in</strong>g the voltage clamp<br />

technique, where the membrane potential is stepped accord<strong>in</strong>g to the experimental<br />

protocol and kept constant with the feedback amplifier. Based on these experiments they<br />

developed a biophysical model of the excitable membrane. Their electrical model<br />

conta<strong>in</strong>ed not only the passive membrane with fixed capacitance and resistance, but also<br />

the contribution of active sodium and potassium ion <strong>channels</strong>, which change their<br />

properties as a function of voltage.<br />

In the Hodgk<strong>in</strong>-Huxley model (Fig. 1) the total current, I, can be written as a sum of<br />

capacitive, Icap, and resistive currents, Ii. The capacitive current follows the equation:<br />

dVm<br />

Icap = Cm dt<br />

(5)<br />

where Cm is the membrane capacitance and Vm is the membrane potential. The resistive<br />

currents, Ii, are described with the equation:<br />

( )<br />

I = g V −E<br />

i i m i<br />

where gi is the conductance for each ion. The equation for the membrane voltage, Vm, can<br />

be derived by comb<strong>in</strong><strong>in</strong>g the current equations (5) and (6):<br />

( − ) + ( − ) + ( − )<br />

dV g V E g V E g V E<br />

m Im<br />

= −<br />

dt C C<br />

Na m Na K m K L m L<br />

m m<br />

where Im is the current <strong>in</strong>jected by the experimenter.<br />

Hodgk<strong>in</strong> and Huxley observed that dur<strong>in</strong>g voltage clamp the sodium conductance first<br />

rises with short delay and then falls aga<strong>in</strong> to a low value, processes they called activation<br />

and <strong>in</strong>activation, respectively. Based on these f<strong>in</strong>d<strong>in</strong>gs they developed an empirical<br />

k<strong>in</strong>etic model that predicted correctly the major features of the experimental<br />

observations. The k<strong>in</strong>etic behaviour of the open<strong>in</strong>g and clos<strong>in</strong>g of the ion <strong>channels</strong> are<br />

controlled by a vary<strong>in</strong>g number of <strong>in</strong>dependent membrane-bound particles. The state of<br />

the gat<strong>in</strong>g particle, γk, is assumed to undergo first-order transitions between permissive<br />

and non-permissive forms:<br />

α<br />

γ<br />

1−<br />

γ ←⎯⎯ ⎯⎯→ γ<br />

k<br />

β<br />

k<br />

γ<br />

where αk and βk are the voltage dependent rate constants. The values for the gat<strong>in</strong>g<br />

particles, describ<strong>in</strong>g the probability of the permissive state, γk, can be obta<strong>in</strong>ed from the<br />

differential equation:<br />

d<br />

dt<br />

γ k =αγ( 1−γ<br />

k) − βγ<br />

γ k<br />

Based on the theory of gat<strong>in</strong>g particles, the conductance for each ion channel, gi, can be<br />

calculated with the equation:<br />

(6)<br />

(7)<br />

(8)<br />

(9)


19<br />

(10)<br />

( )<br />

where gimax is the maximum conductance and n is the number of gat<strong>in</strong>g particles, which,<br />

while <strong>in</strong>creas<strong>in</strong>g from 1 to 2, 3, 4…, changes the ris<strong>in</strong>g phase of the conductance from<br />

exponential to sigmoidal. S<strong>in</strong>ce the orig<strong>in</strong>al publications, numerous models have been<br />

developed follow<strong>in</strong>g the Hodgk<strong>in</strong>-Huxley formalism for a variety of systems (Hille<br />

2001). Hodgk<strong>in</strong>-Huxley type of models still form the reference for study<strong>in</strong>g the role of<br />

ion <strong>channels</strong> <strong>in</strong> electrical signall<strong>in</strong>g (Meunier & Segev 2002).<br />

n<br />

gi = gimax∏⎡⎣γk V m,t<br />

⎤⎦<br />

k<br />

2.3 K + <strong>channels</strong><br />

The history of the K + <strong>channels</strong> starts from the year 1902 when Julius Bernste<strong>in</strong> postulated<br />

a selective potassium permeability <strong>in</strong> excitable cell membranes (Bernste<strong>in</strong> 1902).<br />

Presently K + <strong>channels</strong> are known to be expressed <strong>in</strong> all animal cells, <strong>in</strong>volv<strong>in</strong>g <strong>in</strong><br />

numerous functions (Hille 2001). Indeed, K + <strong>channels</strong> are generally considered to<br />

represent the most diverse of the ion channel families (Rudy 1988), each type of neuron<br />

express<strong>in</strong>g its own blend of <strong>channels</strong> to suit its special purposes. The best known example<br />

is the axonal membrane express<strong>in</strong>g predom<strong>in</strong>antly one major class of K + <strong>channels</strong> – the<br />

delayed rectifier type – that repolarises the membrane after the action potential (Hodgk<strong>in</strong><br />

& Huxley 1952).<br />

An A-type channel from <strong>Drosophila</strong> was the first K + channel to be cloned (Tempel et<br />

al. 1987, Kamb et al. 1988, Pongs et al. 1988). The identification of this channel started<br />

with mutant flies called Shaker, named after their tendency to shake legs while under<br />

ether anesthesia. This one gene locus was found to code for a variety of K + <strong>channels</strong> by<br />

mechanisms of alternative splic<strong>in</strong>g (Iverson et al. 1988, Pongs et al. 1988, Schwarz et al.<br />

1988, Timpe et al. 1988, Hardie et al. 1991) and heteromultimeric comb<strong>in</strong>ations (Isacoff<br />

et al. 1990, McCormack et al. 1990). Later three other genes were identified <strong>in</strong><br />

<strong>Drosophila</strong>, complement<strong>in</strong>g the K + channel family, named Shaker, Shal, Shab and Shaw,<br />

cod<strong>in</strong>g for K + <strong>channels</strong> rang<strong>in</strong>g from a rapidly <strong>in</strong>activat<strong>in</strong>g A current to a slow delayed<br />

rectifier current (Butler et al. 1989, Wei et al. 1990).<br />

The equilibrium potential for the K + <strong>channels</strong> is typically slightly below the rest<strong>in</strong>g<br />

potential of a neuron. Therefore, the open<strong>in</strong>g of the K + <strong>channels</strong> draws the membrane<br />

potential back to the direction of the rest<strong>in</strong>g potential, produc<strong>in</strong>g a stabilis<strong>in</strong>g effect. In<br />

spik<strong>in</strong>g neurons, the functions of K + <strong>channels</strong> <strong>in</strong>clude sett<strong>in</strong>g the rest<strong>in</strong>g potential,<br />

keep<strong>in</strong>g fast action potentials short, term<strong>in</strong>at<strong>in</strong>g periods of <strong>in</strong>tense activity, tim<strong>in</strong>g the<br />

<strong>in</strong>ter-spike <strong>in</strong>tervals dur<strong>in</strong>g repetitive fir<strong>in</strong>g, and generally lower<strong>in</strong>g the effectiveness of<br />

excitatory <strong>in</strong>puts on a neuron when they are open (Hille 2001). In graded potential<br />

neurons the potassium <strong>channels</strong>, <strong>in</strong> addition to their stabilis<strong>in</strong>g role, may tune the cell<br />

membrane of sensory neurons <strong>in</strong>to an optimal filter for the transduced signal (Laughl<strong>in</strong> &<br />

Weckstrom 1993, Weckström & Laughl<strong>in</strong> 1995).


20<br />

2.4 Efficiency of neural cod<strong>in</strong>g<br />

Information theory quantifies our <strong>in</strong>tuitive notions of ga<strong>in</strong><strong>in</strong>g <strong>in</strong>formation and puts it on<br />

the absolute scale (Rieke 1997, Dayan & Abbott 2001). The <strong>in</strong>formation capacity of a<br />

system gives an upper limit for the number of states a signall<strong>in</strong>g system can transmit <strong>in</strong> a<br />

given time w<strong>in</strong>dow. Energy as well as <strong>in</strong>formation capacity is an important consideration<br />

<strong>in</strong> understand<strong>in</strong>g neural process<strong>in</strong>g. Potentially, energetic costs could limit absolute<br />

numbers of neurons and synaptic connections <strong>in</strong> the neural systems and <strong>in</strong>fluence the<br />

expression and distribution of ion <strong>channels</strong> (Laughl<strong>in</strong> et al. 1998).<br />

Usage of energy can be l<strong>in</strong>ked to two fundamental measures of signal quality: signalto-noise<br />

ratio (SNR) and bandwidth (a measure of speed of response) (Weckström &<br />

Laughl<strong>in</strong> 1995). Improv<strong>in</strong>g the SNR <strong>in</strong>creases the energy demands, because reliability<br />

<strong>in</strong>creases as the square root of the number of uncorrelated stochastic events that generate<br />

signals. Each stochastic signall<strong>in</strong>g event, such as open<strong>in</strong>g an ion channel or releas<strong>in</strong>g a<br />

synaptic vesicle, requires extra energy. Rais<strong>in</strong>g the bandwidth means that the membrane<br />

time constant must be reduced by <strong>in</strong>creas<strong>in</strong>g conductance. Increased conductance means<br />

larger ion fluxes across the membrane that has to be compensated by ATP driven ion<br />

transport. Evidence is accumulat<strong>in</strong>g that neurons economize by restrict<strong>in</strong>g the SNR and<br />

bandwidth to the m<strong>in</strong>imum required to get the job “done just right-enough” (Laughl<strong>in</strong><br />

2001).<br />

2.5 <strong>Drosophila</strong> <strong>photoreceptors</strong><br />

The <strong>Drosophila</strong> eye is composed of ca. 800 ommatidia, each of them consist<strong>in</strong>g of<br />

sophisticated optics transmitt<strong>in</strong>g photons to sensory neurons (Stavenga 2003b, Stavenga<br />

2003a, Stavenga 2004). Photoreceptors (Fig. 2) are the light detectors of visual systems<br />

convert<strong>in</strong>g the energy of <strong>in</strong>com<strong>in</strong>g photons <strong>in</strong>to electrical signals by a process of<br />

phototransduction (Hardie 2001). Discrete electrical events (bumps) can be recorded<br />

under very dim illum<strong>in</strong>ation correspond<strong>in</strong>g to s<strong>in</strong>gle photon absorptions <strong>in</strong> the photosensitive<br />

membrane (Wu & Pak 1975). The <strong>photoreceptors</strong> respond to light by graded<br />

changes <strong>in</strong> their membrane potential consist<strong>in</strong>g of numerous fused bumps.<br />

Although the photoreceptor voltage responses are <strong>in</strong>itiated by the currents generated <strong>in</strong><br />

the photo-sensitive membrane, the f<strong>in</strong>al shape of these responses is determ<strong>in</strong>ed by the<br />

properties of the photo-<strong>in</strong>sensitive membrane conta<strong>in</strong><strong>in</strong>g voltage-<strong>gated</strong> K + <strong>channels</strong><br />

(Shaker and slow delayed rectifier <strong>channels</strong> be<strong>in</strong>g characterised as the predom<strong>in</strong>ant ones,<br />

(Hardie 1991a, Hardie et al. 1991, Hevers & Hardie 1995). In <strong>Drosophila</strong> the<br />

contribution of each type of K + channel to the performance can be studied <strong>in</strong> relative<br />

isolation by comb<strong>in</strong><strong>in</strong>g electrophysiological experiments with genetic approaches and<br />

modell<strong>in</strong>g. This makes it a good model system to <strong>in</strong>vestigate the role of voltage-<strong>gated</strong> K +<br />

<strong>channels</strong> <strong>in</strong> neural process<strong>in</strong>g.


21<br />

Fig. 2. Schematic <strong>in</strong>troduction to the <strong>Drosophila</strong> <strong>photoreceptors</strong>. Scale bar <strong>in</strong> the electron<br />

microscope picture <strong>in</strong> the right lower corner is 1 µm. Figure is modified from pictures<br />

adapted from the <strong>Drosophila</strong> FlyBase (http://fly.ebi.ac.uk:7081/) and from (Hardie & Raghu<br />

2001).


3 Aims of the research<br />

The follow<strong>in</strong>g questions were <strong>in</strong>vesti<strong>gated</strong> to systematically estimate the contribution of<br />

voltage-<strong>gated</strong> K + <strong>channels</strong> to the performance of <strong>Drosophila</strong> <strong>photoreceptors</strong>:<br />

1. The contribution of the Shaker <strong>channels</strong> and its functional homologues (Rudy 1988,<br />

Coetzee et al. 1999) to neuronal function rema<strong>in</strong>s unclear, although they are thought to<br />

attenuate the amplitude of graded potentials and back-propa<strong>gated</strong> action potentials <strong>in</strong><br />

dendrites (Laurent 1990, Hoffman & Johnston 1998, Magee et al. 1998), to <strong>in</strong>fluence<br />

the fir<strong>in</strong>g frequency of spik<strong>in</strong>g neurones (Connor & Stevens 1971) and to determ<strong>in</strong>e<br />

the reliability of spike propagation (Debanne et al. 1997). What is the contribution of<br />

Shaker K + <strong>channels</strong> to the <strong>in</strong>formation process<strong>in</strong>g <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong>?<br />

2. Sensory systems are thought to be tuned to the statistical properties of their natural<br />

stimuli through evolutionary and/or developmental processes (Weckström & Laughl<strong>in</strong><br />

1995, Katz & Shatz 1996). Therefore, understand<strong>in</strong>g how specific components<br />

function with<strong>in</strong> these sensory systems should benefit from us<strong>in</strong>g stimuli with<br />

appropriate dynamic properties. How do the voltage-<strong>gated</strong> currents <strong>in</strong>teract with each<br />

other and with the light-<strong>in</strong>duced currents <strong>in</strong> natural stimulus regimes <strong>in</strong> <strong>Drosophila</strong><br />

<strong>photoreceptors</strong>?<br />

3. Despite their importance <strong>in</strong> regulat<strong>in</strong>g neural activity, many genetic deletions of<br />

specific ion <strong>channels</strong> have caused relatively little changes <strong>in</strong> neuron performance<br />

(Wickman et al. 1998, Brickley et al. 2001), even when pharmacological studies<br />

suggest there should be severe phenotypes. This mismatch suggests that altered<br />

channel expression may be compensated for by changes <strong>in</strong> other ion <strong>channels</strong> (Liu et<br />

al. 1998, Namkung et al. 1998, MacLean et al. 2003). Do regulatory changes <strong>in</strong> the<br />

electrical properties, potentially restor<strong>in</strong>g the photoreceptor’s performance, follow<br />

from deletion of the Shaker <strong>channels</strong>?<br />

4. The efficient expenditure of a limited energy budget is a major factor <strong>in</strong>fluenc<strong>in</strong>g all<br />

aspects of animal design. Therefore, energy is thought to have been an important<br />

constra<strong>in</strong>t upon the evolution of nervous systems (Laughl<strong>in</strong> 2001). S<strong>in</strong>ce neural<br />

cod<strong>in</strong>g is metabolically expensive (Kety 1957, Ames 1997, Laughl<strong>in</strong> et al. 1998), the<br />

metabolic cost of a particular conductance may be an important consideration. How do<br />

the Shaker K + <strong>channels</strong> affect the energy efficiency of neural <strong>in</strong>formation <strong>in</strong><br />

<strong>Drosophila</strong> <strong>photoreceptors</strong>?


23<br />

5. <strong>Voltage</strong>-<strong>gated</strong> K + <strong>channels</strong> have different biophysical properties and, therefore, their<br />

role <strong>in</strong> shap<strong>in</strong>g the voltage signals is likely to be different. What is the role of slow<br />

delayed rectifier K + <strong>channels</strong> <strong>in</strong> neural process<strong>in</strong>g of <strong>Drosophila</strong> <strong>photoreceptors</strong>? Are<br />

potential changes <strong>in</strong> the electrical properties due to reduction <strong>in</strong> the slow delayed<br />

rectifier conductance similar to that of the Shaker mutant <strong>photoreceptors</strong> or do they<br />

possibly provide robustness for different aspects of function?


