• Nenhum resultado encontrado

Strength–duration relationship for intra- versus 3 extracellular stimulation with microelectrodes

N/A
N/A
Protected

Academic year: 2021

Share "Strength–duration relationship for intra- versus 3 extracellular stimulation with microelectrodes"

Copied!
13
0
0

Texto

(1)

1

2

STRENGTH–DURATION RELATIONSHIP FOR INTRA- VERSUS

3

EXTRACELLULAR STIMULATION WITH MICROELECTRODES

4Q1 F. RATTAY,a*L. P. PAREDESaAND R. N. LEAOb,c

5 aInstitute for Analysis and Scientific Computing, Vienna University 6Q2 of Technology, A-1040 Vienna, Austria

7 bNeurodynamics Laboratory, Department of Neuroscience, 8 Uppsala University, Uppsala, Sweden

9 cBrain Institute, Federal University of Rio Grande do Norte, Natal-RN, 10 Brazil

11 Abstract—Chronaxie, a historically introduced excitability time parameter for electrical stimulation, has been assumed to be closely related to the time constant of the cell mem-brane. Therefore, it is perplexing that significantly larger chronaxies have been found for intracellular than for extra-cellular stimulation. Using compartmental model analysis, this controversy is explained on the basis that extracellular stimulation also generates hyperpolarized regions of the cell membrane hindering a steady excitation as seen in the intracellular case. The largest inside/outside chronaxie ratio for microelectrode stimulation is found in close vicinity of the cell. In the case of monophasic cathodic stimulation, the length of the primarily excited zone which is situated between the hyperpolarized regions increases with electrode–cell distance. For distant electrodes this results in an excitation process comparable to the temporal behav-ior of intracellular stimulation. Chronaxie also varies along the neural axis, being small for electrode positions at the nodes of Ranvier and axon initial segment and larger at the soma and dendrites. As spike initiation site can change for short and long pulses, in some cases strength–duration curves have a bimodal shape, and thus, they deviate from a classical monotonic curve as described by the formulas of Lapicque or Weiss.Ó 2012 Published by Elsevier Ltd on behalf of IBRO.

Key words: chronaxie, strength–duration curve, electrical stimulation, microelectrode, activating function.

12

13 INTRODUCTION

14 Selective neural stimulation is a great challenge in the

15 development of neural prostheses. As an example, active 16 contacts of an electrode array implanted at the retina or

17 other structures along the visual pathway should stimulate 18 elements that elicit visual sensations corresponding to the

19 place in the array (Brindley, 1955; Zrenner, 2002; Dowling,

20 2005; Ferna´ndez et al., 2005; Fried et al., 2006; Cohen,

21

2007; Sekirnjak et al., 2008; Horsager et al., 2009; Pezaris

22

and Reid, 2009; Tehovnik et al., 2009)while avoiding side Q3 23

effects like co-stimulation of bypassing axons arising from

24

distant locations (Greenberg et al., 1999; Rattay and

Re-25

satz, 2004). Considering the fact that neural substructures

26

have different strength–duration characteristics may give

27

an opportunity for more selective stimulation. For example,

28

in the treatment of chronic pain during spinal cord

stimula-29

tion smaller dorsal column fibers can only be activated

30

when pulse width is sufficiently large (Holsheimer et al.,

31

2011).

32

Recently, the strength–duration relationship for

extra-33

cellular neural stimulation was analyzed under the

34

assumption of a constant electrical field in flat, spherical

35

and cylindrical cells (Boinagrov et al., 2010). These

inves-36

tigations provided a biophysical basis for the stimulation

37

with large electrodes and also explained effects of

stimula-38

tion with ultrashort pulses (<5

l

s). In accordance with

ret-39

inal ganglion cell experiments (Jensen et al., 2005), for

40

pulse durations <2 ms, it was found that a 500

l

m

diam-41

eter electrode placed above the soma caused excitation

42

with significantly lower thresholds compared to a position

43

above the axon. This result is unexpected given the

44

assumption that the axon is the most excitable part of a

45

neuron externally stimulated (Nowak and Bullier, 1998;

46

Rattay, 1999). Differences in experimental setups are the

47

reason for the controversy about such observations.

Stim-48

ulation of the retina with large electrodes generates a

49

rather constant field and consequently transverse currents

50

depolarize the cell membrane at one side and

hyperpolar-51

ize the cell at the opposite side. On the other hand, external

52

excitation with small electrodes is mainly based on

stimu-53

lating effects resulting from extracellular potential

varia-54

tions along the neural structure (Rushton, 1927; Ranck,

55

1975; Rattay, 1986, 1987, 1999). These facts vary the

ex-56

pected excitation paradigm and should be carried in mind.

57

Voltage sensitive ion channel types and densities are

58

other important elements for neuron excitability. In

rela-59

tion to retinal implant, electrode positions applied at a

60

dense two-dimensional grid around the soma region of

61

rabbit ganglion cells showed the lowest thresholds along

62

a small section of the axon, about 40

l

m from the soma.

63

At this axonal section, immunochemical staining revealed

64

a dense band of voltage-gated sodium channels (Fried et

65

al., 2009).Similarly, high sodium channel densities at the Q4 66

axon initial segments (AIS) of cortical cells define sites for

67

action potential initiation under natural conditions (Stuart

68

et al., 1997; Yu et al., 2008; Hu et al., 2009) as well as

69

spike initiation candidates for external stimulation with

70

microelectrodes (Rattay and Wenger, 2010).

0306-4522/12 $ - see front matterÓ 2012 Published by Elsevier Ltd on behalf of IBRO.

http://dx.doi.org/10.1016/j.neuroscience.2012.04.004

*Corresponding author. Tel: +43-1-58801-10114. E-mail address:frank.rattay@tuwien.ac.at(F. Rattay).