4 Material and methods<br />

4.1 Fly stocks<br />

Flies (<strong>Drosophila</strong> melanogaster) were raised on standard medium at 19°C <strong>in</strong> darkness.<br />

The wild-type stra<strong>in</strong> was red-eyed <strong>Drosophila</strong> melanogaster Oregon Red. The Shaker<br />

channel mutation was Shaker 14 (a missense mutation <strong>in</strong> the core region result<strong>in</strong>g <strong>in</strong> nonfunctional<br />

Shaker <strong>channels</strong>) (Kaplan & Trout 1961, Salkoff & Wyman 1981). Three<br />

alleles of the Shab gene (Shab 1 , Shab 2 and Shab 3 ) were used either <strong>in</strong> ebony or wild-type<br />

background. Shab 1 and Shab 2 are missense mutations whereas Shab 3 is a null allele<br />

(Hegde et al. 1999). Shaker 14 Shab 3 double mutants lacked functional Shaker and Shab<br />

<strong>channels</strong>. All the mutants were expressed <strong>in</strong> red-eyed flies.<br />

4.2 Optical neutralisation of the cornea<br />

For observ<strong>in</strong>g rhabdomeres <strong>in</strong> the <strong>in</strong>tact eye, heads were mounted on microscope slides<br />

us<strong>in</strong>g clear nail varnish and observed under a 63× oil immersion objective us<strong>in</strong>g<br />

antidromic illum<strong>in</strong>ation for photography (Francesch<strong>in</strong>i & Kirschfeld 1971).<br />

4.3 Electrophysiological experiments<br />

4.3.1 In vitro patch-clamp record<strong>in</strong>gs from isolated <strong>photoreceptors</strong><br />

Whole-cell record<strong>in</strong>gs from dissociated <strong>Drosophila</strong> <strong>photoreceptors</strong> were performed as<br />

previously described (Hardie 1991a, Hardie 1991b, Hardie et al. 1991, Hardie et al.<br />

2001). Dissociated ommatidia from recently eclosed adult flies were prepared and<br />

transferred to the bottom of a record<strong>in</strong>g chamber on an <strong>in</strong>verted Nikon Diaphot


25<br />

microscope. The bath was composed of (<strong>in</strong> mM): 120 NaCl, 5 KCl, 10 TES, 4 MgCl2, 1.5<br />

CaCl2, 25 prol<strong>in</strong>e and 5 alan<strong>in</strong>e. The standard <strong>in</strong>tracellular solution was (<strong>in</strong> mM): 140 K +<br />

gluconate, 10 TES, 4 Mg ATP, 2 MgCl2, 1 NAD and 0.4 Na + -GTP. The pH of all<br />

solutions was 7.15. Whole-cell voltage clamp record<strong>in</strong>gs were made us<strong>in</strong>g electrodes<br />

with a resistance of ~10-15MΩ. The series resistance values were generally below 25MΩ<br />

and were rout<strong>in</strong>ely compensated to >80%. Data were collected and analyzed us<strong>in</strong>g an<br />

Axopatch 1-D amplifier and pCLAMP 8 or 9 software (Axon Instruments, Foster City,<br />

CA, USA).<br />

Cells were stimulated via a green LED (560 nm), with maximum effective <strong>in</strong>tensity of<br />

~2x10 5 photons/s per photoreceptor. Relative <strong>in</strong>tensities were calibrated us<strong>in</strong>g a l<strong>in</strong>ear<br />

photodiode and converted to absolute <strong>in</strong>tensities <strong>in</strong> terms of effectively absorbed photons<br />

by count<strong>in</strong>g quantum bumps at low <strong>in</strong>tensities (Henderson et al. 2000). Bumps were<br />

elicited either by cont<strong>in</strong>uous dim illum<strong>in</strong>ation or by repeated brief flashes conta<strong>in</strong><strong>in</strong>g on<br />

average less than one effective photon and were detected and analysed off-l<strong>in</strong>e us<strong>in</strong>g<br />

M<strong>in</strong>ianalysis software (Synaptosoft) (Henderson et al. 2000). All the patch-clamp data<br />

used <strong>in</strong> this study was provided by Dr. Roger Hardie (Department of Anatomy, University<br />

of Cambridge, Cambridge, UK).<br />

4.3.2 In vivo <strong>in</strong>tracellular experiments<br />

Intracellular record<strong>in</strong>gs, stimulation and data acquisition have been described previously<br />

<strong>in</strong> detail (Juusola & Hardie 2001a). Flies were fixed us<strong>in</strong>g wax <strong>in</strong> a custom built ceramic<br />

holder and a small w<strong>in</strong>dow was cut <strong>in</strong>to the surface of the compound eye. Record<strong>in</strong>gs<br />

were made us<strong>in</strong>g quartz microelectrodes (resistances between 150 and 220 MΩ, filled<br />

with 3 M KCl) placed to the ret<strong>in</strong>a through a small hole cut <strong>in</strong> the surface of the eye. As<br />

<strong>in</strong>different electrode a second blunt microelectrode filled with fly R<strong>in</strong>ger was also placed<br />

<strong>in</strong> the flies' head close to the eye. All record<strong>in</strong>gs were made us<strong>in</strong>g a switch clamp<br />

amplifier (SEC 10L, npi Electronic, Germany) <strong>in</strong> current-clamp mode. The temperature<br />

of the flies was ma<strong>in</strong>ta<strong>in</strong>ed at 25 °C throughout the experiments to with<strong>in</strong> 1 °C accuracy<br />

us<strong>in</strong>g a thermocouple feed<strong>in</strong>g back to a Peltier device.<br />

Impaled <strong>photoreceptors</strong> were stimulated by a high <strong>in</strong>tensity green LED with peak<br />

wavelength of 525 nm (3× 10 6 photons/s; Marl Optosource, UK). Data acquisition,<br />

stimulus generation and signal analysis were performed by a purpose built MATLAB<br />

<strong>in</strong>terface (Biosyst, © Mikko Juusola). For the natural stimulus experiments, the light<br />

<strong>in</strong>tensity was derived from a published time series obta<strong>in</strong>ed from a light detector mov<strong>in</strong>g<br />

through a natural environment (van Hateren 1997). Photoreceptors were considered for<br />

analysis only if their rest<strong>in</strong>g membrane potential was less than -55mV and they had at<br />

least a 45 mV saturat<strong>in</strong>g impulse response <strong>in</strong> dark-adapted conditions. All the<br />

<strong>in</strong>tracellular experiments were carried out by Dr. Jeremy Niven and Dr. Mikko Juusola <strong>in</strong><br />

the Physiological Laboratory at University of Cambridge, Cambridge, UK.


26<br />

4.4 L<strong>in</strong>ear system analysis<br />

The goal of system analysis is to establish a cause-effect relationship that would allow the<br />

prediction of the system response to any arbitrary stimulus. For l<strong>in</strong>ear systems the<br />

property of superposition allows measurements from a restricted range of stimuli to<br />

predict the response to any stimulus. Traditional l<strong>in</strong>ear system identification relies on<br />

analys<strong>in</strong>g responses to the impulse, step and Gaussian white noise stimuli (to determ<strong>in</strong>e<br />

the frequency response function of a system). Gaussian white noise is a stationary, zeroaverage<br />

random process, which has the fundamental property that any two samples of it<br />

are statistically <strong>in</strong>dependent. Additionally, the power spectrum of ideal Gaussian white<br />

noise is constant over all frequencies. Due to experimental limitations (limited signal<br />

lengths) a band limited Gaussian white noise is used with a frequency band exceed<strong>in</strong>g<br />

that of the <strong>in</strong>vesti<strong>gated</strong> system (Marmarelis & Marmarelis 1978).<br />

Repeated presentations (10-30 times) of an identical pseudorandom Gaussian white<br />

noise modulated current, i(t), or light contrast, c(t), evoke variable voltage responses due<br />

both to record<strong>in</strong>g noise and the stochastic nature of the underly<strong>in</strong>g biological processes.<br />

Averag<strong>in</strong>g the responses gives the noise-free light contrast or current-evoked<br />

photoreceptor voltage signal, sv(t). Subtraction of this signal from the <strong>in</strong>dividual<br />

responses, rv(t)i, gives the noise component, nv(t)i, of each <strong>in</strong>dividual response period:<br />

() = () − ()<br />

n t r t s t<br />

V i V i V<br />

Errors due to residual noise <strong>in</strong> sV(t) are small and proportional to (noise power)/ n<br />

(Kouvala<strong>in</strong>en et al. 1994).<br />

To calculate the signal-to-noise ratio <strong>in</strong> the frequency doma<strong>in</strong>, SNRv(f), signal and<br />

noise were segmented <strong>in</strong>to 50% overlapp<strong>in</strong>g stretches and w<strong>in</strong>dowed with a Blackman-<br />

Harris fourth term w<strong>in</strong>dow before their correspond<strong>in</strong>g spectra, Sv(f) and Nv(f), were<br />

calculated with an FFT algorithm. SNRv(f) is given by:<br />

( )<br />

2<br />

Sf<br />

SNR (12)<br />

V ( f ) =<br />

2<br />

N( f)<br />

From the SNRv(f) an estimate of the <strong>in</strong>formation capacity (bits/s), C, can be calculated<br />

us<strong>in</strong>g Shannon's formula (Shannon 1948):<br />

∞<br />

∫<br />

0<br />

( 2 ⎡ ( ) ⎤)<br />

C= log ⎣SNR f + 1⎦ df<br />

The photoreceptor frequency response function, G(f), for the light contrast stimulus, c(t),<br />

as well as membrane impedance, Z(f), for the current stimulus, i(t), can be calculated<br />

with:<br />

( ) ( )<br />

( ) ( )<br />

( ) ( )<br />

( ) ( )<br />

* *<br />

SV f ⋅C f SV f ⋅I<br />

f<br />

G( f) = and Z( f)<br />

=<br />

* *<br />

Cf ⋅CfIf⋅If (11)<br />

(13)<br />

(14)


27<br />

where asterisk stands for the complex conjugate of the correspond<strong>in</strong>g Fourier transform.<br />

How do the voltage-<strong>gated</strong> potassium <strong>channels</strong> affect the membrane impedance?<br />

Activation of a channel means that more ion <strong>channels</strong> are opened, reduc<strong>in</strong>g the<br />

impedance <strong>in</strong> the frequency range def<strong>in</strong>ed by the activation time constant, τ. The 3dB<br />

corner frequency, f3dB, of this effect can be calculated with equation:<br />

(15)<br />

On the other hand, <strong>in</strong>activation of the ion <strong>channels</strong> <strong>in</strong>creases the impedance. Time<br />

constants for the <strong>in</strong>activation are usually much longer than those of the activation limit<strong>in</strong>g<br />

this effect to the low frequencies. If a significant proportion of the ion <strong>channels</strong> is open at<br />

the steady-state, the <strong>in</strong>activation may reduce the conductance below the steady-state<br />

value. This causes the membrane impedance to exceed that of the steady-state at the<br />

frequencies determ<strong>in</strong>ed by the <strong>in</strong>activation time constant. A schematic presentation of the<br />

relationships between the membrane impedance and the dynamics of voltage-<strong>gated</strong> K +<br />

1<br />

f3dB<br />

=<br />

2πτ<br />

<strong>channels</strong> is shown <strong>in</strong> Fig 3.<br />

Fig. 3. The effect of voltage-<strong>gated</strong> K + <strong>channels</strong> on the membrane impedance<br />

4.5 Nonl<strong>in</strong>ear system analysis<br />

For nonl<strong>in</strong>ear systems the pr<strong>in</strong>ciple of superposition does not hold. Responses of neural<br />

systems or components to one set of stimuli may not predict their responses to other<br />

<strong>in</strong>puts (Rieke et al. 1995, van Hateren 1997, Lewen et al. 2001, Vickers et al. 2001,<br />

R<strong>in</strong>berg & Davidowitz 2002, Burton & Laughl<strong>in</strong> 2003, Chacron et al. 2003, Juusola & de<br />

Polavieja 2003, Niven & Burrows 2003). For example, responses to white noise<br />

stimulation of lateral geniculate nucleus neurons did not predict responses to naturalistic<br />

stimuli (Dan et al. 1996). Kernels are a way to describe the pattern of the nonl<strong>in</strong>ear


28<br />

<strong>in</strong>teraction among the past events with regard to the effect that this <strong>in</strong>teraction has upon<br />

the system response.<br />

Nonl<strong>in</strong>ear system identification <strong>in</strong> this study was based on estimat<strong>in</strong>g the kernels of a<br />

Volterra series, K0, K1(u), K2(u,v), ... where u, v, ... are time lags (French & Marmarelis<br />

1999) with light <strong>in</strong>tensity as the <strong>in</strong>put, x(t), as a function of time, t, and receptor potential<br />

as the output, y(t).<br />

∞ ∞ ∞<br />

y( t) = K0 + ∫K1( u) x ( t − u) du + ∫∫K2(<br />

u, v) x ( t −u) x ( t −v)<br />

dudv<br />

0 0 0<br />

Several methods have been developed for kernel estimation. Earlier methods relied on<br />

stimulat<strong>in</strong>g the unknown system with Gaussian white noise, but more recent methods<br />

avoid this requirement (French & Marmarelis 1999). A completely general approach<br />

based on the parallel cascade method (Korenberg 1991, Juusola et al. 2003) was used<br />

hav<strong>in</strong>g the l<strong>in</strong>ear filters of the cascades formed from Gaussian distributed random<br />

numbers (French et al. 2001, Juusola et al. 2003). This method makes no <strong>in</strong>itial<br />

assumptions about the forms of the kernels or the nature of the <strong>in</strong>put or output signals and<br />

it can be applied to systems conta<strong>in</strong><strong>in</strong>g relatively high order of nonl<strong>in</strong>earities. However, it<br />

is not necessary to construct all of the higher order kernels.<br />

After kernel estimation, percentage mean square error (MSE) values (French &<br />

Marmarelis 1999) were calculated for the zero and first-order kernels alone and for the<br />

comb<strong>in</strong>ed, zero-, first-, second and third-order kernels from:<br />

2<br />

(y(t) − y s(t))<br />

MSE =<br />

(17)<br />

2 2<br />

y(t) − (y(t))<br />

where ys(t) is the Volterra series output and the bars <strong>in</strong>dicate time averages. The kernel<br />

estimates were used to predict the output of the nonl<strong>in</strong>ear system to the <strong>in</strong>put signal. All<br />

predictions were based on recorded data that had not been used for system identification.<br />

Based on the kernel analysis, a NLN (Nonl<strong>in</strong>ear static-L<strong>in</strong>ear dynamic-Nonl<strong>in</strong>ear<br />

static) cascade model (French et al. 1993) was developed to simulate the voltage<br />

responses <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong>. The two nonl<strong>in</strong>ear components were polynomial<br />

functions and the l<strong>in</strong>ear component was the Wong and Knight photoreceptor model<br />

(Wong et al. 1980):<br />

1 ⎛ t ⎞<br />

g(t) = ⎜ ⎟ e<br />

n! τ ⎝τ ⎠<br />

n<br />

-t/ τNLN<br />

NLN NLN<br />

where n and τNLN are parameters to be fitted. To remove redundant parameters, only one<br />

constant term was <strong>in</strong>cluded <strong>in</strong> the model, as an offset <strong>in</strong> the output of the l<strong>in</strong>ear<br />

component. Similarly, the first nonl<strong>in</strong>ear component had a fixed first-order coefficient of<br />

unity. The numbers of unknown parameters were therefore: N1-1 (first polynomial) + 3<br />

(Wong and Knight model plus offset) + N2 (second polynomial), where N1 and N2 are the<br />

order of the polynomials. The NLN cascade model was fitted by us<strong>in</strong>g the simulated<br />

anneal<strong>in</strong>g method (Press et al. 1990). The first nonl<strong>in</strong>earity required a fifth-order<br />

polynomial, whereas the second polynomial was third-order, giv<strong>in</strong>g a total count of 10<br />

(16)<br />

(18)


29<br />

parameters. Nonl<strong>in</strong>ear kernel analysis and NLN modell<strong>in</strong>g were carried out <strong>in</strong><br />

collaboration with Prof. Andrew French (Department of Physiology and Biophysics,<br />

Dalhousie University, Halifax, Nova Scotia, Canada).<br />

4.6 Hodgk<strong>in</strong>-Huxley type photoreceptor models<br />

A Hodgk<strong>in</strong>-Huxley type model of the <strong>photoreceptors</strong> was developed us<strong>in</strong>g MATLAB<br />

software (Mathworks, USA). Photoreceptor models used <strong>in</strong> the orig<strong>in</strong>al papers I-III<br />

<strong>in</strong>cluded Shaker and slow delayed rectifier voltage-<strong>gated</strong> potassium conductances, <strong>in</strong><br />

addition to K + and Cl - chloride leak conductances (the fast delayed rectifier conductance<br />

was omitted because it is present only <strong>in</strong> some <strong>photoreceptors</strong>) (Hardie 1991a).<br />

Photoreceptor models used <strong>in</strong> the orig<strong>in</strong>al paper IV also <strong>in</strong>cluded a slowly activat<strong>in</strong>g,<br />

non-<strong>in</strong>activat<strong>in</strong>g novel voltage-<strong>gated</strong> K + conductance and the partial failure of Shaker<br />

channel <strong>in</strong>activation (~11 %). The light-dependent conductance due to the<br />

phototransduction cascade was modelled as an additional light leak conductance (light<br />

adapted <strong>photoreceptors</strong>), as a simulation specific dynamic light conductance or as a<br />

reconstructed light <strong>in</strong>duced conductance (LIC; see 4.6.4). The electrical equivalent circuit<br />

for the photoreceptor models is shown <strong>in</strong> Fig 4.<br />

Fig. 4. Electrical circuit of the photoreceptor models. Abbreviations: Sh, Shaker channel; Dr,<br />

delayed rectifier channel; Novel, novel K + channel; Kleak, potassium leak conductance; Cl,<br />

chloride leak conductance; LIC, light <strong>gated</strong> conductance. The gNovel branch does not exist <strong>in</strong><br />

the photoreceptor models used <strong>in</strong> (I)-(III).<br />

4.6.1 <strong>Voltage</strong>-dependent properties of Shaker, slow delayed rectifier and<br />

novel non<strong>in</strong>activat<strong>in</strong>g K + <strong>channels</strong><br />

Steady-state activation and <strong>in</strong>activation curves for voltage-<strong>gated</strong> K + <strong>channels</strong> were fitted<br />

with the Boltzmann function:<br />

⎛ g ⎞ 1<br />

⎜ ⎟=<br />

⎝g⎠ 1+ e<br />

∞<br />

V50 −Vm<br />

max s<br />

(19)


30<br />

where V50 is the voltage produc<strong>in</strong>g a steady-state conductance of 50 % of the maximum<br />

value and s is the slope factor of the function. Steady-state activation and <strong>in</strong>activation<br />

curves for Shaker and slow delayed rectifier <strong>channels</strong> were obta<strong>in</strong>ed from previously<br />

published data (Hardie 1991a, Hevers & Hardie 1995) (Figure 5a). The activation curve<br />

for the novel voltage-<strong>gated</strong> K + conductance was fitted to experimental data from the<br />

whole-cell record<strong>in</strong>gs of Shaker 14 Shab 3 mutant <strong>photoreceptors</strong> (Figure 5b). The fraction<br />

of Shaker <strong>channels</strong> fail<strong>in</strong>g to <strong>in</strong>activate was determ<strong>in</strong>ed from the whole-cell currents of<br />

Shab 3 mutant <strong>photoreceptors</strong> (Fig 1c, IV).<br />

Fig. 5. Steady-state activation and <strong>in</strong>activation curves. (a) Shaker and slow delayed rectifier<br />

(dashed l<strong>in</strong>e). (b) Experimental data po<strong>in</strong>ts (star) and the fitted activation curve for the novel<br />

K + conductance<br />

The time constants of the activation and <strong>in</strong>activation for the Shaker and slow delayed<br />

rectifier <strong>channels</strong> were derived either from previously published data (Hardie 1991a,<br />

Hevers & Hardie 1995) or from additional whole-cell record<strong>in</strong>gs (I). The time constant of<br />

activation for the novel K + conductance was derived from the whole-cell record<strong>in</strong>gs of<br />

Shaker 14 Shab 3 mutant <strong>photoreceptors</strong> (IV). Experimental data was fitted with the bell<br />

shaped function:<br />

1<br />

τ= p (20)<br />

2−Vm p p ( )<br />

3 4⋅ p5 −Vm<br />

p1⋅ e + p5−Vm p6<br />

e −1<br />

where pi are the free parameters for fitt<strong>in</strong>g (Fig. 6). The time constant of the slow delayed<br />

rectifier <strong>in</strong>activation is not significantly voltage-dependent (Hardie 1991a) and it was set<br />

to 1200 ms <strong>in</strong> the model. All time constants were scaled down by a factor 1.35 (calculated<br />

from the Q10 factor; (Juusola & Hardie 2001b), to take <strong>in</strong>to account the temperature<br />

difference between <strong>in</strong> vitro whole-cell (20 °C) and <strong>in</strong> vivo <strong>in</strong>tracellular (25 °C)<br />

record<strong>in</strong>gs.