Abbreviations: AIS, axon initial segments; VSD, voltage sensitive dye. Neuroscience xxx (2012) xxx–xxx

(2)

71 A strength–duration curve describes electrode thresh-72 old current as function of stimulus pulse duration.

Chron-73 axie is defined as the time on such a strength–duration 74 curve for twice the minimum (rheobase) current needed

75 for very long pulses. Established by experimental

find-76 ings, this excitability parameter was assumed to be rather

77 independent from the distance between the current

78 source and the excited cell (Weiss, 1901; Lapicque,

79 1907, 1929; Blair, 1932).

80 As different neural structures have different

chronax-81 ies, stimulus pulse duration is an important element for

82 selective stimulation as well as for electrophysiological 83 classification. For example, the shorter chronaxie of

mye-84 linated motorneurons compared to that of unmyelinated

85 fibers of the myocardium allows safe stimulation with

86 short pulses during artificial respiration. Motoneurons

87 then become activated with little disturbance of the

88 myocardial function (Voorhees et al., 1992). In

electro-89 physiological studies, extracellular gray matter stimulation 90 shows larger chronaxies for somas than for axons,

sup-91 porting the idea that spike initiation occurs in the axon

92

(Nowak and Bullier, 1998). Additionally, in cochlear

im-93

plants, significantly shorter chronaxies have been

de-94

scribed in long-term (compared to acute) deafened

95

cochleae, indicating the loss of peripheral processes in

96

many spiral ganglion cells (Shepherd et al., 2001).

97

Based on theoretical considerations, chronaxie has

98

been suggested to be about 0.7 times the time constant

99

of the cell membrane (Blair, 1932; Ranck, 1975; Reilly,

100

1992). Therefore, some authors assume chronaxie as

101

independent of in- or outside stimulation of the cell

102

(Nowak and Bullier, 1998). This hypothesis contrasts with

103

reports of considerably larger chronaxies for intracellular

104

than for extracellular stimulation (Ranck, 1975).

105

Discrepancy among different works prompts

specula-106

tions that experimental differences could be caused by

107

artifacts like sampling errors or neuronal damage during

108

recordings (Ranck, 1975). To address this controversy,

109

we systematically reproduced experimental findings using

110

computer simulations of a pyramidal cell and its simplified

111

rectified version for electrode positions at the dendrite, the

112

soma, the lateral ending of the AIS and the myelinated

0.0001 0.001 0.01 0.1 1 µA 10 100 0.01 0.1 1 10 ms SOMA soma soma ais ais AIS

dend is not excitable for shorter pulses

DEND AIS NODE NODE NODE soma ais internode dend SOMA DEND AIS dend dend DEND SOMA PULSE DURATION CHRONAXIE NEURON 2 NEURON 1 extracellular stimulation intracellular stimulation node node node basal dendrite main dendrite d~4,7µm electrode positions 50 µm above first internode naked (thin) axon N.R. = node of Ranvier 200µm dendrite unmyel. axon myelin. axon N.R. N.R. N.R. N.R. rectified neuron soma AIS naked axon

axon hillock AIS

NODE soma dend node apical dendrite N.R. N.R. N.R. N.R. N.R. N.R. N.R. N.R. N.R. N.R. N.R. N.R. N.R. x N.R.

B

A

Fig. 1.Strength–duration curves of the investigated pyramidal cell. (A) Neuron 1 (left) and Neuron 2 (right) with selected electrode positions. (B) Strength–duration curves with chronaxie values for electrode positions as marked in A. Lower and upper case letters are used for electrode positions for Neuron 1 and Neuron 2, respectively. Color coding of lines in B corresponds to electrode positions in A. For better discrimination some extracellular strength–duration curves are shown as broken lines. Colors of break point markers are without meaning. Anodic and cathodic square pulses for intra- and extracellular stimulation, respectively.

(3)

113 axon (Fig. 1). In all cases shown inFig. 1B, the intracellu-114 lar/extracellular chronaxie ratio is two or higher. It is

115 important to note that these ratios depend on cell proper-116 ties as well as on the distance and type of electrodes.

117 Such variants could be the reason for the diversity in

118 chronaxie data.

119 Here, we demonstrate that (i) hyperpolarized regions

120 are responsible for the shorter extracellular chronaxies, 121 (ii) for short electrode distances, excitability and chronaxie

122 differ essentially along a selected neuron and conse-123 quently (iii) strength duration curves are expected to

devi-124 ate from the classical form when short and long pulses 125 cause action potential initiation at sites with different

elec-126 trical membrane properties.

127 EXPERIMENTAL PROCEDURES

128 An analysis of two model neurons is presented. In both models 129 excitation is based on recently measured densities of high thresh-130 old sodium channels Nav1.2 in dendrites and the soma and low 131 threshold sodium channels Nav1.6 in the axon (Hu et al., 132 2009). The phenomenon of short chronaxies for extracellular 133 stimulation was also tested in simulations with other neural struc-134 tures (not shown) and did not depend on the specific examples. 135 Model Neuron 1 (Fig. 1A left,Rattay and Wenger, 2010), a 136 simplification of Neuron 2 concerning geometry and ion channel 137 types, is straightforward to analyze and with less computational 138 cost. It has a straight axis and consists of a single non-branching 139 dendrite (500lm, d = 5lm), spherical soma (d = 20lm), axon 140 hillock (10lm, d = 3.1lm), AIS (50lm, d = 1.22lm), naked 141 axon (unmyelinated, 200lm, d = 1lm), myelinated axon 142 (500lm, d = 1lm) and unmyelinated terminal (50lm, 143 d = 1lm). Assumptions for ion channel distribution and ion cur-144 rent computations are quite similar as inHu et al. (2009): the 145 same constant Nav1.2 channel density for the dendrite and soma 146 (gna = 8 mS/cm2), but 40 times higher sodium channel density 147 in hillock and AIS with a change to the low threshold type 148 Nav1.6 in the axon. Intracellular resistivity is 150Xcm, mem-149 brane capacity c = 1lF/cm2. Using ACSL (Advanced Continu-150 ous Simulation Language) software, Neuron 1 is simulated by 151 evaluating the compartment model equations (1)–(3) as de-152 scribed below.

153 Model Neuron 2 (Fig. 1A right) is based on geometrical 154 parameters of a traced cortical layer 5 pyramidal cell with ion 155 channel assumptions as available in the NEURON (Carnevale 156 and Hines, 2006) Model DB (Hines et al., 2004; Hu et al., 157 2009). In contrast to Neuron 1, soma diameter is increased to 158 30lm, naked axon diameter is reduced to 0.4lm and the capac-159 ity of the cell membrane of all compartments with exception of the 160 soma and internodes is reduced to 0.5lF/cm2. Moreover, 161 Neuron 2 incorporates branching dendrites, tapering diameters, 162 uneven ion channel distribution within compartments of a single 163 type and a membrane capacity of 0.02lF/cm2in the axonal inter-164 nodes. High threshold sodium Nav1.2, low threshold Nav1.6 and 165 fast voltage-gated K+of Neuron 1 are complemented by slow 166 non-inactivating potassium current, high-voltage activated Ca2+ 167 and calcium dependent K+ in dendritic and somatic compart-168 ments. All parameters are used as inHu et al. (2009). However, 169 in order to demonstrate current–distance relations in a clear way, 170 the axial 3d structure of the cell was compressed into the x–y 171 plane. Neuron 2 is simulated with NEURON.

172 In more detail, Neuron 1 soma and dendrite were imple-173 mented with the same constant maximum conductances (in mS/ 174 cm2) g

Nav1.2= 8, gNav1.6= 0 and gKv= 10, the axon hillock with 175 gNav1.2= 320, gNav1.6= 0, gKv= = 100, the AIS with gNav1.2= 176 100, gNav1.6= 320, gKv= 100, the unmyelinated axon with 177 gNav1.2= 0, gNav1.6= 300, gKv= 150 and the nodes of Ranvier 178 with gNav1.2= 0, gNav1.6= 160, gKv= 20. Sodium current kinetics

179 are calculated by INav1.j= gNav1.jm3h (V ENa) with j equals to

180 either 2 or 6 and ENa= 60 mV. Details on the differential equations

181 of the different variables are shown inMainen et al. (1995). The

182 values for the half (in)activation voltages V1/2, the slopes k and

183 the coefficients A were obtained from a previously published

mod-184 el in the NEURON Model DB (Hu et al., 2009) after subtracting the

185 corresponding value for the shift of voltage dependence of the

186 kinetics. Consequently, the currents INav1.jhave the same values

187 for A, i.e., A(am) = 0.182, A(bm) = 0.124, A(ah) = 0.024, A(bh

188 ) = 0.0091, and the slope of inactivation, i.e., k(sh) = 5 and

189 k(h1) = 6.2, in contrast to altered slope of activation, i.e.,

190 k(sm) = k(m1) = 7 for Nav1.2 but k(sm) = k(m1) = 6 for

191 Nav1.6. To account for the reduced threshold of Nav1.6 channels

192 V1/2(m) is decreased to41 mV compared to the calculated value

193 of28 mV for activation of Nav1.2 channels. The corresponding

194 values in mV for the inactivation of Nav1.2/Nav1.6 channels are

195 V1/2(ah) =35/41, V1/2(bh) =60/73, V1/2(h1) =57/70.

196 The potassium currents are determined by IK= gKn(V EK) with

197 EK =90 mV. To be consistent the corresponding values of

198 A(a) = 0.02, A(b) = 0.002, V1/2(a) = V1/2 (b) = 25 mV and

199 k(a) = k(b) = 9 were also obtained from the NEURON Model

200 DB (Hu et al., 2009). Internodes are simulated with 17 sheets of

201 membrane with a conductance of 1 mS/cm2and C = 1lF/cm2

202 per sheet (Rattay, 1999). The presented results are simulated

203 for 37°C.

204 In the first modeling step, the extracellular potential Ve

gener-205 ated by an electrode tip is approximated considering a monopolar

206 spherical electrode in an infinite homogeneous extracellular

med-207 ium with resistivityqe= 300Xcm. This is equivalent to a point

208 source stimulation resulting in spherical isopotentials with Ve=

209 qeIel/4pr when a current pulse with amplitude Ielis applied; r is

210 the distance from a point of interest to the point source.