31<br />

Fig. 6. Fitted activation and <strong>in</strong>activation time constants for the experimental data (stars) (a)<br />

Shaker activation, (b) Shaker <strong>in</strong>activation, (c) Slow delayed rectifier activation, (d) Novel K +<br />

conductance activation<br />

Based on the orig<strong>in</strong>al Hodgk<strong>in</strong>-Huxley formalism, follow<strong>in</strong>g equations for the rate<br />

constants, αk and βk were derived:<br />

1<br />

nk<br />

⎡⎛g∞⎞ ⎤<br />

(21)<br />

⎢⎜ g ⎟<br />

⎝ max ⎠<br />

⎥<br />

k<br />

α<br />

⎣ ⎦<br />

k =<br />

τk<br />

( )<br />

1<br />

nk<br />

⎡⎛g∞⎞ ⎤<br />

⎢ g ⎥<br />

⎣⎝ max ⎠k⎦<br />

1−<br />

⎜ ⎟<br />

β k =<br />

τ<br />

k<br />

here g∞g are the steady-state activation and <strong>in</strong>activation curves of each voltage-<br />

max k<br />

<strong>gated</strong> conductance, nk and τk are the correspond<strong>in</strong>g number of gat<strong>in</strong>g particles and the<br />

time constants, respectively. Differential equations for the membrane voltage and gat<strong>in</strong>g<br />

particles, <strong>in</strong> addition to the equation for the conductance, followed traditional Hodgk<strong>in</strong>-<br />

Huxley formalism (equations (7)-(9), respectively). Parameters of the voltage-dependent<br />

properties were fixed throughout the simulations and are given <strong>in</strong> Table 1.<br />

(22)


32<br />

Table 1. <strong>Voltage</strong>-dependent properties of the K + <strong>channels</strong>. Act and Inact refer to<br />

activation and <strong>in</strong>activation, respectively.<br />

Variable<br />

V50 (mV) -23.7<br />

Shaker Shab Novel K + channel<br />

Act Inact * Act Inact Act<br />

1st -55.3<br />

2nd -74.8<br />

-1.0 -25.7 -14.0<br />

S (mV) 12.8<br />

1st -3.9<br />

2nd -10.7<br />

9.1 -6.4 10.6<br />

n 3 1 2 1 1<br />

* Shaker <strong>in</strong>activation had two components<br />

4.6.2 Derivation of the other model parameters<br />

In the photoreceptor models, the maximum conductance for the voltage-<strong>gated</strong> K +<br />

<strong>channels</strong>, the size of the leak conductances, the membrane capacitance and the current<br />

<strong>in</strong>put for the simulations are given as per area. To be able to compare the simulations with<br />

the experiments, the value of the membrane area, Am, was def<strong>in</strong>ed. The membrane area<br />

used <strong>in</strong> the model (1.571*10 -5 cm 2 ) was calculated accord<strong>in</strong>g to average photoreceptor<br />

dimensions (length 100 µm, diameter 5 µm).<br />

The reversal potential for the potassium (-85 mV) and light leak conductances (+10<br />

mV) were obta<strong>in</strong>ed from published data (Hardie 1991a, Reuss et al. 1997, Oberw<strong>in</strong>kler &<br />

Stavenga 2000). The reversal potential for the Cl - leak conductance was obta<strong>in</strong>ed through<br />

iteration (the reversal potential for the non-potassium leak and the sizes of these two<br />

different leak conductances were adjusted until the experimental rest<strong>in</strong>g potential and<br />

<strong>in</strong>put resistance at rest were reproduced). A reversal potential of -30 mV was obta<strong>in</strong>ed,<br />

which was concluded to represent the chloride ions (Hardie, unpublished data). The<br />

reversal potentials of each ion and the membrane area were fixed parameters.<br />

The properties of the <strong>photoreceptors</strong>, like the rest<strong>in</strong>g potential, the <strong>in</strong>put resistance and<br />

the capacitance of the cell, are highly variable from one record<strong>in</strong>g to another. This is<br />

because of the natural variability between the cells, but also because of the vary<strong>in</strong>g<br />

quality of the experiments. All photoreceptor models were fitted based on the simulated<br />

current <strong>in</strong>jection experiments. To take <strong>in</strong>to account the variation <strong>in</strong> photoreceptor<br />

capacitance, first order exponentials were fitted to the voltage response of a small<br />

hyperpolaris<strong>in</strong>g current <strong>in</strong>jection and the specific membrane capacitance, cm, was<br />

calculated with:<br />

(23)<br />

where τm is the membrane time constant, I the amplitude of current <strong>in</strong>jection, Am the<br />

membrane area and U the amplitude of the voltage response. The K + and Cl - τmI<br />

cm<br />

=<br />

AmU leak<br />

conductances were adjusted until experimental rest<strong>in</strong>g potential and <strong>in</strong>put resistance<br />

followed. The maximum conductance of each voltage-<strong>gated</strong> channel was adjusted to


33<br />

reproduce the best fit with the experimental data. Exemplary values for these free<br />

parameters are given <strong>in</strong> the Table 2.<br />

Table 2. Example set of the free parameters.<br />

Parameter Value<br />

Rest<strong>in</strong>g potential -66 mV<br />

Specific membrane capacitance * 4 µF/cm 2<br />

Maximum Shaker conductance 0.8 mS/cm 2<br />

Maximum Shab conductance 3.0 mS/cm 2<br />

Maximum novel K + conductance 0.11 mS/cm 2<br />

Potassium leak conductance 0.0855 mS/cm 2<br />

Chloride leak conductance 0.0585 mS/cm 2<br />

*<br />

The contribution of the microvillar membrane is <strong>in</strong>cluded <strong>in</strong> this parameter<br />

4.6.3 MATLAB© implementation of the Hodgk<strong>in</strong>-Huxley type of<br />

photoreceptor model<br />

MATLAB © software provides all the basic and a vast variety of specialised functions (mfiles)<br />

for example for the statistics, optimisation and signal process<strong>in</strong>g. I used it as a<br />

model development environment, because it also has efficient, <strong>in</strong>-built functions for<br />

solv<strong>in</strong>g differential equations. As an example, m-files of the WT model used <strong>in</strong> current<br />

<strong>in</strong>jection simulations (Fig 4a, IV) are given <strong>in</strong> Appendix.<br />

Normally, numerical <strong>in</strong>tegration of the differential equation models works well for the<br />

smooth <strong>in</strong>put (gradually chang<strong>in</strong>g, cont<strong>in</strong>uous derivatives). In this study it was necessary<br />

to drive the models with highly irregular <strong>in</strong>put, such as white noise. To ensure that the<br />

models could be driven with these dynamic <strong>in</strong>puts, all the voltage-<strong>gated</strong> conductances<br />

were set to zero thereby reduc<strong>in</strong>g the model to an analytically solvable RC-circuit.<br />

Comparison of the simulations with the exact solution of the RC-circuit (step response<br />

and frequency response function) validated that the numerical methods used for<br />

simulations did not <strong>in</strong>troduce significant errors.<br />

4.6.4 Reconstruction of the light <strong>in</strong>duced conductance<br />

Assum<strong>in</strong>g that the photo-<strong>in</strong>sensitive membrane of the photoreceptor is accurately<br />

modelled, it is possible to use the model for reconstruct<strong>in</strong>g the light <strong>in</strong>duced conductance<br />

(LIC) from the correspond<strong>in</strong>g experimental voltage response. Although the voltagedependent<br />

properties of the K + <strong>channels</strong> are characterized <strong>in</strong> darkness, these properties<br />

were assumed to be <strong>in</strong>sensitive to light. This was justified both by the comparison of<br />

simulated voltage responses to current stimuli with experimental data, and by the<br />

experimental results from Calliphora vic<strong>in</strong>a <strong>photoreceptors</strong> (Weckström et al. 1991).


34<br />

The absorption of a s<strong>in</strong>gle photon happens <strong>in</strong> an all-or-none fashion, caus<strong>in</strong>g open<strong>in</strong>g<br />

of light-dependent <strong>channels</strong> followed by the <strong>in</strong>flux of calcium and sodium ions <strong>in</strong> a<br />

microvillus. This leads to rapid changes <strong>in</strong> ionic concentrations due to the small volume<br />

of the microvillus. Dur<strong>in</strong>g light stimulation the photon flux activates numerous<br />

microvilli, caus<strong>in</strong>g dynamic reversal potential fluctuations for the LIC <strong>in</strong> each microvillus<br />

as it receives photons. These fluctuations were assumed to be <strong>in</strong>dependent, averag<strong>in</strong>g to<br />

the reversal potential of +10 mV (Reuss et al. 1997, Oberw<strong>in</strong>kler & Stavenga 2000).<br />

Us<strong>in</strong>g this reversal potential, the LIC was modelled as a s<strong>in</strong>gle conductance <strong>in</strong>put to the<br />

photo-<strong>in</strong>sensitive membrane. This conductance was used to drive the model and its values<br />

were iterated at each sample po<strong>in</strong>t until the experimental voltage response was<br />

reproduced. A schematic presentation of the reconstruction procedure is shown <strong>in</strong> Fig. 6b<br />

(IV) and the correspond<strong>in</strong>g model m-files are given <strong>in</strong> the Appendix.<br />

Predicted LICs always produced voltage responses that closely matched experimental<br />

data. Small discrepancies were observed <strong>in</strong> simulations with the reconstructed LIC used<br />

<strong>in</strong> the naturalistic stimuli simulations dur<strong>in</strong>g the periods of darkness preceded by high<br />

light <strong>in</strong>tensities (Fig. 2, III). In such conditions, the experimental data conta<strong>in</strong>s clear afterhyperpolarisations,<br />

as well as small depolarisations preced<strong>in</strong>g the afterhyperpolarisations.<br />

These are caused by the Na + /K + ATPase and the 3Na + /Ca 2+ exchanger<br />

currents, respectively (Gerster 1997, Gerster et al. 1997, Oberw<strong>in</strong>kler & Stavenga 2000).<br />

These discrepancies were concluded not to affect the results, based on their small size and<br />

the nonl<strong>in</strong>ear analysis (III).<br />

4.6.5 Model simulations<br />

To simulate the current <strong>in</strong>jection experiments (for example, Fig. 4, I), models were driven<br />

with the experimental current stimuli divided by the photoreceptor membrane area.<br />

Modell<strong>in</strong>g enabled the monitor<strong>in</strong>g of each <strong>in</strong>dividual conductance dur<strong>in</strong>g the simulations.<br />

This feature was used to separate the effect of each voltage-<strong>gated</strong> conductance to the<br />

voltage response (I-IV). Impedance functions for the models were determ<strong>in</strong>ed by us<strong>in</strong>g<br />

band-limited Gaussian white noise current <strong>in</strong>put that was filtered by a fourth-order<br />

Butterworth low pass filter with 500 Hz corner frequency (I). The frequency band of this<br />

<strong>in</strong>put was the same as <strong>in</strong> the experiments and it exceeds that of the <strong>photoreceptors</strong><br />

(Juusola & Hardie 2001a). The current-voltage (I-V) relationships for the models <strong>in</strong> (IV)<br />

were determ<strong>in</strong>ed by driv<strong>in</strong>g them with 70 evenly spaced, 100 ms current pulses, rang<strong>in</strong>g<br />

from -0.15 nA to 0.4 nA and sampl<strong>in</strong>g the membrane voltage at the end of each pulse.<br />

To simulate photoreceptor responses <strong>in</strong> the light adaptation, light dependent leak was<br />

added until the same amount of depolarization was achieved as <strong>in</strong> the experiments. The<br />

log-normal shape of the light conductance pulses used <strong>in</strong> (I) and (II) were fitted to<br />

experimentally derived light impulse responses and the size of the pulses was adjusted so<br />

that the largest conductance pulse produced a saturated voltage response hav<strong>in</strong>g<br />

amplitude similar to those <strong>in</strong> the experiments (Juusola & Hardie 2001a). For the LICvoltage<br />

(LIC-V) relationships <strong>in</strong> (IV), 100 ms LIC pulses of <strong>in</strong>creas<strong>in</strong>g amplitude were<br />

used as the model <strong>in</strong>put and the correspond<strong>in</strong>g voltages were measured relative to the<br />

dark rest<strong>in</strong>g potential and normalized by the amplitude of the saturat<strong>in</strong>g maximal


35<br />

response. The simulated naturalistic stimuli experiments (III) and the determ<strong>in</strong>ation of the<br />

contrast ga<strong>in</strong> (frequency response function between the white noise modulated light<br />

contrast and the voltage response) for the photoreceptor models were carried out by us<strong>in</strong>g<br />

the reconstructed LIC (IV).<br />

In addition to experimentally fitted photoreceptor models, additional hypothetical<br />

membrane models were also used <strong>in</strong> (I) and (IV). In (I) the hypothetical WT without<br />

Shaker membrane model was used to study the effect of the Shaker <strong>in</strong>activation to the<br />

amplification observed <strong>in</strong> the impedance function (I, supplementary material). In (IV)<br />

two hypothetical membrane models, WT with 30% Shab and WT without Shab, were<br />

generated to <strong>in</strong>vestigate the role of regulation of the electrical properties of the photo<strong>in</strong>sensitive<br />

membrane of Shab mutant <strong>photoreceptors</strong>.<br />

4.6.6 Estimation of the metabolic costs<br />

To estimate the metabolic costs, the electrical circuit model (Fig. 4) was extended to<br />

<strong>in</strong>clude the Na + /K + ATPase and the Na + /K + /2Cl - co-transporter (Fig. 1, II). The activity of<br />

the Na + /K + ATPase, the pump current (IP), determ<strong>in</strong>es the cost of ma<strong>in</strong>ta<strong>in</strong><strong>in</strong>g the <strong>in</strong>ternal<br />

K + concentration and external Na + concentration. A Na + /K + /2Cl - co-transporter, the<br />

probable Cl - accumulat<strong>in</strong>g mechanism <strong>in</strong> neurones (Delpire 2000), ma<strong>in</strong>ta<strong>in</strong>s the <strong>in</strong>ternal<br />

Cl - concentration. A fraction of the LIC (~20%) is carried by calcium ions (40% of the<br />

current), which are exchanged <strong>in</strong> a 1:3 ratio for sodium ions by the Na + /Ca 2+ exchanger<br />

(Reuss et al. 1997, Oberw<strong>in</strong>kler & Stavenga 2000). The Na + /K + ATPase hydrolyses one<br />

ATP molecule for every three Na + ions expelled and two K + ions imported. Thus, if the<br />

pump ma<strong>in</strong>ta<strong>in</strong>s the concentrations of these ions across the cell membrane, the pump<br />

current can be derived from <strong>in</strong>dividual currents <strong>in</strong> the electrical circuit:<br />

1 1<br />

IP= ( IKSh+ IKdr + IKleak ) − IClleak<br />

(24)<br />

2 4<br />

S<strong>in</strong>ce the membrane potential is stable <strong>in</strong> long term, the sum of all currents across the<br />

model membrane must eventually equal zero:<br />

IKSh + IKdr + IKleak + IClleak + IP+ ILIC= 0<br />

These currents were adjusted to satisfy equation (25) us<strong>in</strong>g a simplex algorithm (function<br />

called fm<strong>in</strong>search <strong>in</strong> MATLAB © ) to m<strong>in</strong>imise the changes for each current. Us<strong>in</strong>g these<br />

adjusted values, the estimated number of ATP molecules hydrolysed per second can be<br />

calculated:<br />

ATP IN P A =<br />

s F<br />

(25)<br />

(26)


36<br />

The number of ATP molecules per bit of <strong>in</strong>formation was calculated by divid<strong>in</strong>g the<br />

estimated number of ATP molecules hydrolysed per second by the measured <strong>in</strong>formation<br />

capacity, C (bits/s):<br />

ATP IN P A F<br />

=<br />

bit C<br />

(27)


5 Summary of the results <strong>in</strong> the orig<strong>in</strong>al papers<br />

5.1 The contribution of Shaker K + <strong>channels</strong> to the <strong>in</strong>formation<br />

capacity of <strong>Drosophila</strong> <strong>photoreceptors</strong> (I)<br />

The performance of a photoreceptor <strong>in</strong> cod<strong>in</strong>g a light signal may be quantitatively<br />

described by its sensitivity, signal-to-noise ratio and frequency response, enabl<strong>in</strong>g<br />

specific components of the signall<strong>in</strong>g mach<strong>in</strong>ery, <strong>in</strong>clud<strong>in</strong>g ion <strong>channels</strong>, to be related to<br />

specific aspects of cellular function (Weckström & Laughl<strong>in</strong> 1995). These quantitative<br />

measures, comb<strong>in</strong>ed with a mathematical model of the <strong>photoreceptors</strong> were used to<br />

assess the contribution of the Shaker K + -conductance to photoreceptor performance.<br />

The signall<strong>in</strong>g efficiency of <strong>photoreceptors</strong> <strong>in</strong> wild type (WT) and Shaker 14 flies, was<br />

studied by present<strong>in</strong>g sequences of dynamically modulated light with a mean contrast of<br />

0.32, close to that of natural sceneries (Laughl<strong>in</strong> 1981), over the range of light <strong>in</strong>tensities<br />

to which the <strong>photoreceptors</strong> are normally exposed (Fig. 1B, I). The basic properties of the<br />

phototransduction process were determ<strong>in</strong>ed to be unaffected <strong>in</strong> Shaker mutant<br />

<strong>photoreceptors</strong>. At all light background <strong>in</strong>tensities, WT flies responded with a larger<br />

signal (measured as the signal variance) than Shaker 14 flies (Fig. 1I, I). There was no<br />

significant difference <strong>in</strong> the noise variance between Shaker 14 and WT flies over all light<br />

<strong>in</strong>tensities (p


38<br />

<strong>in</strong>duced by current <strong>in</strong>jection opens additional Shaker <strong>channels</strong> that <strong>in</strong>activate rapidly,<br />

thereby <strong>in</strong>creas<strong>in</strong>g the membrane resistance beyond the rest<strong>in</strong>g value and produc<strong>in</strong>g a<br />

larger voltage response to the same current step. The loss of the Shaker conductance<br />

would be expected to result <strong>in</strong> a higher <strong>in</strong>put resistance at rest, however, Shaker 14<br />

<strong>photoreceptors</strong> had reduced resistances compared to those of WT <strong>photoreceptors</strong><br />

(Shaker 14 , 225 ± 26 MΩ at -64.3 ± 5.4 mV, N=13; WT, 411 ± 34 MΩ at -68.1 ± 3.2 mV,<br />

N=21). Increase <strong>in</strong> the leak conductance was used to model the decrease <strong>in</strong> the Shaker 14<br />

photoreceptor <strong>in</strong>put resistance (Fig. 4H, I).<br />

The <strong>in</strong>put resistance affects the photoreceptor’s current to voltage ga<strong>in</strong>. The<br />

biophysical properties of the Shaker <strong>channels</strong> should lead to an <strong>in</strong>crease <strong>in</strong> the spread of<br />

the signal across the photoreceptor voltage range. The spread of the voltage responses of<br />

the Shaker 14 photoreceptor was compressed compared to those of the WT, suggest<strong>in</strong>g that<br />

the WT <strong>photoreceptors</strong> were us<strong>in</strong>g the available voltage range more fully than Shaker 14<br />

<strong>photoreceptors</strong>. Vary<strong>in</strong>g the size of the leak conductance <strong>in</strong> the WT and Shaker 14<br />

photoreceptor models showed that the amount of leak conductance found experimentally,<br />

i.e., the <strong>in</strong>put resistance of the both photoreceptor types, optimised the available voltage<br />

ranges (Fig. 5C,D; I).<br />

5.2 Metabolic cost of Shaker K + <strong>channels</strong> <strong>in</strong> <strong>Drosophila</strong><br />

<strong>photoreceptors</strong> (II)<br />

The contribution of the Shaker conductance to the cost of neural <strong>in</strong>formation was<br />

estimated by compar<strong>in</strong>g wild type (WT) and Shaker mutant <strong>photoreceptors</strong>. The currents<br />

flow<strong>in</strong>g through the ion <strong>channels</strong> either <strong>in</strong> steady-state or dur<strong>in</strong>g light stimulation were<br />

estimated with the Hodgk<strong>in</strong>-Huxley type of model (I). These currents were subsequently<br />

used to derive metabolic costs by us<strong>in</strong>g an electrical circuit model of the <strong>photoreceptors</strong><br />

that also <strong>in</strong>cluded the Na + /K + pump and the Na + /K + /2Cl - co-transporter (Fig. 2, II). The<br />

estimated cost of ma<strong>in</strong>ta<strong>in</strong><strong>in</strong>g the ionic gradients <strong>in</strong> both dark and light conditions was<br />

greater <strong>in</strong> Shaker mutant than <strong>in</strong> the WT (Table 1, II). The <strong>in</strong>creased rate of ATP<br />

hydrolysis <strong>in</strong> Shaker <strong>photoreceptors</strong> coupled with their reduced <strong>in</strong>formation capacity led<br />

to a ~2-fold <strong>in</strong>crease <strong>in</strong> their bit cost when compared to their WT counterparts. The<br />

calculated bit costs may be an underestimate for the <strong>Drosophila</strong> <strong>photoreceptors</strong> dur<strong>in</strong>g<br />

natural stimuli because of the lower <strong>in</strong>formation transmission rates and the energy<br />

required for early steps <strong>in</strong> phototransduction (e.g., phosphorylation or amplification)<br />

(Laughl<strong>in</strong> et al. 1998, Hardie 2001, Juusola & de Polavieja 2003). However, similar<br />

calculations for Calliphora vic<strong>in</strong>a <strong>photoreceptors</strong> (Laughl<strong>in</strong> et al. 1998) produce<br />

estimates of metabolic cost close to experimentally determ<strong>in</strong>ed values (Hamdorf et al.<br />

1988), <strong>in</strong>dicat<strong>in</strong>g that the estimates of the bit cost were valid.<br />

The <strong>in</strong>creased bit cost <strong>in</strong> Shaker 14 <strong>photoreceptors</strong> was due to reduced <strong>in</strong>put resistance<br />

follow<strong>in</strong>g from the loss of the Shaker <strong>channels</strong>. In (I) the amount of leak conductance <strong>in</strong><br />

the model needed to reproduce the experimentally determ<strong>in</strong>ed <strong>in</strong>put resistances were<br />

shown to maximise the available voltage range <strong>in</strong> Shaker 14 and WT <strong>photoreceptors</strong>. The<br />

metabolic cost may also be an important factor affect<strong>in</strong>g the regulation of the <strong>in</strong>put<br />

resistance. Both Shaker 14 and WT <strong>photoreceptors</strong> were optimised <strong>in</strong> terms of metabolic


39<br />

cost and available voltage range (Fig. 2, II), however, WT <strong>photoreceptors</strong> <strong>in</strong>curred a<br />

lower metabolic cost due to the Shaker conductance.<br />

5.3 Interactions between light-<strong>in</strong>duced currents, voltage-<strong>gated</strong><br />

currents, and <strong>in</strong>put signal properties <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong><br />

(III)<br />

In vivo <strong>in</strong>tracellular record<strong>in</strong>gs from WT and Shaker 14 <strong>photoreceptors</strong> were carried out<br />

while present<strong>in</strong>g light modulated by a Natural Time Series of Intensities (NTSIs). The<br />

naturalistic stimuli conta<strong>in</strong>ed periods of both very low and high light <strong>in</strong>tensity that were<br />

reflected <strong>in</strong> the voltage responses of both the WT and mutant <strong>photoreceptors</strong> (Fig. 1A,<br />

III). Differences <strong>in</strong> the voltage responses over the entire stimulus sequences are shown <strong>in</strong><br />

the response histograms (Fig. 1C, III), and the cumulative frequency distributions of<br />

these histograms (Fig. 1D, III), which demonstrate that the voltage response of WT<br />

<strong>photoreceptors</strong> was spread over a greater voltage range than the mutant.<br />

A nonl<strong>in</strong>ear analysis was used to characterize the effects of ion <strong>channels</strong> upon natural<br />

stimulus cod<strong>in</strong>g <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong> (Figs. 2-4, III). First-order kernels, K1(u),<br />

derived from WT and Shaker 14 voltage responses to naturalistic stimuli had the same<br />

general form as flash responses <strong>in</strong> light-adapted <strong>Drosophila</strong> <strong>photoreceptors</strong> (Fig. 3A, III;<br />

(Juusola & Hardie 2001a). All of the first-order kernels had a delay of approximately<br />

6 ms, but the kernels derived from Shaker 14 voltage responses consistently were larger,<br />

had faster time-to-peak (approximately 3-4 ms) and narrower half-width than those from<br />

WT (Fig. 3A, III). These differences may be expla<strong>in</strong>ed, at least partially, by the presence<br />

or absence of Shaker <strong>channels</strong>. An <strong>in</strong>creas<strong>in</strong>g light <strong>in</strong>tensity depolarizes WT<br />

<strong>photoreceptors</strong>, rapidly activates Shaker <strong>channels</strong>, which partially shunts the LIC through<br />

the <strong>in</strong>creased membrane conductance. The Shaker K + <strong>channels</strong> then rapidly <strong>in</strong>activate,<br />

amplify<strong>in</strong>g the effect of the LIC at later times <strong>in</strong> the photoreceptor voltage response.<br />

Second-order kernels, K2(u,v), had positive peaks on the diagonals, hav<strong>in</strong>g approximately<br />

the same time courses as the first-order kernels (Fig. 3D,E, III).<br />

The pattern of kernel forms suggested that the WT and Shaker 14 photoreceptor<br />

responses could be described by a model <strong>in</strong> which a dynamic l<strong>in</strong>ear filter is followed by a<br />

static nonl<strong>in</strong>earity (Korenberg & Hunter 1986, French & Korenberg 1989). The NLN<br />

model was developed with a ‘Wong and Knight’ or gamma function model of<br />

phototransduction (Wong et al. 1980) as the l<strong>in</strong>ear dynamic component surrounded by<br />

two static nonl<strong>in</strong>earities (Figs. 2, 4, III). Shaker 14 <strong>photoreceptors</strong> differed most strongly<br />

from WT <strong>photoreceptors</strong> <strong>in</strong> the second and third stages of the NLN model (Fig. 4, Table<br />

2; III). The first static nonl<strong>in</strong>earity was a positive rectification and the l<strong>in</strong>ear components<br />

were very similar to first-order Volterra kernels. The f<strong>in</strong>al static nonl<strong>in</strong>earity always had a<br />

region of zero, or negative slope <strong>in</strong> the middle of an overall positive characteristic,<br />

correspond<strong>in</strong>g to an <strong>in</strong>termediate region where <strong>in</strong>creas<strong>in</strong>g light <strong>in</strong>tensity produced no<br />

change, or a slight hyperpolarisation, <strong>in</strong> the membrane potential. In Shaker 14<br />

<strong>photoreceptors</strong> this effect occurred over a smaller stimulus range and at lower stimulus<br />

levels, correspond<strong>in</strong>g to less depolarized membrane potentials.


40<br />

The Hodgk<strong>in</strong>-Huxley type models of the WT and Shaker 14 <strong>photoreceptors</strong> were used<br />

to determ<strong>in</strong>e if possible differences <strong>in</strong> the LIC or voltage-<strong>gated</strong> K + <strong>channels</strong> contribute to<br />

the observed differences between WT and Shaker 14 responses. These models were used to<br />

reconstruct currents generated by the LIC, slow delayed rectifier currents and, <strong>in</strong> WT<br />

<strong>photoreceptors</strong>, the Shaker current to the NTSI stimulus (Fig. 5A, D, F; III). Volterra<br />

kernels between the naturalistic light stimulus and the LIC were calculated (Fig. 6, Table<br />

3; III). Comparison of LIC kernels showed that the kernel of the Shaker mutant was<br />

larger and faster than that of WT, which was also true for Volterra kernels from the<br />

responses (Fig. 3, III). Similarity between the LIC and voltage response kernels suggests<br />

that the LIC dom<strong>in</strong>ates the dynamic response.<br />

The NLN model was developed also for the LIC (Fig. 7, Table 4; III). As for the<br />

Volterra kernels, the l<strong>in</strong>ear Wong and Knight component, which may represent LIC<br />

generation, was larger and had a faster time-to-peak <strong>in</strong> Shaker mutant <strong>photoreceptors</strong><br />

suggest<strong>in</strong>g possible differences <strong>in</strong> the LIC of the two phenotypes. Association of the f<strong>in</strong>al<br />

nonl<strong>in</strong>earity with the photo-<strong>in</strong>sensitive membrane was supported by comparison of WT<br />

and Shaker 14 NLN models for the LIC with those of the voltage response. The f<strong>in</strong>al<br />

nonl<strong>in</strong>earity conta<strong>in</strong>s significant differences <strong>in</strong> the NLN models of the voltage responses<br />

that are absent <strong>in</strong> the NLN models of the LICs suggest<strong>in</strong>g that the f<strong>in</strong>al nonl<strong>in</strong>earity is<br />

most affected by the properties of the photo-<strong>in</strong>sensitive membrane.<br />

Comparison of the slow delayed rectifier currents from WT and Shaker 14 models<br />

revealed that <strong>in</strong> the absence of the Shaker current there was a marked <strong>in</strong>crease <strong>in</strong> the slow<br />

delayed rectifier current (Fig. 5D-G, III). Additionally, <strong>in</strong> WT the delayed rectifier current<br />

was slower than <strong>in</strong> the Shaker 14 mutant (Fig. 5G, III). The Shaker <strong>channels</strong> effectively<br />

shunt the photoreceptor currents, not only reduc<strong>in</strong>g the overall size of the slow delayed<br />

rectifier current but also chang<strong>in</strong>g the time course of activation.<br />

5.4 Role of Shab K + <strong>channels</strong> <strong>in</strong> the <strong>in</strong>formation process<strong>in</strong>g of<br />

<strong>Drosophila</strong> <strong>photoreceptors</strong> (IV)<br />

<strong>Voltage</strong> clamp record<strong>in</strong>gs of isolated <strong>photoreceptors</strong> from Shab 3 flies were performed to<br />

show that the slow delayed rectifier was encoded by the Shab gene <strong>in</strong> <strong>Drosophila</strong><br />

<strong>photoreceptors</strong>. To ensure that mutation of the Shab gene did not affect photoreceptor<br />

morphology we illum<strong>in</strong>ated the <strong>photoreceptors</strong> antidromically to determ<strong>in</strong>e the structure<br />

and spac<strong>in</strong>g of their rhabdomeres. Each Shab 3 ommatidium conta<strong>in</strong>ed a standard<br />

configuration of <strong>photoreceptors</strong> with normal rhabdoms, and the phototransduction<br />

cascade was shown to be unaffected <strong>in</strong> the mutant (Fig. 1, IV).<br />

In vivo current <strong>in</strong>jection experiments showed that Shab 3 <strong>photoreceptors</strong> had different<br />

passive membrane properties from those of the WT (Fig. 2A,D; IV). As <strong>in</strong> the previous<br />

studies (I-III), Hodgk<strong>in</strong>-Huxley type modell<strong>in</strong>g was used to isolate the effect of the Shab<br />

<strong>channels</strong> on the photo-<strong>in</strong>sensitive membrane properties. Comparison between the model<br />

simulations and the experimental Shab 3 photoreceptor responses suggested that, <strong>in</strong><br />

addition to changes <strong>in</strong> the passive membrane properties, an additional voltage-<strong>gated</strong><br />

conductance was present. To isolate this conductance, currents were recorded from<br />

Shaker 14 Shab 3 double mutant <strong>photoreceptors</strong> <strong>in</strong> whole cell patch-clamp. In some


41<br />

Shaker 14 Shab 3 double mutant <strong>photoreceptors</strong> the fast delayed rectifier voltage-<strong>gated</strong><br />

potassium current was present, show<strong>in</strong>g that it was not encoded by the Shab gene (Fig.<br />

3A, IV). An additional, novel, slowly activat<strong>in</strong>g, non-<strong>in</strong>activat<strong>in</strong>g potassium current was<br />

also present <strong>in</strong> Shaker 14 Shab 3 <strong>photoreceptors</strong> (Fig. 3A, IV) and, <strong>in</strong> those <strong>photoreceptors</strong><br />

lack<strong>in</strong>g the fast delayed rectifier, appeared to be the only rema<strong>in</strong><strong>in</strong>g voltage-<strong>gated</strong><br />

potassium current (Fig. 3B, IV). The voltage-dependent properties of this conductance<br />

were characterized (Fig. 3C-E, IV) and it was <strong>in</strong>corporated <strong>in</strong>to the photoreceptor<br />

models. Close exam<strong>in</strong>ation of Shab 3 photoreceptor whole cell currents (Fig. 1C, IV)<br />

revealed a component that was not due to the novel voltage-<strong>gated</strong> K + conductance (Fig.<br />

3A,B; IV). These residual currents followed from partial failure of the Shaker channel<br />

<strong>in</strong>activation (~11%), a feature, which was also <strong>in</strong>corporated to the models.<br />

The I-V relationships of the WT and Shab mutant <strong>photoreceptors</strong> were simulated to<br />

<strong>in</strong>vestigate if the altered passive membrane properties contribute to the ma<strong>in</strong>tenance of<br />

the electrical properties. To model the changes <strong>in</strong> the passive membrane, an <strong>in</strong>crease <strong>in</strong><br />

the leak conductances <strong>in</strong> the Shab photoreceptor models was needed to reproduce the<br />

correspond<strong>in</strong>g experimental data. The I-V relationships of the Shab photoreceptor models<br />

were shown to closely match those of the WT over the physiological voltage range (-70<br />

mV to -15 mV) (Fig. 4C-F, IV).<br />

To quantify the reduction <strong>in</strong> <strong>in</strong>put resistance suggested by the <strong>in</strong>creased leak<br />

conductances <strong>in</strong> the models, the mean <strong>in</strong>put resistances <strong>in</strong> dark for the WT, Shab 1 and<br />

Shab 3 flies were calculated (Fig. 5A, IV). Shab 1 and Shab 3 <strong>photoreceptors</strong> had<br />

significantly reduced <strong>in</strong>put resistances (-54% and -59% respectively), although Shab<br />

<strong>channels</strong> are not activated at the photoreceptor rest<strong>in</strong>g potential (-66 mV). Photoreceptor<br />

models were used to predict the correspond<strong>in</strong>g <strong>in</strong>put resistances <strong>in</strong> fully light-adapted<br />

<strong>photoreceptors</strong> at -40 mV (Fig. 5A, IV). The Shab 1 and Shab 3 photoreceptor <strong>in</strong>put<br />

resistances <strong>in</strong> light were close to that of the WT photoreceptor despite the large<br />

differences <strong>in</strong> the dark. Therefore, the light-voltage relationship of the WT, Shab 1 and<br />

Shab 3 <strong>photoreceptors</strong> were compared over a wide range of light <strong>in</strong>tensities. Over all<br />

adapt<strong>in</strong>g light backgrounds, the mean steady-state membrane potentials of WT, Shab 1 and<br />

Shab 3 <strong>photoreceptors</strong> showed a similar light dependency and were not significantly<br />

different (Fig. 5B, IV).<br />

The LIC <strong>in</strong>put-voltage output relationships of the Shab 1 and Shab 3 photoreceptor<br />

models as well as those of their hypothetical counterparts (reduced Shab conductance but<br />

passive membrane properties same as <strong>in</strong> the WT models) were determ<strong>in</strong>ed to study how<br />

regulation of the electrical properties of the photo-<strong>in</strong>sensitive membrane contribute to the<br />

observed robustness <strong>in</strong> the light-voltage relationship (Fig. 5B, IV). In contrast to the<br />

hypothetical models, the LIC-voltage relationships of the experimentally fitted ones<br />

<strong>in</strong>tersected with that of the WT at a LIC value produc<strong>in</strong>g a voltage response 50% of the<br />

maximum (Fig. 5D, IV). Comparison of WT with the Shab 1 and Shab 3 photoreceptor<br />

LIC-voltage relationships showed that the voltage-dependent Shab conductance produces<br />

a more efficient use of the available voltage range compared to the voltage-<strong>in</strong>dependent<br />

leak conductances used to model the regulatory reduction <strong>in</strong> the <strong>in</strong>put resistance (Fig. 5D,<br />

IV). However, due to this regulation, both Shab 1 and Shab 3 <strong>photoreceptors</strong> have an<br />

<strong>in</strong>creased spread of the voltages due to the LIC over the available voltage range <strong>in</strong><br />

comparison to their respective hypothetical models (Fig. 5C, IV).