211 In the second step, the response of a neuron is simulated by

212 a compartment model. That is a network of resistances and

213 capacitances, where the current to the centre of compartment n

214 consists of the following components: a capacitive current, ion

215 currents across the membrane and intracellular currents to

neigh-216 bored compartments. Applying Kirchhoff’s law for compartment n

217 results in 218 dðVi;n Ve;nÞ dt  Cnþ Iion;nþ Vi;n Vi;n1 Rn=2þ Rn1=2 þ Vi;n Vi;nþ 1 Rn=2þ Rnþ1=2 ¼ 0 ð1Þ 220 220 221 with intracellular potential Vi, axial resistance R and membrane

222 capacity C. The following system of differential equations is deduced

223 by introducing the transmembrane voltage V¼ Vi Veto compute

224 the time courses of Vnin every compartment (Rattay, 1999):

225 dVn dt ¼ Iion;nþ Vn1 Vn Rn1=2þ Rn=2 þ Vnþ1 Vn Rnþ1=2þ Rn=2  þ Ve;n1 Ve;n Rn1=2þ Rn=2 þ Ve;nþ1 Ve; n Rnþ1=2þ Rn=2  Cn: ð2Þ 227227 228 The direct stimulating influence of the extracellular potential on

229 compartment n is defined by the activating function (Rattay, 1999)

230 fn¼ Ve;n1 Ve;n Rn1=2þ Rn=2 þ Ve;nþ1 Ve;n Rnþ1=2þ Rn=2   Cn: ð3Þ 232 232 233 For a fiber with constant diameter d, constant compartment length

234 Dx, intracellular resistivityqiand specific capacity c,(3)appears

235 in a simpler form 236 fn¼ d 4c

q

i

Ve;nþ1 2Ve;nþ Ve;nþ1

Dx2 : ð3aÞ 238238

239 The value within the brackets of(3)corresponds to a virtual

in-240 jected current applied to compartment n. In regions where this

cur-241 rent is positive, the membrane depolarizes and where it is

242 negative, it tends to hyperpolarize.

(4)

243 VOLTAGE SENSITIVE DYE IMAGING 244 Transverse hippocampal slices were obtained from P21

245 C57B6 mice as previously described (Leao et al., 2009)

246 and according to the rules of Animal Experimentation of

247 the Uppsala University. Slices were maintained in artificial 248 ACSF (in mM: 124 NaCl, 3.5 KCl, 1.25 NaH2PO4, 1.5

249 MgCl2, 1.5 CaCl2, 30 NaHCO3, 10 glucose), constantly

250 bubbled with 95% O2and 5% CO2. Recordings/imaging

251 were obtained at 25°C, in the presence of 10

l

M CNQX, 252 30

l

M dAP5 and 10

l

M bicuculline methochloride to

min-253 imize the effect of synaptic currents (Leao et al., 2005).

254 Voltage sensitive dye (VSD) loading (JPW3027 obtained 255 from Prof Leslie Loew, University of Connecticut, USA)

256 was performed exactly as described inPalmer and Stuart 257Q5 (2006).Images were acquired using a EM-CCD camera

258 (Luca, Andor, Ireland). Excitation was produced by a 259 200W metal-halide lamp through a bypass filter centered

260 at 535 nm (510 to 560 nm pass) and emission was

low-261 pass filtered at 590 nm. Image acquisition and extracellular

262 stimulation were synchronized by a National Instruments

263 digital device and to guarantee time precision, we recorded 264 the stimulator and the camera ‘fire’ outputs (that flags

im-265 age acquisition) using a National Instruments DAQ card. 266 Current clamp recordings were also obtained from the

im-267 aged cell using winWCP (Dr John Dempster, Strathclyde 268 University, UK). The stimulating electrode (tungsten,

269 10

l

m tip) was placed 50 and 75

l

m above axons. Pulses

270 had 100

l

s durations and extracellular current intensities 271 were adjusted to 75% of the minimum threshold current.

272 Images were taken 100

l

s after the stimulus using

273 100

l

s exposure. Pixel intensity was measured in 4 2

274 pixel region of interest and averaged from 10 images ac-275 quired with 1s delay between each other.

276 RESULTS

277 In the following several computer experiments are

278 proposed to analyze step by step characteristics of

279 strength–duration curves and related historical formulas.

280 Intracellular stimulation: space clamp versus cable

281 model

282 In order to describe the neural excitation process

inde-283 pendently from the intracellular current needed for spike 284 conduction,Hodgkin et al. (1952)used the space clamp

285 technique. They stimulated with a long inserted wire that

286 clamped the squid axon uniformly along its entire length. 287 An equivalent model situation is created by current

injec-288 tion into the soma of Neuron 1 by cutting all soma pro-289 cesses. Stimulation of the resting soma with a rather

290 long weak pulse results in a transmembrane voltage

291 curve V(t) with an asymptotic exponential increase during

292 the subthreshold response (Fig. 2A). The time constant

293 for the exponential voltage of the passive membrane is 294 the product of the ohmic resistance and the capacity of

295 the cell membrane (

s

= R C). The time constant

s

of 296 such an exponential increase can be found graphically

297 by linear extrapolation at stimulus onset as the intersec-298 tion of the tangent in V at pulse onset with the steady state

299 indicated by the horizontal dashed line inFig. 2A. At the

300

end of the 8 ms pulse, the subthreshold approach of V

301

is very close to its stationary value.

302

The shape of the subthreshold membrane voltage V(t)

303

changes considerably when the stimulus current is

304

allowed to flow not only across the membrane but also

305

intracellularly from the stimulated soma into the dendrite

306

and the axon. Finding

s

by the graphical method of

307

Fig. 2A becomes contradictory when applied in Fig. 2B

308

and C as it results in shorter time values t1 while needing

Iel=50pA Iel=50pA I =50pAel I =150pAel 1ms 1ms 1ms 10mV 2mV 0.63*Vstationary 0.63*Vstationary 0.63*Vstationary 1ms 1mV 30mV transmembrane voltage V =3.4ms t1=1.4ms, t 2=4.3ms t1=1.4ms, t 2=5.3ms STIMULUS CURRENT STIMULUS CURRENT STIMULUS CURRENT Multiplication of subthreshold response with a factor 3 corresponding to Iel=150pA

Subthreshold response Iel=50pA Soma without processes, intracellular stimulation

Soma with dendrite, intracellular stimulation

C

Soma with axon and dendrite, intracellular stimulation

A

B

Fig. 2.Transmembrane voltage V of Neuron 1 for intracellular stimulation at the soma. (A) Space clamp condition with no axial current flow. The top graph demonstrates the coincidence of the passive and active membrane response in the subthreshold regime. Linear extrapolation of V at stimulus onset up to the steady state value of the single RC circuit (gray arrow) defines the time constant of the passive membrane response ass= R C (lower graph of A). With the same electrode current Iel= 50 pA, the steady state value of V is reduced by a factor5 when axial current flows into the dendritic branch (B) and by a factor10 for conductance into dendrite and axon (C). In comparison with the space clamp condition, intracellular current flow into dendrite and axon results in shorter times t1 but longer times to reach the steady state of the subthreshold membrane.

(5)

309 longer times for reaching the steady state (pulse dura-310 tions >20 ms). Obviously, the fit by a single exponential

311 function corresponding to a single RC element (as in case 312 A, where 63% of Vstationary is determined by

s

) cannot

313 describe the subthreshold response of the cable model

314 in an acceptable way. We have to change the method

315 and define the time constant as the time t2 when the cable

316 model reaches 63% (1 1/e) of the stationary membrane 317 voltage V (Fig. 2B and C). Note that the time constant of

318 the cable model is not an average value of the time

con-319 stants of the compartments. This is demonstrated by

320 Fig. 2B where both the dendrite and the soma of Neuron1 321 have quite the same electric membrane properties. The

322 main difference between

s

and its replacing values t1

323 and t2 inFig. 2A and B is a consequence of a distinctly 324 reduced steady state membrane voltage when the greater

325 part of the stimulus current flows into the dendrite. The dif-326 ferent temporal excitation profiles shown inFig. 2 have

327 consequences on the corresponding strength–duration

328 curves.

329 The middle solid line in Fig. 3 shows the strength–

330 duration relation under the space clamp condition where

331 chronaxie is independent from the soma area. Allowing

332 part of the injected current to flow into dendrites and the 333 axon demands for stronger threshold currents for all pulse

334 durations (Fig. 3, upper solid line). The intracellular

335 strength–duration curves for Neuron 1 and its space

336 clamped version deviate in their shapes for long pulses. 337 With the same electrical cell properties, the calculated

338 intracellular chronaxie ratio 2.5/1.79 = 1.4 (Fig. 3)

339 becomes smaller by enlargement of the soma surface

340 or by reducing the number or diameters of processes.

341 The variable chronaxie ratio decreases to 1 if the diame-342 ters of the soma processes converge to zero, which is the

343 space clamp condition.

344 This dependence of chronaxie on diameter disproves

345 the often applied rule mentioned in the introduction, 346 namely chronaxie0.7

s

with

s

= R C, where the

prod-347 uct R C is independent of membrane size. However,

348 even more surprising is the huge deviation from this for-349 mula for extracellular stimulation. With the time constant

350 of the soma membrane (Fig. 2A) we obtain chronaxie 351 0.7  3.4 ms = 2.38 ms instead of 0.38 ms for the case 352 presented inFig. 3.

353 In contrast to intracellular stimulation, in most

applica-354 tions extracellular stimulations are optimally achieved with

355 cathodic currents (Ranck, 1975; Rattay, 1986, 1999). 356 Hence, we compare anodic inside with cathodic outside

357 stimulation in order to explain in the next subsections 358 the large chronaxie differences between intra- and

extra-359 cellular stimulation shown in Figs.1B and3.

360 Extracellularly activated region increases with

361 electrode distance

362 A straight fiber with the properties of the naked axon of 363 Neuron 1 is shown inFig. 4A–C for external stimulation

364 with a cathodic 100

l

s pulse at threshold intensity. In con-365 trast to intracellular stimulation, the externally positioned

366 microelectrode causes in every compartment an injected

367 virtual current. As second important effect, outside

368 stimulation with a monophasic pulse produces virtual

369

currents of both polarities where the sum of all the virtual

370

currents is zero.1The driving forces of excitation are the

371

currents in the region with positive activating function

val-372

ues (red arrows2inFig. 4C).

373

The length of the region where a cathodic point source

374

causes positive activating function ispffiffiffi2 zel, where zelis

375

the electrode distance. This formula is correct for a

376

straight nonmyelinated fiber with constant diameter in a

377

homogeneous medium and the relation can be described

378

by a 70-degree angle (Fig. 4A, tan(35°) =pffiffiffi2=2 zel=zel),

379

(Rattay, 1986). According to the activating function

380

concept, this angle is independent of fiber diameter and

381

electrical membrane properties and can also be used as

382

an approach for myelinated axons (Rattay, 1986). Large

383

positive and negative isolated activating functions’ values

384

appear at locations with considerable diameter changes,

385

in branching or in fiber endings. As axial intracellular

cur-386

rents level such local effects, the 70° rule is a rough

ap-387

proach even for these cases. A specific chronaxie value

388

of a region can be expected as long as the zone defined

389

by the 70° seen from the point source is concentrated in a

390

cell region with common electrical properties.