42<br />

The voltage responses of WT, Shab 1 and Shab 3 <strong>photoreceptors</strong> were recorded whilst<br />

present<strong>in</strong>g repetitions of the white noise modulated light stimulus sequence to determ<strong>in</strong>e<br />

how changes <strong>in</strong> the photo-<strong>in</strong>sensitive membrane of the Shab mutant <strong>photoreceptors</strong> affect<br />

the process<strong>in</strong>g of dynamically fluctuat<strong>in</strong>g light stimuli. The LIC was reconstructed from<br />

the Shab 3 photoreceptor record<strong>in</strong>gs and used to determ<strong>in</strong>e the contrast ga<strong>in</strong>s of the WT,<br />

Shab 1 and Shab 3 models. These contrast ga<strong>in</strong>s closely matched those of the<br />

experimentally determ<strong>in</strong>ed ones (Fig. 6D,E; IV), show<strong>in</strong>g that regulation of the electrical<br />

properties enables ma<strong>in</strong>tenance of the contrast ga<strong>in</strong> <strong>in</strong> Shab 1 and Shab 3 <strong>photoreceptors</strong>.<br />

Photoreceptor performance is constra<strong>in</strong>ed not only by signal ga<strong>in</strong> but also by the<br />

noise. To quantify the reliability of contrast cod<strong>in</strong>g, the mean <strong>in</strong>formation capacity of the<br />

WT, Shab 1 and Shab 3 <strong>photoreceptors</strong> was calculated over the range of light <strong>in</strong>tensity<br />

backgrounds (Fig. 7C, IV). The mean <strong>in</strong>formation capacities were similar over almost all<br />

light <strong>in</strong>tensities for the WT, Shab 1 and Shab 3 <strong>photoreceptors</strong>. In particular, the<br />

convergence of <strong>in</strong>formation capacities <strong>in</strong> high light <strong>in</strong>tensities where changes <strong>in</strong> the<br />

photo-<strong>in</strong>sensitive membrane were shown to ma<strong>in</strong>ta<strong>in</strong> light-voltage relationships <strong>in</strong> Shab<br />

mutants (Fig 5, IV), demonstrates the robustness of contrast cod<strong>in</strong>g <strong>Drosophila</strong><br />

<strong>photoreceptors</strong>.


6 Discussion<br />

6.1 Modell<strong>in</strong>g of <strong>Drosophila</strong> <strong>photoreceptors</strong><br />

The Hodgk<strong>in</strong>-Huxley type of models used <strong>in</strong> this study are shown to be capable for<br />

reproduc<strong>in</strong>g the experimentally determ<strong>in</strong>ed photoreceptor responses, which justifies their<br />

essential role <strong>in</strong> provid<strong>in</strong>g evidence for the conclusions presented. However, there are<br />

limitations with this modell<strong>in</strong>g approach that one has to consider. Hodgk<strong>in</strong>-Huxley type<br />

of models are macroscopic descriptions of the ion <strong>channels</strong> on the isopotential cell<br />

membrane, <strong>in</strong> contrast to the stochastic state-variable or electrically coupled<br />

compartmental models (Koch 1999, Hille 2001). Stochastic modell<strong>in</strong>g was not used<br />

because of the success of the macroscopic approach <strong>in</strong> addition to the drastically<br />

<strong>in</strong>creased complexity and computational burden of stochastic models. Can <strong>Drosophila</strong><br />

<strong>photoreceptors</strong> be assumed as isopotential cells?<br />

The photo-sensitive membrane of the <strong>Drosophila</strong> photoreceptor is formed from tightly<br />

packed tubular microvilli (Fig 2) (Hardie & Raghu 2001). Despite of their small<br />

dimensions, microvilli are shown to be electrically well coupled to the cell body (Postma<br />

et al. 1999). Photoreceptors also have a relatively short axon lack<strong>in</strong>g the sodium spike<br />

generation. However, <strong>in</strong> vitro experiments have shown that the voltage-<strong>gated</strong> potassium<br />

<strong>channels</strong> are located <strong>in</strong> the cell body membrane. This excludes the most common source<br />

of <strong>in</strong>accuracies <strong>in</strong> the Hodgk<strong>in</strong>-Huxley type of models, the errors <strong>in</strong> voltage dependent<br />

channel properties due to the poor space clamp. The photoreceptor axons may still<br />

possess different types of voltage-<strong>gated</strong> ion <strong>channels</strong> (Weckström et al. 1992). These<br />

<strong>channels</strong> could potentially affect the voltage signals generated <strong>in</strong> the cell body, but based<br />

on the experimental data and the model simulations this effect is very subtle and unlikely<br />

to affect the results of this study.<br />

The phototransduction and the concurrent generation of the currents due to the LIC<br />

were not explicitly modelled <strong>in</strong> this study, although experimental data needed for<br />

develop<strong>in</strong>g a detailed model has been accumulat<strong>in</strong>g dur<strong>in</strong>g the recent years (Hardie 2001,<br />

Hardie 2003). This restriction was kept <strong>in</strong> m<strong>in</strong>d while plann<strong>in</strong>g the simulations to avoid<br />

mis<strong>in</strong>terpretations caused by <strong>in</strong>adequate modell<strong>in</strong>g. In (III-IV) it was circumvented by<br />

reconstruct<strong>in</strong>g the LIC from <strong>in</strong> vivo voltage responses.


44<br />

6.2 K + <strong>channels</strong> <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong><br />

<strong>Drosophila</strong> <strong>photoreceptors</strong> can be assumed to be tuned dur<strong>in</strong>g evolution accord<strong>in</strong>g to<br />

their visual ecology by adjust<strong>in</strong>g the properties of the phototransduction and by match<strong>in</strong>g<br />

the photo-<strong>in</strong>sensitive membrane to the transduced currents with voltage-<strong>gated</strong> K +<br />

<strong>channels</strong> (Weckström & Laughl<strong>in</strong> 1995). The biophysical properties of these <strong>channels</strong><br />

determ<strong>in</strong>e their role <strong>in</strong> process<strong>in</strong>g the voltage signals. Shaker <strong>channels</strong> activate rapidly <strong>in</strong><br />

voltages near the rest<strong>in</strong>g potential, followed by the fast <strong>in</strong>activation. Therefore, Shaker<br />

<strong>channels</strong> selectively attenuate the transient signal components while amplify<strong>in</strong>g the<br />

slower ones, which effectively spreads responses over the available voltage range (I).<br />

Photoreceptors generate a ma<strong>in</strong>ta<strong>in</strong>ed depolarization <strong>in</strong> response to <strong>in</strong>creas<strong>in</strong>g light<br />

<strong>in</strong>tensity. Due to their voltage dependency (Hardie 1991a, Hevers & Hardie 1995), Shab<br />

<strong>channels</strong> open at high light <strong>in</strong>tensities, thereby attenuat<strong>in</strong>g the light-<strong>in</strong>duced<br />

depolarization and prevent<strong>in</strong>g response saturation. Activation of Shab <strong>channels</strong> also<br />

reduces the membrane time constant of the <strong>Drosophila</strong> <strong>photoreceptors</strong> allow<strong>in</strong>g them to<br />

encode faster events (IV).<br />

Further <strong>in</strong>sight <strong>in</strong>to the role of each type of ion channel <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong><br />

may come from identification of the genes encod<strong>in</strong>g fast delayed rectifier and novel K +<br />

<strong>channels</strong>. It is shown that the novel non-<strong>in</strong>activat<strong>in</strong>g K + <strong>channels</strong> and the fast delayed<br />

rectifier <strong>channels</strong> are not encoded by either Shab or Shaker (IV) and are likely to be<br />

formed from separate gene products, not heteromultimers (Covarrubias et al. 1991). The<br />

most likely candidates are the rema<strong>in</strong><strong>in</strong>g two members of the Shaker-like gene family <strong>in</strong><br />

<strong>Drosophila</strong>, Shal and Shaw (Butler et al. 1989, Wei et al. 1990). In vitro characterization<br />

of these <strong>channels</strong> suggests that Shaw encodes non-<strong>in</strong>activat<strong>in</strong>g potassium channel whilst<br />

Shal encodes a fast activat<strong>in</strong>g-fast <strong>in</strong>activat<strong>in</strong>g channel, which have similar k<strong>in</strong>etics to the<br />

novel conductance and the fast delayed rectifier, respectively (Butler et al. 1989, Wei et<br />

al. 1990).<br />

6.3 The robustness of neural cod<strong>in</strong>g <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong><br />

The regulatory changes <strong>in</strong> other ion <strong>channels</strong> or <strong>in</strong> the passive membrane properties<br />

follow<strong>in</strong>g from imposed changes <strong>in</strong> the channel composition may be required for normal<br />

neural function and could potentially ensure that neurons ma<strong>in</strong>ta<strong>in</strong> relatively robust<br />

response characteristics despite variable levels of channel expression (I, IV, (Goldman et<br />

al. 2001, Marder & Pr<strong>in</strong>z 2002, Niven 2004). The regulatory changes may also provide<br />

<strong>in</strong>sight to the function of the channel affected by the mutation. A decrease <strong>in</strong> the <strong>in</strong>put<br />

resistance of the Shaker mutant <strong>photoreceptors</strong> may compensate for the compressed<br />

voltage range follow<strong>in</strong>g from the mutation (I). Shab <strong>channels</strong> activate <strong>in</strong> depolarized<br />

voltages be<strong>in</strong>g dom<strong>in</strong>ant K + conductance <strong>in</strong> light. Reduced <strong>in</strong>put resistance of the Shab<br />

mutant <strong>photoreceptors</strong> ma<strong>in</strong>ta<strong>in</strong>ed a similar operat<strong>in</strong>g voltage, sensitivity and contrast<br />

cod<strong>in</strong>g <strong>in</strong> light as the WT <strong>photoreceptors</strong> (IV). Although these changes for both the Shab<br />

and Shaker mutant <strong>photoreceptors</strong> significantly improved their performance (I, IV),<br />

several clear disadvantages may follow from loss of the voltage-<strong>gated</strong> <strong>channels</strong>. Both<br />

Shaker and Shab mutant <strong>photoreceptors</strong> show a compressed voltage-range available for


45<br />

cod<strong>in</strong>g the variations <strong>in</strong> light <strong>in</strong>tensity. This may result <strong>in</strong> impaired performance <strong>in</strong> cod<strong>in</strong>g<br />

the natural sceneries, which conta<strong>in</strong> fast and large <strong>in</strong>tensity fluctuations (III) (van Hateren<br />

1997). The reduced <strong>in</strong>put resistance is also energetically more expensive, especially at<br />

rest, than the voltage-<strong>gated</strong> <strong>channels</strong>, which activate dur<strong>in</strong>g the signall<strong>in</strong>g events (II,<br />

(Laughl<strong>in</strong> et al. 1998, Laughl<strong>in</strong> 2001).<br />

What could be the nature of the regulatory mechanism decreas<strong>in</strong>g the <strong>in</strong>put resistance<br />

<strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong>? Regulation of the composition of ion <strong>channels</strong> of a neuron<br />

was orig<strong>in</strong>ally shown to be produced by learn<strong>in</strong>g tasks (Brons & Woody 1980, Alkon<br />

1984, Moyer et al. 1996, Antonov et al. 2001). Subsequent studies <strong>in</strong> both <strong>in</strong>vertebrate<br />

and vertebrate neurons have demonstrated activity-dependent regulation of their electrical<br />

properties tak<strong>in</strong>g place <strong>in</strong> hours to days after imposed changes <strong>in</strong> their activity<br />

(LeMasson et al. 1993, Turrigiano et al. 1994, Turrigiano et al. 1995, Desai et al. 1999,<br />

Golowasch et al. 1999, Aizenman et al. 2003). Persistent changes <strong>in</strong> electrical properties<br />

can also be <strong>in</strong>duced rapidly (Aizenman & L<strong>in</strong>den 2000) and <strong>in</strong> a graded manner for both<br />

<strong>in</strong>creases and decreases <strong>in</strong> activity (Egorov et al. 2002, Nelson et al. 2003). Recently, an<br />

activity-<strong>in</strong>dependent mechanism was reported to take place <strong>in</strong> the absence of significant<br />

changes <strong>in</strong> physiological activity (MacLean et al. 2003). Activity-dependent regulation is<br />

thought to be mediated by changes <strong>in</strong> gene transcription (MacLean et al. 2003), whereas<br />

the activity-<strong>in</strong>dependent regulation was suggested to be controlled by a<br />

posttranscriptional mechanism. The results of this study show regulatory changes <strong>in</strong><br />

Shab 3 <strong>photoreceptors</strong> (IV), a null mutation of the Shab <strong>channels</strong> (Hegde et al. 1999);<br />

Shab 1 and Shab 2 <strong>photoreceptors</strong> (IV), missense mutations of the Shab <strong>channels</strong> (Hegde et<br />

al. 1999); as well as <strong>in</strong> the Shaker 14 <strong>photoreceptors</strong> (I), a null mutation produc<strong>in</strong>g<br />

nonfunctional Shaker <strong>channels</strong> (Kaplan & Trout 1961, Salkoff & Wyman 1981). These<br />

results suggest that the change <strong>in</strong> ion channel function would be sufficient to cause<br />

concurrent regulation of the electrical properties. Therefore, by us<strong>in</strong>g the established<br />

term<strong>in</strong>ology, the regulation is activity-dependent <strong>in</strong> the <strong>Drosophila</strong> <strong>photoreceptors</strong>.<br />

What could be the possible neuronal mechanism beh<strong>in</strong>d the decrease <strong>in</strong> the <strong>in</strong>put<br />

resistance? The decrease <strong>in</strong> the <strong>in</strong>put resistances was modelled by <strong>in</strong>creased leak<br />

conductances. In Hodgk<strong>in</strong>-Huxley type of models the leak conductances are rout<strong>in</strong>ely<br />

used to cover all the undef<strong>in</strong>ed processes to fit these models to the experimental data.<br />

Thus, leak conductances can be considered as a descriptive representation of several<br />

possible neuronal mechanisms, although a recent study demonstrated a ma<strong>in</strong>tenance of<br />

the electrical properties by the compensatory <strong>in</strong>crease <strong>in</strong> K + leak conductance <strong>in</strong><br />

celebellar granule cells after genetically remov<strong>in</strong>g a tonic synaptic conductance (Brickley<br />

et al. 2001). If the leak conductances used <strong>in</strong> the modell<strong>in</strong>g have cellular counterparts or<br />

if they describe a more complicated process, rema<strong>in</strong>s to be determ<strong>in</strong>ed <strong>in</strong> the future<br />

works.<br />

How could the regulation of the electrical properties be signalled <strong>in</strong> the <strong>Drosophila</strong><br />

<strong>photoreceptors</strong>? Dur<strong>in</strong>g a neuron’s development, a sequential acquisition of the<br />

expression of different ion channel types may depend on the <strong>in</strong>tracellular Ca 2+ that is<br />

associated with spontaneous activity (Gu & Spitzer 1997). Substantial theoretical and<br />

experimental evidence of Ca 2+ dependent regulation also <strong>in</strong> mature neurons makes Ca 2+ a<br />

prime candidate for the <strong>in</strong>duction mechanism of the robustness <strong>in</strong> electrical properties<br />

(Zhang & L<strong>in</strong>den 2003). In <strong>Drosophila</strong> <strong>photoreceptors</strong>, 40% of current generated by the<br />

LIC is carried by Ca 2+ ions lead<strong>in</strong>g to concentration changes <strong>in</strong> the cell body (Reuss et al.


46<br />

1997, Postma et al. 1999, Hardie 2001). The logarithm of <strong>in</strong>crease <strong>in</strong> cytosolic Ca 2+<br />

concentration <strong>in</strong> the cell body has been shown to be l<strong>in</strong>ear with the logarithm of the light<br />

<strong>in</strong>tensity <strong>in</strong> vivo <strong>in</strong> Calliphora vic<strong>in</strong>a <strong>photoreceptors</strong> (Oberw<strong>in</strong>kler & Stavenga 1998) and<br />

similar concentrations have also been determ<strong>in</strong>ed <strong>in</strong> vitro <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong><br />

(Hardie 1996). The changes <strong>in</strong> the K + conductance without a concurrent decrease <strong>in</strong> the<br />

<strong>in</strong>put resistance would alter the light-voltage relationship, potentially lead<strong>in</strong>g to changes<br />

<strong>in</strong> the Ca 2+ concentration (I, IV). Changes <strong>in</strong> Ca 2+ concentrations could modify the<br />

voltage-dependent properties of the ion <strong>channels</strong> through phosphorylation or direct<br />

modulation by the calcium (Zhang & L<strong>in</strong>den 2003), or it could regulate their expression.<br />

Therefore, Ca 2+ is suggested to play an essential role <strong>in</strong> the regulatory signall<strong>in</strong>g of the<br />

electrical properties.<br />

6.4 Naturalistic stimulus<br />

Natural visual <strong>in</strong>puts often conta<strong>in</strong> strong temporal and spatial correlations, as well as<br />

large <strong>in</strong>tensity fluctuations (Burton & Moorhead 1981, Sr<strong>in</strong>ivasan et al. 1982, Field 1987,<br />

Tolhurst et al. 1992, Ruderman & Bialek 1994) that may cause sparse, brief, <strong>in</strong>tense<br />

fir<strong>in</strong>g events (V<strong>in</strong>je & Gallant 2000, Re<strong>in</strong>agel 2001, V<strong>in</strong>je & Gallant 2002), and activate<br />

components such as ion <strong>channels</strong> that strongly <strong>in</strong>fluence neural cod<strong>in</strong>g. However, sensory<br />

systems are usually <strong>in</strong>vesti<strong>gated</strong> us<strong>in</strong>g simpler, easily characterised stimuli, such as<br />

pulses, steps or white noise (Zettler 1969, Laughl<strong>in</strong> 1978, French 1980, Juusola et al.<br />

1994, Juusola & Hardie 2001a). These experiments have yielded considerable <strong>in</strong>sights<br />

<strong>in</strong>to sensory systems but are unlikely to predict responses to natural stimuli.<br />

Nonl<strong>in</strong>ear analysis of the voltage responses of WT and Shaker mutant <strong>photoreceptors</strong><br />

to the naturalistic stimuli produced results that differed from those for white noise stimuli<br />

published previously (I), (Juusola et al. 2003), suggest<strong>in</strong>g that Shaker <strong>channels</strong> behave<br />

differently under different stimulus regimes (Fig. 3, III). Differences between kernels<br />

derived from naturalistic and white noise stimuli are due to the stimulus structure.<br />