391

The length of the activated portion flanked by the

hyper-392

polarized regions increases with electrode–neuron

dis-393

tance, and consequently, the influence of the

394

hyperpolarized region on the excitation process is

gradu-395

ally reduced when moving the current source away. This 1 10 100 1000 µA 0.01 0.1 1 10ms electrode current Iel rheobase I0 2 rheobase* pulse duration extracellular chronaxie = 0.38ms space clamp chronaxie = 1.79ms intracellular chronaxie = 2.5ms

Fig. 3.Strength–duration curves of the somatic membrane of Neuron 1 for stimulation with square pulses. Chronaxie defined as the pulse duration needed for twice the rheobase I0 (horizontal arrows) is different for the space clamp condition (central group of three curves) and the case of current flowing into dendrites and the axon (upper curves). Injected anodic current amplitudes (upper and central groups) are shown as current densities for 1 cm2 membrane. Extracellular stimulation (lower curves) is with cathodic currents for a point source 50lm above soma center. Approximation of the strength–duration curves according to the classical formulas of Lapicque (1907): Iel= I0/(1 2k) (dashed lines) andWeiss (1901): Iel= I0(1 + 1/k) (dotted lines) where k is pulse duration/chronaxie.

1 Eq.(1)is a current balance where the last two terms describe the intracellular current flow to the left and right neighboring compartments. Shifting to the next compartment (n ? n + 1) includes the ‘old’ right neighbor as new left neighbor, but the current flow has changed signs. Thus, the sum of these two currents is 0 and consequently the sum of all axial currents is 0. Applying this principle in Eq. (2) one find that the sum of all virtual stimulating currents is 0. For details seeRattay (1990).

2 For interpretation of color in Figs. 1, 4 and 10, the reader is referred to the web version of this article.

(6)

396

effect could be also demonstrated using VSD imaging. The

397

length of the depolarization detected by the intracellular

398

VSD increases with electrode distance (Fig. 4D). The area

399

under the curve in fluorescence (normalized by the

maxi-400

mum) versus length graph was used as a measurement

401

of depolarization ‘spread’ caused by the extracellular

stim-402

ulus. When the stimulating electrode was placed 50

l

m

403

(zel= 50

l

m) above the axon, the fluorescence (DF/F0,

404

normalized) versus length integral was equal to 0.038 ±

405

0.002

l

m1 and 0.058 ± 0.004

l

m1 for zel= 75

l

m

406

(n = 5 cells, p = 0.008, paired t test).

407

Intracellular versus extracellular stimulation

408

The spatio-temporal evolution of transmembrane voltage

409

profiles is significantly different for a microelectrode

posi-410

tioned either above or in the soma (Fig. 5A–D). InFig. 5A

411

and B, pulse duration is 100

l

s and membrane voltages

412

along the neural axis are compared for threshold intensity.

413

At the end of the pulse, extracellular stimulation shows a

414

larger voltage maximum, but a shorter depolarized region.

415

For 1 ms threshold pulses the same voltage profiles

ap-416

pear again with smaller amplitudes (due to lower

stimula-417

tion currents) as the firsts of ten advancing 100

l

s time

418

steps (thick lines inFig. 5C and D). The rather constant

in-419

crease of the maximum and spatial extension of the

volt-420

age profile for the case of intracellular stimulation in the

421

advancing 100

l

s time steps demonstrate that intracellular

422

stimulation is more effective as the injected current is

avail-423

able to load the membrane capacity in the vicinity of the

424

electrode (Fig. 5D). Virtual negative injected currents

dur-425

ing pulse application (Eq.(3)) cause strong

hyperpolariza-426

tion (Fig. 5A and C). An essential part of the virtual positive

427

current, the driving force for excitation, is lost as a result of

428

the counterbalancing axial current flow between the

depo-429

larized and the two hyperpolarized regions. Note that half

430

of the positive voltage profile (its extension and maximum

431

value) is already reached at approx. 10% of the 1 ms

stim-432

ulation pulse (thick line inFig. 5C). This is contrary to the

433

gradual increase of the voltage profile seen inFig. 5D.

434

A second important effect is demonstrated in the

435

10 ms pulse example (Fig. 6). The subthreshold response

436

for extracellular stimulation has a maximum at half of the

437

stimulus pulse time with a decay related to the inactivation

438

gating variable h (Mainen et al., 1995) that act as a factor

439

for the sodium current3(Fig. 6A). The quicker

extracellu-440

larly elicited voltage increase affects the lower h values.

441

The maximum appears earlier for extracellular electrode

442

positions above AIS and node of Ranvier (Fig. 6C). In

con-443

trast,Fig. 6B and D shows a rather constant increase in

444

membrane voltages indicating that stimulation is still

possi-445

ble with longer and weaker pulses.

446

Chronaxie increases with electrode distance

447

In the previous section and with Fig. 3, it was

demon-448

strated that axial current flow from the activated area into

A

B

C

D

hyperpolarized region activated region hyperpolarized region z =50µmel 70° 2 x zel * z axon 0 -25 -25mV -50mV -100mV -50 -75 -100 750 500 250 -250 0 electrode I =-25µA, Vel el=-500mV extracellular potential Ve along the fibre

activating function and virtual injected currents 100µm 0.2% F/F0 ED=50 m ED=75 m ED=75 m ED=50 m Soma

Arbitrary Pixel Intensity

Axon Rest 1 0 mV/ms mV

Fig. 4.Extracellular stimulation. (A) Geometry and isopotentials for a point source 50lm above a fiber positioned at the x-axis. (B) Extracellular potential Ve=qeIel/4pr with r¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ðx  xelÞ2þ z2el q

is used to calculate the activating function. (C) A fiber with d = 1lm and compartment lengthDx = 10lm results in a peak activating function value of 740 mV/ms. According to(2)this is the slope of the membrane voltage in the compartment below the electrode at the beginning of the25lA pulse. According to(3), the virtual currents are fn Cn. With Cn= dp Dx  c = 3.14159 * 107lF and specific membrane capacity c = 1lF/cm2, the maximum injected current is 232 pA at the center of the activated region. As the length of the activated region is defined by an angle of 70°, this region increases with electrode distance. (D) VSD imaging of a pyramidal cell axon at rest (upper photomicrograph), 100ls after the stimulus for electrode distance zel= 50lm (middle photomicrograph) and zel= 75lm (lower photomicrograph) (scale bar=10lm;⁄

electrode position in the horizontal axis). Fluorescent traces for zel= 50 and 75lm versus axon position is shown below the photomicrographs and are aligned to the region in the axon activated by the extracellular stimulus. Each point in the traces represents average fluorescence of an axon region of approx. 10lm length in five images.

3 Sodium current, the driving component in excitation, is modeled as g Nam3 h(V ENa), with maximum conductance gNa, gating variables m and h, membrane voltage V and Nernst potential ENa. In the resting state m has a quite small value in contrast to the inactivation variable h which starts with a high value that decreases (with some delay) when V increases.

(7)

449 the hyperpolarized regions causes shorter chronaxies for 450 cathodic extracellular than for intracellular stimulation. It

451 was also shown that the activated length is related to an

452 angle of 70° at the electrode and this activated length in-453 creases with microelectrode distance (Fig. 4).

Conse-454 quently, stimulating current loss along the neural axis 455 into the side lobes as defined by the activating function

456

shows a reduction when intracellular resistance between

457

depolarized and hyperpolarized regions increases by

458

increasing electrode distance. Comparison of excitation

459

profiles for electrodes at 50 and 200

l

m above the soma

460

(Fig. 7) emphasizes the trend to longer chronaxies for

lar-461

ger electrode distances. As seen in the case at 200

l

m

462

above the soma, the larger distance between the primarily 500 µm

20 mV extracellular stimulation intracellular stimulation

pulse duration 0.1 ms pulse duration 1 ms

A

B

C

D

Iel =-37 µA I =-8.6 µAel I =1.0 nAel I =7.5 nAel resting voltage -70 mV

dendrite axon dendrite axon

soma soma

electrode electrode

t=100 µs t=100 µs

Fig. 5.Spatial transmembrane voltage profiles for extra- and intracellular stimulation computed for Neuron 1. Extracellular stimulation 50lm above the soma (A, C) with cathodic threshold currents. Corresponding intracellular anodic threshold current cases at the right side (B, D). Voltage along the neural axis at the end of a 100ls stimulation pulse (A, B) and in ten advancing 100ls time steps for a 1 ms stimulation pulse (C, D).

20 mV 2 ms

extracellular stimulation intracellular stimulation pulse duration 10 ms pulse duration 5 ms

A

C

D

B

I =-7.1 µAel I =-7.0 µAel I =0.27 nAel I =0.28 nAel AIS soma soma AIS node of Ranvier node of Ranvier h=1 h=0

Fig. 6.Comparison of threshold and subthreshold membrane voltages as functions of time of the compartments closest to the electrode, computed for Neuron 1. Extracellular cathodic stimulation 50lm above soma (A), AIS and node of Ranvier (C). Corresponding intracellular anodic stimulation at the right side. The dashed lines in A and B show the inactivation gating variable h for subthreshold cases; note its quicker decay and the lower minimum for extracellular stimulation. During pulse application there is a monotonous voltage increase in all intracellularly stimulated cases, whereas extracellular stimulation causes a maximum in every of the subthreshold voltage curves, most pronounced in the node of Ranvier example. As shown in A, the maximum is a consequence of the decreasing inactivation variable h. Pulses loose their stimulating properties as soon as the maximum is reached. Same scaling in all graphs.

(8)

463 depolarized and hyperpolarized regions results in a more 464 constant increase of the voltage profile which becomes

465 comparable to that of the intracellular case.

466 Strength–duration curves for electrode positions

467 above the soma and a node of Ranvier underline the

468 association of increasing chronaxies with increasing elec-469 trode distance (Fig. 8). Chronaxie was determined by the

470 intersection of strength–duration curves with the horizon-471 tal line 2⁄rheobase, with electrode currents normalized to

472 rheobase (Fig. 8). The lower part of the figure uses the 473 chronaxie values of the shown strength–duration curves

474 for electrode distances of 50, 100, 200 and 300

l

m to

475 quantitatively describe chronaxie versus distance rela-476 tionships for electrodes above the third node of Ranvier

477 and the soma. For the extracellular stimulation cases, 478 when increasing electrode distance, different functional

479 neural parts become excited. For this reason, and more 480 remarkably for the soma case, the curves containing the

481 circles deviate from the direction pointed by the arrows

482 (Fig. 8, bottom). Note the quite large difference between 483 intra- and extracellular chronaxies.

484 A shift of the spike initiation zone by electrode dis-485 tance increase is analyzed in the next example with two

486 electrode positions at zel= 50

l

m and zel= 200

l

m

487 above the center of the naked axon of Neuron 1

488 (Fig. 9). In contrast to the 50

l

m case (A) the depolarized 489 region predicted by the activating function with the

70-de-490 gree angle (Fig. 4) exceeds the 200

l

m long naked axon

491

for zel= 200

l

m. There is a change in the sign of the

492

slopes at the onset of the 200

l

s stimulus within

493

the naked axon region in (A), whereas for zel= 200

l

m

494

the slopes are always positive in the larger region marked

495

by the gray rectangle indicating the activated region

496

(Fig. 9B). In case B, the activating function values at the

497

border between AIS and the naked axon are essentially

498

smaller than below the electrode. However, the high

so-499

dium channel density of AIS supports spike initiation at

500

the naked axon, acting as a favorable neighboring

com-501

partment. The activating function values for the AIS are

502

very similar to those values for the naked axon close to

503

it, but its position next to the soma results unfavorable.

504

The convenient neighboring compartments together with

505

axial current flow are crucial components for excitation

506

during intra- and extracellular stimulation.