Naturalistic stimuli can conta<strong>in</strong> prolonged dark periods <strong>in</strong>terspersed with periods of high<br />

light <strong>in</strong>tensity, evok<strong>in</strong>g large fluctuations <strong>in</strong> photoreceptor voltage (van Hateren 1997),<br />

whereas dark and bright periods <strong>in</strong> band-limited Gaussian white noise stimuli of similar<br />

duration are typically brief, allow<strong>in</strong>g photoreceptor responses to be modulated around a<br />

relatively constant mean voltage. Shaker <strong>channels</strong> behave differently under these two<br />

stimulation regimes because their prom<strong>in</strong>ent activation and <strong>in</strong>activation dur<strong>in</strong>g voltage<br />

transients are reduced by ma<strong>in</strong>ta<strong>in</strong>ed depolarization.<br />

The results <strong>in</strong> (III) suggest that mutation of the Shaker <strong>channels</strong> might affect the<br />

currents generated by the LIC through a feedback mechanism. Such feedback could occur<br />

due to changes <strong>in</strong> Ca 2+ concentration as discussed <strong>in</strong> 6.3. However, bump generation was<br />

found to be unaffected <strong>in</strong> the Shaker 14 <strong>photoreceptors</strong>. Individual bumps are responses to<br />

s<strong>in</strong>gle photons, whereas the nonl<strong>in</strong>ear analysis and modell<strong>in</strong>g <strong>in</strong> (III) was done for the<br />

responses to ~10 6 photons/s or more (III). As mean light <strong>in</strong>tensity <strong>in</strong>creases, average<br />

bump size decreases markedly and time course is reduced (Wu & Pak 1975, Juusola &<br />

Hardie 2001a). These changes <strong>in</strong> bump waveform may be due to Ca 2+ mediated<br />

adaptation (Henderson et al. 2000), which is a feature of many <strong>in</strong>vertebrate and


47<br />

vertebrate <strong>photoreceptors</strong> (Hardie & M<strong>in</strong>ke 1995, Montell 1999, Pugh et al. 1999, Burns<br />

& Baylor 2001, Fa<strong>in</strong> et al. 2001). Changes <strong>in</strong> light and dark adaptation mediated by Ca 2+<br />

have been proposed to expla<strong>in</strong> differences between photoreceptor responses with<strong>in</strong> an<br />

<strong>in</strong>dividual compound eye (Burton et al. 2001). Therefore, the changes <strong>in</strong> the currents<br />

generated by the LIC could possibly follow from the altered dynamics <strong>in</strong> light adaptation.<br />

This would extend previous f<strong>in</strong>d<strong>in</strong>gs that activity of the phototransduction cascade may<br />

alter photoreceptor membrane properties (Burton 2002, Wolfram & Juusola 2004) by<br />

show<strong>in</strong>g that these <strong>in</strong>teractions may be bidirectional.<br />

As <strong>in</strong> the Shaker <strong>photoreceptors</strong>, Shab mutant <strong>photoreceptors</strong> may also respond<br />

differently <strong>in</strong> the naturalistic stimuli regime. Indeed, the faster and larger impulse<br />

responses <strong>in</strong> light adapted Shab 3 mutants could provide an explanation for the <strong>in</strong>creased<br />

high frequency content <strong>in</strong> the contrast ga<strong>in</strong> (Fig 6D, IV). However, an <strong>in</strong>-depth study of a<br />

possible tun<strong>in</strong>g of the LIC properties after imposed changes <strong>in</strong> the photo-<strong>in</strong>sensitive<br />

membrane is needed to test the hypothesis and it rema<strong>in</strong>s to be <strong>in</strong>vesti<strong>gated</strong> <strong>in</strong> future<br />

works.<br />

6.5 Efficiency of neural cod<strong>in</strong>g<br />

The efficient expenditure of a limited energy budget is a major factor <strong>in</strong>fluenc<strong>in</strong>g all<br />

aspects of animal design, therefore, energy is thought to have been an important<br />

constra<strong>in</strong>t upon the evolution of nervous systems (Laughl<strong>in</strong> 2001), <strong>in</strong>clud<strong>in</strong>g that of<br />

humans (Aiello & Wheeler 1995). Potentially, energetic costs could limit absolute<br />

numbers of neurones and synaptic connections (<strong>in</strong>formation <strong>channels</strong>) <strong>in</strong> the bra<strong>in</strong><br />

(Laughl<strong>in</strong> et al. 1998), and <strong>in</strong>fluence the expression and distribution of ion <strong>channels</strong> and<br />

second messenger pathways <strong>in</strong> <strong>in</strong>dividual neurones. Indeed, the human bra<strong>in</strong> utilizes<br />

energy at close to 20 Watts, account<strong>in</strong>g for 20% of the rest<strong>in</strong>g oxygen consumption of the<br />

human body (Kety 1957), of which over half is required to fuel ion pumps (Ames 1997).<br />

The success of an <strong>in</strong>dividual also depends upon accurately determ<strong>in</strong><strong>in</strong>g and execut<strong>in</strong>g<br />

appropriate behaviour. Nervous systems are, therefore, also constra<strong>in</strong>ed by the need to<br />

extract biologically germane, reliable <strong>in</strong>formation from the environment to generate such<br />

adaptive behaviour (Meister & Berry 1999, de Polavieja 2002). The need for reliability,<br />

<strong>in</strong> the presence of noise, constra<strong>in</strong>s <strong>in</strong>formation process<strong>in</strong>g <strong>in</strong> nervous systems sett<strong>in</strong>g<br />

lower limits for the numbers of neurones and synaptic connections (Laughl<strong>in</strong> et al. 1998)<br />

and <strong>in</strong>fluenc<strong>in</strong>g neuronal design. Although metabolic energy is clearly important <strong>in</strong><br />

determ<strong>in</strong><strong>in</strong>g neural function, the basic data on the quantitative relationships between<br />

energy and <strong>in</strong>formation is lack<strong>in</strong>g for the nervous system. Precisely how much energy<br />

must a neuron consume to do a given amount of useful work, transmitt<strong>in</strong>g and process<strong>in</strong>g<br />

<strong>in</strong>formation? How does energy consumption scale with the quantity of <strong>in</strong>formation that<br />

neurons handle?<br />

Metabolic efficiency will only be an important determ<strong>in</strong>ant of the evolution and design<br />

of a signall<strong>in</strong>g system when metabolic costs impose a significant penalty on the parent<br />

organism. The estimates of photoreceptor ATP consumption <strong>in</strong> this study compares with a<br />

figure for the human bra<strong>in</strong> that is widely considered to be significant and could have<br />

shaped its evolution (Kety 1957, Laughl<strong>in</strong> 2001). The significance of ret<strong>in</strong>al ATP


48<br />

consumption is re<strong>in</strong>forced by comparative studies. Calliphora vic<strong>in</strong>a <strong>photoreceptors</strong><br />

consume more energy, gram for gram, than active mammalian muscle (Laughl<strong>in</strong> et al.<br />

1998).<br />

Understand<strong>in</strong>g the relationship between <strong>in</strong>dividual conductances and the metabolic<br />

cost of <strong>in</strong>formation may also elucidate the function of other components <strong>in</strong> neural<br />

signall<strong>in</strong>g such as neuromodulators. For example, seroton<strong>in</strong> positively shifts the voltage<br />

range of Shaker and Shab <strong>channels</strong> <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong> so that they are<br />

completely <strong>in</strong>active at rest (Hevers & Hardie 1995). This would remove a large<br />

proportion of the K + conductance <strong>in</strong> the WT <strong>photoreceptors</strong> <strong>in</strong> the dark, significantly<br />

reduc<strong>in</strong>g the metabolic cost of ma<strong>in</strong>ta<strong>in</strong><strong>in</strong>g the rest<strong>in</strong>g potential. However, the shifted<br />

voltage range would <strong>in</strong>crease the metabolic cost of <strong>in</strong>formation at high light <strong>in</strong>tensities,<br />

suggest<strong>in</strong>g seroton<strong>in</strong> should be released at night to m<strong>in</strong>imize metabolic cost. Thus,<br />

metabolic cost may be an important factor not only <strong>in</strong> determ<strong>in</strong><strong>in</strong>g which comb<strong>in</strong>ations of<br />

ion <strong>channels</strong> are expressed <strong>in</strong> particular neurones but also <strong>in</strong> the plasticity with<strong>in</strong> the<br />

properties of these ion <strong>channels</strong>.<br />

6.6 Relevance of the f<strong>in</strong>d<strong>in</strong>gs<br />

This study is a systematic survey on how voltage-<strong>gated</strong> K + <strong>channels</strong> contribute to the<br />

efficiency of neural process<strong>in</strong>g <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong>. The functions discovered<br />

are likely to be widespread <strong>in</strong> process<strong>in</strong>g graded signals <strong>in</strong> sensory receptors and post<br />

synaptic potentials <strong>in</strong> both vertebrates (Rudy 1988, Sheng et al. 1993, Wang et al. 1993,<br />

Debanne et al. 1997, Hoffman et al. 1997, Magee et al. 1998, Coetzee et al. 1999) and<br />

<strong>in</strong>vertebrates (Connor & Stevens 1971, Laurent 1990, Hardie 1991a, Weckström &<br />

Laughl<strong>in</strong> 1995).<br />

This study also shows how regulatory changes <strong>in</strong> the electrical properties can restore<br />

certa<strong>in</strong> aspects of function after imposed changes to the ion channel composition. One<br />

possible objective for these changes may be to ma<strong>in</strong>ta<strong>in</strong> the operat<strong>in</strong>g po<strong>in</strong>t of the<br />

<strong>Drosophila</strong> <strong>photoreceptors</strong> (I, IV). This would be well <strong>in</strong> accordance with the hypothesis<br />

that regulation of the electrical properties could maximize <strong>in</strong>formation storage by keep<strong>in</strong>g<br />

the dynamic range of signall<strong>in</strong>g with<strong>in</strong> useful limits (Stemmler & Koch 1999, Turrigiano<br />

& Nelson 2004).<br />

Results of this study demonstrate the importance of the voltage-<strong>gated</strong> K + <strong>channels</strong> for<br />

neural cod<strong>in</strong>g precision and highlight the effect of regulatory changes <strong>in</strong> the electrical<br />

properties upon neuronal <strong>in</strong>formation process<strong>in</strong>g. I believe this study improves our<br />

understand<strong>in</strong>g of the role of K + <strong>channels</strong> <strong>in</strong> neuronal signall<strong>in</strong>g and provides new <strong>in</strong>sight<br />

on how the remarkable robustness of neuronal systems may be achieved.


7 Conclusions<br />

1. Biophysical properties of the voltage-<strong>gated</strong> K + <strong>channels</strong> determ<strong>in</strong>e their role <strong>in</strong><br />

shap<strong>in</strong>g the voltage signals <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong>. Shaker <strong>channels</strong> attenuate<br />

the fast transients and amplify the slower signal components spread<strong>in</strong>g the signals<br />

efficiently over the available voltage range (I).<br />

2. Shab <strong>channels</strong> activate at high light <strong>in</strong>tensities, thereby attenuat<strong>in</strong>g the light-<strong>in</strong>duced<br />

depolarization and prevent<strong>in</strong>g response saturation. Activation of Shab <strong>channels</strong> also<br />

reduces the membrane time constant of <strong>Drosophila</strong> <strong>photoreceptors</strong> allow<strong>in</strong>g them to<br />

encode faster events (IV).<br />

3. Nonl<strong>in</strong>ear analysis and the modell<strong>in</strong>g of the WT and Shaker mutant <strong>photoreceptors</strong><br />

showed that, <strong>in</strong> addition to elim<strong>in</strong>at<strong>in</strong>g the Shaker current, the mutation also changed<br />

the current flow<strong>in</strong>g through the LIC and <strong>in</strong>creased the Shab current. Part of these<br />

changes could be attributed to direct feedback from the Shaker <strong>channels</strong> via the<br />

membrane potential. However, other changes may occur <strong>in</strong> the LIC of Shaker mutants,<br />

possibly dur<strong>in</strong>g photoreceptor development (III).<br />

4. Comparison of the Shaker and Shab mutant <strong>photoreceptors</strong> with the WT revealed that<br />

a concurrent decrease <strong>in</strong> the steady-state <strong>in</strong>put resistance followed from deletion of the<br />

voltage-<strong>gated</strong> K + <strong>channels</strong>. This allowed partial compensation of the compression and<br />

saturation caused by the loss of Shaker <strong>channels</strong> and it ma<strong>in</strong>ta<strong>in</strong>ed the characteristics<br />

of the light-voltage relationship <strong>in</strong> Shab mutant <strong>photoreceptors</strong> (I&IV).<br />

5. Although the regulation of the electrical properties provided robustness to the<br />

performance of the <strong>Drosophila</strong> <strong>photoreceptors</strong>, it did not fully restore wild type<br />

properties <strong>in</strong> either mutant (I,III,IV). Indeed, it resulted <strong>in</strong> decreased efficacy of neural<br />

process<strong>in</strong>g, assessed by the metabolic cost of <strong>in</strong>formation (II).


References<br />

Aiello LC & Wheeler P (1995) The Expensive-Tissue Hypothesis - the Bra<strong>in</strong> and the Digestive-<br />

System <strong>in</strong> Human and Primate Evolution. Current Anthropology 36(2): 199-221.<br />

Aizenman CD, Akerman CJ, Jensen KR & Cl<strong>in</strong>e HT (2003) Visually driven regulation of <strong>in</strong>tr<strong>in</strong>sic<br />

neuronal excitability improves stimulus detection <strong>in</strong> vivo. Neuron 39(5): 831-42.<br />

Aizenman CD & L<strong>in</strong>den DJ (2000) Rapid, synaptically driven <strong>in</strong>creases <strong>in</strong> the <strong>in</strong>tr<strong>in</strong>sic excitability<br />

of cerebellar deep nuclear neurons. Nat Neurosci 3(2): 109-11.<br />

Alkon DL (1984) Calcium-mediated reduction of ionic currents: a biophysical memory trace.<br />

Science 226(4678): 1037-45.<br />

Ames A (1997) Energy requirements of bra<strong>in</strong> function: when is energy limit<strong>in</strong>g? In: I. Bodis-<br />

Wollner (ed) Mitochondria and free radicals <strong>in</strong> neurodegenerative diseases. New York, Wiley-<br />

Liss, p 17-27.<br />

Antonov I, Antonova I, Kandel ER & Hawk<strong>in</strong>s RD (2001) The contribution of activity-dependent<br />

synaptic plasticity to classical condition<strong>in</strong>g <strong>in</strong> Aplysia. J Neurosci 21(16): 6413-22.<br />

Bernste<strong>in</strong> J (1902) Untersuchungen zur Thermodynamik der bioelektrischen Ströme. Pflügers Arch<br />

92: 521-562.<br />

Brickley SG, Revilla V, Cull-Candy SG, Wisden W & Farrant M (2001) Adaptive regulation of<br />

neuronal excitability by a voltage-<strong>in</strong>dependent potassium conductance. Nature 409(6816): 88-<br />

92.<br />

Brons JF & Woody CD (1980) Long-term changes <strong>in</strong> excitability of cortical neurons after<br />

Pavlovian condition<strong>in</strong>g and ext<strong>in</strong>ction. J Neurophysiol 44(3): 605-15.<br />

Burns ME & Baylor DA (2001) Activation, deactivation, and adaptation <strong>in</strong> vertebrate<br />

photoreceptor cells. Annu Rev Neurosci 24: 779-805.<br />

Burton BG (2002) Long-term light adaptation <strong>in</strong> <strong>photoreceptors</strong> of the housefly, Musca domestica.<br />

J Comp Physiol A Neuroethol Sens Neural Behav Physiol 188(7): 527-38.<br />

Burton BG & Laughl<strong>in</strong> SB (2003) Neural images of pursuit targets <strong>in</strong> the photoreceptor arrays of<br />

male and female houseflies Musca domestica. J Exp Biol 206(Pt 22): 3963-77.<br />

Burton BG, Tatler BW & Laughl<strong>in</strong> SB (2001) Variations <strong>in</strong> photoreceptor response dynamics<br />

across the fly ret<strong>in</strong>a. J Neurophysiol 86(2): 950-60.<br />

Burton GJ & Moorhead IR (1981) Visual form perception and the spatial phase transfer function. J<br />

Opt Soc Am 71(9): 1056-63.<br />

Butler A, Wei AG, Baker K & Salkoff L (1989) A family of putative potassium channel genes <strong>in</strong><br />

<strong>Drosophila</strong>. Science 243(4893): 943-7.<br />

Chacron MJ, Doiron B, Maler L, Longt<strong>in</strong> A & Bastian J (2003) Non-classical receptive field<br />

mediates switch <strong>in</strong> a sensory neuron's frequency tun<strong>in</strong>g. Nature 423(6935): 77-81.


51<br />

Coetzee WA, Amarillo Y, Chiu J, Chow A, Lau D, McCormack T, Moreno H, Nadal MS, Ozaita<br />

A, Pountney D, Saganich M, Vega-Saenz de Miera E & Rudy B (1999) Molecular diversity of<br />

<strong>K+</strong> <strong>channels</strong>. Ann N Y Acad Sci 868: 233-85.<br />

Connor JA & Stevens CF (1971) <strong>Voltage</strong> clamp studies of a transient outward membrane current <strong>in</strong><br />

gastropod neural somata. J Physiol 213(1): 21-30.<br />

Covarrubias M, Wei AA & Salkoff L (1991) Shaker, Shal, Shab, and Shaw express <strong>in</strong>dependent <strong>K+</strong><br />

current systems. Neuron 7(5): 763-73.<br />

Dan Y, Atick JJ & Reid RC (1996) Efficient cod<strong>in</strong>g of natural scenes <strong>in</strong> the lateral geniculate<br />

nucleus: experimental test of a computational theory. J Neurosci 16(10): 3351-62.<br />

Dayan P & Abbott LF (2001) Theoretical neuroscience : computational and mathematical model<strong>in</strong>g<br />

of neural systems. MIT Press, Cambridge, Mass.<br />

de Polavieja GG (2002) Errors drive the evolution of biological signall<strong>in</strong>g to costly codes. J Theor<br />

Biol 214(4): 657-64.<br />

Debanne D, Guer<strong>in</strong>eau NC, Gahwiler BH & Thompson SM (1997) Action-potential propagation<br />

<strong>gated</strong> by an axonal I(A)-like <strong>K+</strong> conductance <strong>in</strong> hippocampus. Nature 389(6648): 286-9.<br />

Delpire E (2000) Cation-Chloride Cotransporters <strong>in</strong> Neuronal Communication. News Physiol Sci<br />

15: 309-312.<br />

Desai NS, Rutherford LC & Turrigiano GG (1999) Plasticity <strong>in</strong> the <strong>in</strong>tr<strong>in</strong>sic excitability of cortical<br />

pyramidal neurons. Nat Neurosci 2(6): 515-20.<br />

Egorov AV, Hamam BN, Fransen E, Hasselmo ME & Alonso AA (2002) Graded persistent activity<br />

<strong>in</strong> entorh<strong>in</strong>al cortex neurons. Nature 420(6912): 173-8.<br />

Fa<strong>in</strong> GL, Matthews HR, Cornwall MC & Koutalos Y (2001) Adaptation <strong>in</strong> vertebrate<br />

<strong>photoreceptors</strong>. Physiol Rev 81(1): 117-151.<br />

Field DJ (1987) Relations between the statistics of natural images and the response properties of<br />

cortical cells. J Opt Soc Am A 4(12): 2379-94.<br />

Francesch<strong>in</strong>i N & Kirschfeld K (1971) Etude optique <strong>in</strong> vivo des éléments photorecepteurs dans<br />

l'oeil composé de <strong>Drosophila</strong>. Kybernetik 9(5): 159-82.<br />

French AS (1980) Phototransduction <strong>in</strong> the fly compound eye exhibits temporal resonances and a<br />

pure time delay. Nature 283(5743): 200-2.<br />

French AS & Korenberg MJ (1989) A nonl<strong>in</strong>ear cascade model for action potential encod<strong>in</strong>g <strong>in</strong> an<br />

<strong>in</strong>sect sensory neuron. Biophys J 55(4): 655-61.<br />

French AS, Korenberg MJ, Järvilehto M, Kouvala<strong>in</strong>en E, Juusola M & Weckström M (1993) The<br />

dynamic nonl<strong>in</strong>ear behavior of fly <strong>photoreceptors</strong> evoked by a wide range of light <strong>in</strong>tensities.<br />

Biophys J 65(2): 832-9.<br />

French AS & Marmarelis VZ (1999) Nonl<strong>in</strong>ear analysis of neuronal systems. In: H. Johansson (ed)<br />

Modern nechniques <strong>in</strong> neuroscience research. Berl<strong>in</strong>, Spr<strong>in</strong>ger-Verlag, p 627-640.<br />

French AS, Sekizawa SI, Hoger U & Torkkeli PH (2001) Predict<strong>in</strong>g the responses of<br />

mechanoreceptor neurons to physiological <strong>in</strong>puts by nonl<strong>in</strong>ear system identification. Ann<br />

Biomed Eng 29(3): 187-94.<br />

Gerster U (1997) A quantitative estimate of flash-<strong>in</strong>duced Ca(2+)- and Na(+)-<strong>in</strong>flux and<br />

Na+/Ca(2+)-exchange <strong>in</strong> blowfly Calliphora <strong>photoreceptors</strong>. Vision Res 37(18): 2477-85.<br />

Gerster U, Stavenga DG & Backhaus W (1997) Na+/K(+)pump activity <strong>in</strong> <strong>photoreceptors</strong> of the<br />

blowfly Calliphora: A model analysis based on membrane potential measurements. Journal of<br />

Comparative Physiology a-Sensory Neural and Behavioral Physiology 180(2): 113-122.<br />

Goldman MS, Golowasch J, Marder E & Abbott LF (2001) Global structure, robustness, and<br />

modulation of neuronal models. J Neurosci 21(14): 5229-38.<br />

Golowasch J, Abbott LF & Marder E (1999) Activity-dependent regulation of potassium currents <strong>in</strong><br />

an identified neuron of the stomatogastric ganglion of the crab Cancer borealis. J Neurosci<br />

19(20): RC33.


52<br />

Gu X & Spitzer NC (1997) Break<strong>in</strong>g the code: regulation of neuronal differentiation by<br />

spontaneous calcium transients. Dev Neurosci 19(1): 33-41.<br />

Hamdorf K, Hochstrate P, Hoglund G, Burbach B & Wiegand U (1988) Light activation of the<br />

sodium-pump <strong>in</strong> blowfly <strong>photoreceptors</strong>. J Comp Physiol A Neuroethol Sens Neural Behav<br />

Physiol 162: 285-300.<br />

Hardie RC (1991a) <strong>Voltage</strong>-sensitive potassium <strong>channels</strong> <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong>. J Neurosci<br />

11(10): 3079-95.<br />

Hardie RC (1991b) Whole-Cell Record<strong>in</strong>gs of the Light-Induced Current <strong>in</strong> Dissociated<br />

<strong>Drosophila</strong> Photoreceptors - Evidence for Feedback by Calcium Permeat<strong>in</strong>g the Light-Sensitive<br />

Channels. Proceed<strong>in</strong>gs of the Royal Society of London Series B-Biological Sciences 245(1314):<br />

203-210.<br />

Hardie RC (1996) INDO-1 measurements of absolute rest<strong>in</strong>g and light-<strong>in</strong>duced Ca2+ concentration<br />

<strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong>. J Neurosci 16(9): 2924-33.<br />

Hardie RC (2001) Phototransduction <strong>in</strong> <strong>Drosophila</strong> melanogaster. J Exp Biol 204(Pt 20): 3403-9.<br />

Hardie RC (2003) Regulation of TRP <strong>channels</strong> via lipid second messengers. Annu Rev Physiol 65:<br />

735-59.<br />

Hardie RC & M<strong>in</strong>ke B (1995) Phospho<strong>in</strong>ositide-mediated phototransduction <strong>in</strong> <strong>Drosophila</strong><br />

<strong>photoreceptors</strong>: the role of Ca2+ and trp. Cell Calcium 18(4): 256-74.<br />

Hardie RC & Raghu P (2001) Visual transduction <strong>in</strong> <strong>Drosophila</strong>. Nature 413(6852): 186-93.<br />

Hardie RC, Raghu P, Moore S, Juusola M, Ba<strong>in</strong>es RA & Sweeney ST (2001) Calcium <strong>in</strong>flux via<br />

TRP <strong>channels</strong> is required to ma<strong>in</strong>ta<strong>in</strong> PIP2 levels <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong>. Neuron 30(1):<br />

149-59.<br />

Hardie RC, Voss D, Pongs O & Laughl<strong>in</strong> SB (1991) Novel potassium <strong>channels</strong> encoded by the<br />

Shaker locus <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong>. Neuron 6(3): 477-86.<br />

Hegde P, Gu GG, Chen D, Free SJ & S<strong>in</strong>gh S (1999) Mutational analysis of the Shab-encoded<br />

delayed rectifier K(+) <strong>channels</strong> <strong>in</strong> <strong>Drosophila</strong>. J Biol Chem 274(31): 22109-13.<br />

Henderson SR, Reuss H & Hardie RC (2000) S<strong>in</strong>gle photon responses <strong>in</strong> <strong>Drosophila</strong><br />

<strong>photoreceptors</strong> and their regulation by Ca2+. J Physiol 524 Pt 1: 179-94.<br />

Hevers W & Hardie RC (1995) Seroton<strong>in</strong> modulates the voltage dependence of delayed rectifier<br />

and Shaker potassium <strong>channels</strong> <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong>. Neuron 14(4): 845-56.<br />

Hille B (2001) Ion <strong>channels</strong> of excitable membranes. S<strong>in</strong>auer, Sunderland, Mass.<br />

Hodgk<strong>in</strong> AL & Huxley AF (1952) A quantitative description of membrane current and its<br />

application to conduction and excitation <strong>in</strong> nerve. J Physiol 117(4): 500-44.<br />

Hoffman DA & Johnston D (1998) Downregulation of transient <strong>K+</strong> <strong>channels</strong> <strong>in</strong> dendrites of<br />

hippocampal CA1 pyramidal neurons by activation of PKA and PKC. J Neurosci 18(10): 3521-<br />

8.<br />

Hoffman DA, Magee JC, Colbert CM & Johnston D (1997) <strong>K+</strong> channel regulation of signal<br />

propagation <strong>in</strong> dendrites of hippocampal pyramidal neurons. Nature 387(6636): 869-75.<br />

Isacoff EY, Jan YN & Jan LY (1990) Evidence for the formation of heteromultimeric potassium<br />

<strong>channels</strong> <strong>in</strong> Xenopus oocytes. Nature 345(6275): 530-4.<br />

Iverson LE, Tanouye MA, Lester HA, Davidson N & Rudy B (1988) A-type potassium <strong>channels</strong><br />

expressed from Shaker locus cDNA. Proc Natl Acad Sci U S A 85(15): 5723-7.<br />

Johnston D, Christie BR, Frick A, Gray R, Hoffman DA, Schexnayder LK, Watanabe S & Yuan<br />

LL (2003) Active dendrites, potassium <strong>channels</strong> and synaptic plasticity. Philos Trans R Soc<br />

Lond B Biol Sci 358(1432): 667-74.<br />

Juusola M & de Polavieja GG (2003) The rate of <strong>in</strong>formation transfer of naturalistic stimulation by<br />

graded potentials. J Gen Physiol 122(2): 191-206.<br />

Juusola M & Hardie RC (2001a) Light adaptation <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong>: I. Response<br />

dynamics and signal<strong>in</strong>g efficiency at 25 degrees C. J Gen Physiol 117(1): 3-25.


53<br />

Juusola M & Hardie RC (2001b) Light adaptation <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong>: II. Ris<strong>in</strong>g<br />

temperature <strong>in</strong>creases the bandwidth of reliable signal<strong>in</strong>g. J Gen Physiol 117(1): 27-42.<br />

Juusola M, Kouvala<strong>in</strong>en E, Järvilehto M & Weckström M (1994) Contrast ga<strong>in</strong>, signal-to-noise<br />

ratio, and l<strong>in</strong>earity <strong>in</strong> light-adapted blowfly <strong>photoreceptors</strong>. J Gen Physiol 104(3): 593-621.<br />

Juusola M, Niven JE & French AS (2003) Shaker <strong>K+</strong> <strong>channels</strong> contribute early nonl<strong>in</strong>ear<br />

amplification to the light response <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong>. J Neurophysiol 90(3): 2014-<br />

21.<br />

Kamb A, Tseng-Crank J & Tanouye MA (1988) Multiple products of the <strong>Drosophila</strong> Shaker gene<br />

may contribute to potassium channel diversity. Neuron 1(5): 421-30.<br />

Kaplan WD & Trout WE (1961) The behaviour of four neurological mutants of <strong>Drosophila</strong>.<br />

Genetics 61: 399-409.<br />

Katz LC & Shatz CJ (1996) Synaptic activity and the construction of cortical circuits. Science<br />

274(5290): 1133-8.<br />

Kety SS (1957) The general metabolism of the bra<strong>in</strong> <strong>in</strong> vivo. In: D. Richter (ed) Metabolism of the<br />

nervous system. London, Pergamon, p 221-237.<br />

Koch C (1999) Biophysics of computation : <strong>in</strong>formation process<strong>in</strong>g <strong>in</strong> s<strong>in</strong>gle neurons. Oxford<br />

University Press, New York.<br />

Korenberg MJ (1991) Parallel Cascade Identification and Kernel Estimation for Nonl<strong>in</strong>ear-<br />

Systems. Annals of Biomedical Eng<strong>in</strong>eer<strong>in</strong>g 19(4): 429-455.<br />

Korenberg MJ & Hunter IW (1986) The identification of nonl<strong>in</strong>ear biological systems: LNL<br />

cascade models. Biol Cybern 55(2-3): 125-34.<br />

Kouvala<strong>in</strong>en E, Weckström M & Juusola M (1994) A method for determ<strong>in</strong><strong>in</strong>g photoreceptor<br />

signal-to-noise ratio <strong>in</strong> the time and frequency doma<strong>in</strong>s with a pseudorandom stimulus. Vis<br />

Neurosci 11(6): 1221-5.<br />

Laughl<strong>in</strong> S (1981) A simple cod<strong>in</strong>g procedure enhances a neuron's <strong>in</strong>formation capacity. Z<br />

Naturforsch [C] 36(9-10): 910-2.<br />

Laughl<strong>in</strong> SB (2001) Energy as a constra<strong>in</strong>t on the cod<strong>in</strong>g and process<strong>in</strong>g of sensory <strong>in</strong>formation.<br />

Curr Op<strong>in</strong> Neurobiol 11(4): 475-80.<br />

Laughl<strong>in</strong> SB, de Ruyter van Steven<strong>in</strong>ck RR & Anderson JC (1998) The metabolic cost of neural<br />

<strong>in</strong>formation. Nat Neurosci 1(1): 36-41.<br />

Laughl<strong>in</strong> SB & Weckstrom M (1993) Fast and Slow Photoreceptors - a Comparative-Study of the<br />

Functional Diversity of Cod<strong>in</strong>g and Conductances <strong>in</strong> the Diptera. Journal of Comparative<br />

Physiology a-Sensory Neural and Behavioral Physiology 172(5): 593-609.<br />

Laughl<strong>in</strong> SBH, R. C. (1978) Common strategies for light adaptation <strong>in</strong> the peripheral systems of fly<br />

and dragonfly. Journal of Comparative Physiology a-Sensory Neural and Behavioral Physiology<br />

128: 319-340.<br />

Laurent G (1990) <strong>Voltage</strong>-dependent nonl<strong>in</strong>earities <strong>in</strong> the membrane of locust nonspik<strong>in</strong>g local<br />

<strong>in</strong>terneurons, and their significance for synaptic <strong>in</strong>tegration. J Neurosci 10(7): 2268-80.<br />

LeMasson G, Marder E & Abbott LF (1993) Activity-dependent regulation of conductances <strong>in</strong><br />

model neurons. Science 259(5103): 1915-7.<br />

Lewen GD, Bialek W & de Ruyter van Steven<strong>in</strong>ck RR (2001) Neural cod<strong>in</strong>g of naturalistic motion<br />

stimuli. Network 12(3): 317-29.<br />

Liu Z, Golowasch J, Marder E & Abbott LF (1998) A model neuron with activity-dependent<br />

conductances regulated by multiple calcium sensors. J Neurosci 18(7): 2309-20.<br />

MacLean JN, Zhang Y, Johnson BR & Harris-Warrick RM (2003) Activity-<strong>in</strong>dependent<br />

homeostasis <strong>in</strong> rhythmically active neurons. Neuron 37(1): 109-20.<br />

Magee J, Hoffman D, Colbert C & Johnston D (1998) Electrical and calcium signal<strong>in</strong>g <strong>in</strong> dendrites<br />

of hippocampal pyramidal neurons. Annu Rev Physiol 60: 327-46.<br />

Marder E & Pr<strong>in</strong>z AA (2002) Model<strong>in</strong>g stability <strong>in</strong> neuron and network function: the role of<br />

activity <strong>in</strong> homeostasis. Bioessays 24(12): 1145-54.


54<br />

Marmarelis PZ & Marmarelis VZ (1978) Analysis of physiological systems : the white-noise<br />

approach. Plenum Press, New York.<br />

McCormack K, L<strong>in</strong> JW, Iverson LE & Rudy B (1990) Shaker <strong>K+</strong> channel subunits from<br />

heteromultimeric <strong>channels</strong> with novel functional properties. Biochem Biophys Res Commun<br />

171(3): 1361-71.<br />

Meister M & Berry MJ, 2nd (1999) The neural code of the ret<strong>in</strong>a. Neuron 22(3): 435-50.<br />

Meunier C & Segev I (2002) Play<strong>in</strong>g the devil's advocate: is the Hodgk<strong>in</strong>-Huxley model useful?<br />

Trends Neurosci 25(11): 558-63.<br />

Montell C (1999) Visual transduction <strong>in</strong> <strong>Drosophila</strong>. Annu Rev Cell Dev Biol 15: 231-68.<br />

Moyer JR, Jr., Thompson LT & Disterhoft JF (1996) Trace eyebl<strong>in</strong>k condition<strong>in</strong>g <strong>in</strong>creases CA1<br />

excitability <strong>in</strong> a transient and learn<strong>in</strong>g-specific manner. J Neurosci 16(17): 5536-46.<br />

Namkung Y, Smith SM, Lee SB, Skrypnyk NV, Kim HL, Ch<strong>in</strong> H, Scheller RH, Tsien RW & Sh<strong>in</strong><br />

HS (1998) Targeted disruption of the Ca2+ channel beta3 subunit reduces N- and L-type Ca2+<br />

channel activity and alters the voltage-dependent activation of P/Q-type Ca2+ <strong>channels</strong> <strong>in</strong><br />

neurons. Proc Natl Acad Sci U S A 95(20): 12010-5.<br />

Nelson AB, Krispel CM, Sekirnjak C & du Lac S (2003) Long-last<strong>in</strong>g <strong>in</strong>creases <strong>in</strong> <strong>in</strong>tr<strong>in</strong>sic<br />

excitability triggered by <strong>in</strong>hibition. Neuron 40(3): 609-20.<br />

Niven JE (2004) Channell<strong>in</strong>g Evolution: Canalization and the nervous system. PLoS Biol 2(1):<br />

E19.<br />

Niven JE & Burrows M (2003) Spike width reduction modifies the dynamics of short-term<br />

depression at a central synapse <strong>in</strong> the locust. J Neurosci 23(20): 7461-9.<br />

Oberw<strong>in</strong>kler J & Stavenga DG (1998) Light dependence of calcium and membrane potential<br />

measured <strong>in</strong> blowfly <strong>photoreceptors</strong> <strong>in</strong> vivo. J Gen Physiol 112(2): 113-24.<br />

Oberw<strong>in</strong>kler J & Stavenga DG (2000) Calcium transients <strong>in</strong> the rhabdomeres of dark- and lightadapted<br />

fly photoreceptor cells. J Neurosci 20(5): 1701-9.<br />

Pongs O, Kecskemethy N, Muller R, Krah-Jentgens I, Baumann A, Kiltz HH, Canal I, Llamazares<br />

S & Ferrus A (1988) Shaker encodes a family of putative potassium channel prote<strong>in</strong>s <strong>in</strong> the<br />

nervous system of <strong>Drosophila</strong>. Embo J 7(4): 1087-96.<br />

Postma M, Oberw<strong>in</strong>kler J & Stavenga DG (1999) Does Ca2+ reach millimolar concentrations after<br />

s<strong>in</strong>gle photon absorption <strong>in</strong> <strong>Drosophila</strong> photoreceptor microvilli? Biophys J 77(4): 1811-23.<br />

Pugh EN, Jr., Nikonov S & Lamb TD (1999) Molecular mechanisms of vertebrate photoreceptor<br />

light adaptation. Curr Op<strong>in</strong> Neurobiol 9(4): 410-8.<br />

Re<strong>in</strong>agel P (2001) How do visual neurons respond <strong>in</strong> the real world? Curr Op<strong>in</strong> Neurobiol 11(4):<br />

437-42.<br />

Reuss H, Mojet MH, Chyb S & Hardie RC (1997) In vivo analysis of the <strong>Drosophila</strong> light-sensitive<br />

<strong>channels</strong>, TRP and TRPL. Neuron 19(6): 1249-59.<br />

Rieke F (1997) Spikes : explor<strong>in</strong>g the neural code. MIT Press, Cambridge, Mass.<br />

Rieke F, Bodnar DA & Bialek W (1995) Naturalistic stimuli <strong>in</strong>crease the rate and efficiency of<br />

<strong>in</strong>formation transmission by primary auditory afferents. Proc R Soc Lond B Biol Sci 262(1365):<br />

259-65.<br />

R<strong>in</strong>berg D & Davidowitz H (2002) A stimulus generat<strong>in</strong>g system for study<strong>in</strong>g w<strong>in</strong>d sensation <strong>in</strong> the<br />

American cockroach. J Neurosci Methods 121(1): 1-11.<br />

Ruderman DL & Bialek W (1994) Statistics of natural images: Scal<strong>in</strong>g <strong>in</strong> the woods. Physical<br />

Review Letters 73(6): 814-817.<br />

Rudy B (1988) Diversity and ubiquity of K <strong>channels</strong>. Neuroscience 25(3): 729-49.<br />

Salkoff L & Wyman R (1981) Genetic modification of potassium <strong>channels</strong> <strong>in</strong> <strong>Drosophila</strong> Shaker<br />

mutants. Nature 293(5829): 228-30.<br />

Schwarz TL, Tempel BL, Papazian DM, Jan YN & Jan LY (1988) Multiple potassium-channel<br />

components are produced by alternative splic<strong>in</strong>g at the Shaker locus <strong>in</strong> <strong>Drosophila</strong>. Nature<br />

331(6152): 137-42.


55<br />

Shannon CE (1948) Communication <strong>in</strong> the precence of noise. Proc Inst Radio Eng 37: 10-21.<br />

Sheng M, Liao YJ, Jan YN & Jan LY (1993) Presynaptic A-current based on heteromultimeric <strong>K+</strong><br />

<strong>channels</strong> detected <strong>in</strong> vivo. Nature 365(6441): 72-5.<br />

Sr<strong>in</strong>ivasan MV, Laughl<strong>in</strong> SB & Dubs A (1982) Predictive cod<strong>in</strong>g: a fresh view of <strong>in</strong>hibition <strong>in</strong> the<br />

ret<strong>in</strong>a. Proc R Soc Lond B Biol Sci 216(1205): 427-59.<br />

Stavenga DG (2003a) Angular and spectral sensitivity of fly <strong>photoreceptors</strong>. I. Integrated facet lens<br />

and rhabdomere optics. Journal of Comparative Physiology a-Neuroethology Sensory Neural<br />

and Behavioral Physiology 189(1): 1-17.<br />

Stavenga DG (2003b) Angular and spectral sensitivity of fly <strong>photoreceptors</strong>. II. Dependence on<br />

facet lens F-number and rhabdomere type <strong>in</strong> <strong>Drosophila</strong>. J Comp Physiol A Neuroethol Sens<br />

Neural Behav Physiol 189(3): 189-202.<br />

Stavenga DG (2004) Angular and spectral sensitivity of fly <strong>photoreceptors</strong>. III. Dependence on the<br />

pupil mechanism <strong>in</strong> the blowfly Calliphora. J Comp Physiol A Neuroethol Sens Neural Behav<br />

Physiol 190(2): 115-29.<br />

Stemmler M & Koch C (1999) How voltage-dependent conductances can adapt to maximize the<br />

<strong>in</strong>formation encoded by neuronal fir<strong>in</strong>g rate. Nat Neurosci 2(6): 521-7.<br />

Tempel BL, Papazian DM, Schwarz TL, Jan YN & Jan LY (1987) Sequence of a probable<br />

potassium channel component encoded at Shaker locus of <strong>Drosophila</strong>. Science 237(4816): 770-<br />

5.<br />

Timpe LC, Schwarz TL, Tempel BL, Papazian DM, Jan YN & Jan LY (1988) Expression of<br />

functional potassium <strong>channels</strong> from Shaker cDNA <strong>in</strong> Xenopus oocytes. Nature 331(6152): 143-<br />

5.<br />

Tolhurst DJ, Tadmor Y & Chao T (1992) Amplitude spectra of natural images. Ophthalmic Physiol<br />

Opt 12(2): 229-32.<br />

Turrigiano G, Abbott LF & Marder E (1994) Activity-dependent changes <strong>in</strong> the <strong>in</strong>tr<strong>in</strong>sic properties<br />

of cultured neurons. Science 264(5161): 974-7.<br />

Turrigiano G, LeMasson G & Marder E (1995) Selective regulation of current densities underlies<br />

spontaneous changes <strong>in</strong> the activity of cultured neurons. J Neurosci 15(5 Pt 1): 3640-52.<br />

Turrigiano GG & Nelson SB (2004) Homeostatic plasticity <strong>in</strong> the develop<strong>in</strong>g nervous system. Nat<br />

Rev Neurosci 5(2): 97-107.<br />

van Hateren JH (1997) Process<strong>in</strong>g of natural time series of <strong>in</strong>tensities by the visual system of the<br />

blowfly. Vision Res 37(23): 3407-16.<br />

Vickers NJ, Christensen TA, Baker TC & Hildebrand JG (2001) Odour-plume dynamics <strong>in</strong>fluence<br />

the bra<strong>in</strong>'s olfactory code. Nature 410(6827): 466-70.<br />

V<strong>in</strong>je WE & Gallant JL (2000) Sparse cod<strong>in</strong>g and decorrelation <strong>in</strong> primary visual cortex dur<strong>in</strong>g<br />

natural vision. Science 287(5456): 1273-6.<br />

V<strong>in</strong>je WE & Gallant JL (2002) Natural stimulation of the nonclassical receptive field <strong>in</strong>creases<br />

<strong>in</strong>formation transmission efficiency <strong>in</strong> V1. J Neurosci 22(7): 2904-15.<br />

Wang H, Kunkel DD, Mart<strong>in</strong> TM, Schwartzkro<strong>in</strong> PA & Tempel BL (1993) Heteromultimeric <strong>K+</strong><br />

<strong>channels</strong> <strong>in</strong> term<strong>in</strong>al and juxtaparanodal regions of neurons. Nature 365(6441): 75-9.<br />

Weckström M, Hardie RC & Laughl<strong>in</strong> SB (1991) <strong>Voltage</strong>-activated potassium <strong>channels</strong> <strong>in</strong> blowfly<br />

<strong>photoreceptors</strong> and their role <strong>in</strong> light adaptation. J Physiol 440: 635-57.<br />

Weckström M, Juusola M & Laughl<strong>in</strong> SB (1992) Presynaptic Enhancement of Signal Transients <strong>in</strong><br />

Photoreceptor Term<strong>in</strong>als <strong>in</strong> the Compound Eye. Proceed<strong>in</strong>gs of the Royal Society of London<br />

Series B-Biological Sciences 250(1327): 83-89.<br />

Weckström M & Laughl<strong>in</strong> SB (1995) Visual Ecology and <strong>Voltage</strong>-Gated Ion Channels <strong>in</strong> Insect<br />

Photoreceptors. Trends <strong>in</strong> Neurosciences 18(1): 17-21.<br />

Wei A, Covarrubias M, Butler A, Baker K, Pak M & Salkoff L (1990) <strong>K+</strong> current diversity is<br />

produced by an extended gene family conserved <strong>in</strong> <strong>Drosophila</strong> and mouse. Science 248(4955):<br />

599-603.


56<br />

Wickman K, Nemec J, Gendler SJ & Clapham DE (1998) Abnormal heart rate regulation <strong>in</strong> GIRK4<br />

knockout mice. Neuron 20(1): 103-14.<br />

Wolfram V & Juusola M (2004) Impact of rear<strong>in</strong>g conditions and short-term light exposure on<br />

signal<strong>in</strong>g performance <strong>in</strong> <strong>Drosophila</strong> <strong>photoreceptors</strong>. J Neurophysiol 92(3): 1918-27.<br />

Wong F, Knight BW & Dodge FA (1980) Dispersion of latencies <strong>in</strong> <strong>photoreceptors</strong> of Limulus and<br />

the adapt<strong>in</strong>g-bump model. J Gen Physiol 76(5): 517-37.<br />

Wu CF & Pak WL (1975) Quantal basis of photoreceptor spectral sensitivity of <strong>Drosophila</strong><br />

melanogaster. J Gen Physiol 66(2): 149-68.<br />

Zettler F (1969) Die Abhangigkeit des Übertragungsverhaltens von Frequenz und<br />

Adaptionszustand; gemessen am e<strong>in</strong>zelnen Lichtrezeptor von Calliphora erythrocephala. Z<br />

Verg Physiol 64: 432-449.<br />

Zhang W & L<strong>in</strong>den DJ (2003) The other side of the engram: experience-driven changes <strong>in</strong> neuronal<br />

<strong>in</strong>tr<strong>in</strong>sic excitability. Nat Rev Neurosci 4(11): 885-900.


Appendix<br />

Examples of the model m-files used <strong>in</strong> this study.<br />

A1. M-file used for simulat<strong>in</strong>g the current <strong>in</strong>jection experiments <strong>in</strong> Fig<br />

4a, IV<br />

%%% Script for current <strong>in</strong>jection simulations %%%<br />

global param param_1;<br />

for Im=[0.00248 0.00802 0.0133 0.0188 -0.00248 -0.00802 -0.0133]<br />

V0=-66; % mV, rest<strong>in</strong>g potential<br />

h0=1/(1+exp((-25.7-V0)/-6.4));<br />

n0=(1/(1+exp((-1-V0)/9.1)))^(1/2);<br />

mKA0=(1/(1+exp((-23.7-V0)/12.8)))^(1/3);<br />

hKA0=(0.8/(1+exp((-55.3-V0)/-3.9))+0.2/(1+exp((-74.8-V0)/-10.7)));<br />

gnew0=1/(1+exp((-14-V0)/10.6));<br />

y<strong>in</strong>it=[V0;h0;n0;mKA0;hKA0;gnew0]; % <strong>in</strong>itial values for diff. equations<br />

options=odeset(‘'OutputFcn','myparams'); % collects parameters<br />

tspan=[0 300]; % timespan for the solution<br />

param=[];<br />

[t,y]=ode15s(@cur_clamp,tspan,y<strong>in</strong>it,options,Im);% ODE solver<br />

% Plott<strong>in</strong>g the results<br />

figure(1); plot(t,y(:,1));title('Current <strong>in</strong>jection simulation');<br />

xlabel('ms)');ylabel('mV'); hold on;<br />

figure(2); plot(t,param(:,1).*15710,'r');hold on; % Shaker<br />

figure(2); plot(t,param(:,2).*15710,'b');hold on; % Shab<br />

figure(2); plot(t,param(:,3).*15710,'g');hold on; % Novel potassium conductance<br />

xlabel('ms)');ylabel('nS'); hold on;<br />

end


%%% WT model %%%<br />

function dy=cur_clamp(t,y,Im)<br />

global param param_1;<br />

Cm=4; % *10^-6 F/cm^2, membrane capacitance<br />

gl=0.0585*10^-3; % S/cm^2, Chloride leak conductance<br />

gl2=0.0855*10^-3; % S/cm^2, Potassium leak conductance<br />

glic=0; % S/cm^2, Light leak conductance<br />

gKsm=3*10^-3; % S/cm^2, Shab max conductance<br />

gKAm=0.8*10^-3; % S/cm^2, Shaker max conductance<br />

gnewm=0.11*10^-3; % S/cm^2, Novel K + channel max conductance<br />

gKAmf=0.087*10^-3; % S/cm^2, partial failure <strong>in</strong> Shaker <strong>in</strong>activation<br />

Vl=-30; % mV, Chloride leak rev. potential<br />

Vlic=10; % mV, LIC rev. potential<br />

VK=-85; % mV, Potassium rev. potential<br />

58<br />

% Current <strong>in</strong>jection protocol<br />

c_start=100; % start<strong>in</strong>g time<br />

t_cur_on=100; % duration<br />

t_cur_off=c_start+t_cur_on;<br />

if tc_start+t_cur_on<br />

I=0;<br />

else<br />

I=Im; % I=current <strong>in</strong>jection<br />

end<br />

end<br />

% Activation-<strong>in</strong>activation -parameters<br />

% Shaker<br />

if abs(exp((-y(1)-59.639)/4.50122)-1)


59<br />

bmKA=(1-((1/(1+exp((-23.7-y(1))/12.8))))^(1/3))/(1/(1.35*…<br />

(0.008174*exp((-y(1)+1.61882)/24.6538)+0.058139*(-y(1)-59.639)/…<br />

(exp((-y(1)-59.639)/4.50122)-1))));<br />

end<br />

if abs(exp((-y(1)+13.4859)/11.11)-1)


60<br />

gKA=((y(4)^3)*y(5)*gKAm)+((y(4)^3)*gKAmf); % Shaker<br />

gKs=(y(3)^2)*y(2)*gKsm; % Shab<br />

gnew=y(6)*gnewm; % Novel <strong>K+</strong> conductance<br />

% Differential equations<br />

dy=[-y(1)*((gKA+gKs+gl+glic+gl2+gnew)/(0.001*Cm))+(((gnew*VK)…<br />

+(gKA*VK)+(gKs*VK)+(gl*Vl)+(glic*Vlic)+(gl2*VK))/(0.001*Cm))+(I/(0.001*Cm))<br />

-y(2)*(ah+bh)+ah<br />

-y(3)*(an+bn)+an<br />

-y(4)*(amKA+bmKA)+amKA<br />

-y(5)*(ahKA+bhKA)+ahKA<br />

-y(6)*(anew+bnew)+anew];<br />

param_1=[gKA,gKs,gnew];<br />

%%% Assist<strong>in</strong>g function %%%<br />

function status=myparams(t,y,flag);<br />

% Date: 12/02/1998,Oulu; Han Chunlei. Modified 17.6.2001 by Mikko Vähäsöyr<strong>in</strong>ki.<br />

% This is for extract<strong>in</strong>g the parameters from ODE solvers,<br />

% the INPUT is PARAM_1; OUTPUT is PARAM;<br />

% code is based on the m-file of ODEPLOT.M.<br />

global param param_1;<br />

if narg<strong>in</strong> < 3 | isempty(flag)<br />

param=[param;param_1];<br />

else<br />

switch(flag)<br />

case '<strong>in</strong>it'<br />

param=[param;param_1];<br />

case 'done'<br />

end<br />

end<br />

status = 0;


61<br />

A2. M-file for reconstruct<strong>in</strong>g the LIC from experimental voltage<br />

response, IV<br />

%%% Reconstruct<strong>in</strong>g the LIC from Shab3 photoreceptor response %%%<br />

% Commented to the extent are not covered <strong>in</strong> A1<br />

global goal param param_1; % goal is the experimental voltage response<br />

% created <strong>in</strong> the workspace<br />

V0=-40;<br />

mKA0=(1/(1+exp((-23.7-V0)/12.8)))^(1/3);<br />

hKA0=(0.8/(1+exp((-55.3-V0)/-3.9))+0.2/(1+exp((-74.8-V0)/-10.7)));<br />

gnew0=1/(1+exp((-14-V0)/10.6)));<br />

y<strong>in</strong>it=[V0;mKA0;hKA0;gnew0];<br />

options=odeset('maxstep',0.5,'OutputFcn','Myparams');<br />

tspan=[0:1:9999];<br />

g=(0.1793*10^-3)*ones(10000,1); % <strong>in</strong>itial value vector for the LIC<br />

counter=0;<br />

while counter < 30 % Number of the iterations, solution converges usually fast<br />

param=[];<br />

[t,y]=ode113(@s3_lic_fit,tspan,y<strong>in</strong>it,options,g);<br />

for <strong>in</strong>d=1:1:length(goal),<br />

diff=goal(<strong>in</strong>d)-y(<strong>in</strong>d,1);<br />

err=abs(diff/goal(<strong>in</strong>d));<br />

if err > 0.00005 % Calculates the error <strong>in</strong> each po<strong>in</strong>t<br />

if y(<strong>in</strong>d,1) >= 0 % and the LIC values are iterated if the error<br />

delta=diff/goal(<strong>in</strong>d); % exceeds the set value<br />

else<br />

delta=-diff/goal(<strong>in</strong>d);<br />

end<br />

g(<strong>in</strong>d)=g(<strong>in</strong>d)*(1+delta);<br />

end<br />

end<br />

counter=counter+1;<br />

end<br />

% Plott<strong>in</strong>g the results<br />

figure(1);plot(t,y(:,1),t,goal); title('<strong>Voltage</strong>');xlabel('ms'); ylabel('mV');<br />

figure(2); plot(t,param*15710); title('LIC'); xlabel('ms'); ylabel('nS');


%%% Shab3 model %%%<br />

function dy=s3_lic_fit(t,y,g)<br />

global goal param param_1;<br />

Cm=2.44;<br />

gl=0.171*10^-3;%<br />

gl2=0.2055*10^-3;<br />

gKAm=0.8*10^-3;<br />

gnewm=0.11*10^-3;<br />

gKAmf=0.087*10^-3;<br />

Vl=-30;<br />

Vlic=10;<br />

VK=-85;<br />

glic=g(length(param)+1);<br />

I=0;<br />

if abs(exp((-y(1)-59.639)/4.50122)-1)


end<br />

anew=((1/(1+exp((-14-y(1))/10.6)))^(1/1))/((13+(6232/(30*sqrt(pi/2)))*…<br />

exp(-2*((y(1)+19.4)/30).^2))/1.35);<br />

bnew=((1-(1/(1+exp((-14-y(1))/10.6))))^(1/1))/((13+(6232/(30*sqrt(pi/2)))*…<br />

exp(-2*((y(1)+19.4)/30).^2))/1.35);<br />

63<br />

gKA=((y(2)^3)*y(3)*gKAm)+((y(2)^3)*gKAmf);<br />

gnew=y(4)*gnewm;<br />

dy=[-y(1)*((gKA+gl+glic+gl2+gnew)/(0.001*Cm))+(((gKA*VK)+(gl*Vl)+…<br />

(glic*Vlic)+(gl2*VK)+(gnew*VK))/(0.001*Cm))+(I/(0.001*Cm))<br />

-y(2)*(amKA+bmKA)+amKA<br />

-y(3)*(ahKA+bhKA)+ahKA<br />

-y(4)*(anew+bnew)+anew];<br />

param_1=[glic];

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!