507

In the next section we show that in some cases axial

508

current flow initiates spikes rather far away from the

stim-509

ulating electrode when long pulses are applied whereas

510

for short pulses spike initiation is close to the electrode.

511

By systematic evaluation and analysis of computer

simu-512

lations we discovered this phenomenon with remarkable

513

consequences for strength–duration curves.

514

Deviations from classical strength–duration curves

515

In order to smoothen recording errors, strength–duration

516

curves are usually fitted by one of the classical

electrode 50µm above soma Iel=-8.6 µA

electrode 200µm above soma Iel=-55 µA

A

B

intracellular soma stimulation Iel=1.0 nA 250 µm 500 µs 10 mV 24% 25% 56% 100% 100% 100%

C

Fig. 7.Temporal evolution of transmembrane voltage profiles in 100ls time steps during a 1 ms pulse at threshold stimulation above the soma (left) and soma membrane voltage as a function of pulse time (right). In case A, 56% of the transmembrane voltage is already reached after the first 100ls. Increasing the electrode distance from zel= 50lm (A) to zel= 200lm (B) causes a similar and consistent voltage increase as for intracellular stimulation. For both cases approx. 25% of the transmembrane voltages are reached after the first 100ls (C) Left figures A and C are replications ofFig. 5C and D.

(9)

517 approaches, the formulas of Weiss and Lapicque (Fig. 3).

518 However, for a constant electrode position our computer

519 simulations demonstrate that deviations from these

520 curves have to be expected when spikes are initiated at 521 different functional parts for short and long pulses.

Intra-522 cellular stimulation of the main dendrite of Neuron 2

re-523 sults in a composite strength–duration curve. This curve 524 consists of two segments of ‘classical curves’ that are

525 connected with a transition part (thick red curve marked 526 as DEND in the lower part ofFig. 1B, the case of

intracel-527 lular stimulation).

528 The occurring phenomena for this main dendrite and

529 other neural structures are explained extensively in 530 Fig. 10. For short pulses, spikes are generated within

531 the dendrite at the site of current injection (green curve

532 inFig. 10A). As expected, but not shown, the spike bifur-533 cates there with one part conducted toward the axon and

534 the other toward peripheral dendritic regions. For a longer 535 stimulation pulse of 5 ms, during the first part of

stimula-536 tion spatial transmembrane voltage distributions similar 537 to the green curve are produced at subthreshold intensity,

538 but with smaller amplitudes (e.g., the two thick blue lines

539 inFig. 10A). During a 5 ms pulse just above threshold 540 intensity, the peak value at this electrode position is not

541 strong enough to produce a dendritic spike, but axial cur-542 rent flow into the axon causes enough sodium current via

543 the low threshold sodium channels Nav1.6 in AIS. As a 544 result a spike initiates at the beginning of the thin part of

545 the axon (t = 5.7 ms, lower thin blue line in Fig. 10A).

546

Thereafter, this spike bifurcates in an asymmetric way

be-547

cause of asymmetric cell properties.

548

In the presented examples of extracellular stimulation,

549

the effect of combined strength–duration curves is difficult

550

to observe for cathodic currents, although the

phenome-551

non is present in a weak form for the cases SOMA and

552

DEND (upper part of Fig. 1B). In contrast, combined

553

strength–duration curves occurs clearly for anodic pulses

554

(Fig. 10B), especially for an electrode above the soma or

555

a node of Ranvier. In contrast to cathodic stimulation,

556

anodic stimulation generates a center of hyperpolarization

557

at a region of the neuron closest to the electrode, and the

558

side lobes of the activating function (marked inFig. 4by

559

green arrows) became the driving forces for excitation.

1 electr ode 50µm 100µm 200µm 300µm

abo ve node of Ran vier

electr ode 50µm 100µm 200µm 300µm above soma 2 3

curr ent inject ed in the soma in a node of Ran vier 4 rheobase*

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 ms Fig. 8.Strength–duration curves for Neuron 1 normalized to rheo-base for electrodes above the soma (solid lines) and the third node of Ranvier (broken lines). Chronaxies marked by vertical arrows increase with electrode–cell axis distance toward the value for inside stimulation (marked by thick lines) which is 1.1 ms for the third node of Ranvier (comp.Fig. 1B) and 2.6 ms for the soma. The circles in the lower part of the figure help to identify the strength–duration curves. The curves containing the circles show a trend toward larger chronaxies with increasing electrode distances.

200µm dendrite unmyel. axon myelin. axon N.R. N.R. N.R. N.R. soma AIS naked axon axon hillock electrode activated region activated region N.R. dendrite unmyel. axon myelin. axon N.R. N.R. N.R. N.R. soma AIS naked axon axon hillock N.R. electrode 70° 100mV 200µs

STIMULUS Iel =-15µA (A) Iel =-150µA (B)

B

A

Fig. 9. Stimulation of Neuron 1 by a microelectrode 50lm (A) and 200lm (B) above the center of the naked axon. Every line on the right side shows the membrane voltage of one compartment in accordance with the cell geometry (left). Short compartment lengths in the unmyelinated axon are because of numerical reasons. Thick gray lines represent the membrane voltage where spikes are initiated which is at the center of the naked axon corresponding to the electrode position in A but is shifted into AIS in B.

(10)

560 Therefore, the first reaction describing the membrane 561 voltage after 10

l

s of anodic stimulation for an electrode

562 at 50

l

m above the soma has two positive peaks

563 (Fig. 10C, black line). Short pulses cause spike initiation

564 at the distal end of the AIS (Fig. 10C, green curves),

565 whereas spikes arise at the centre of the thin axonal

seg-566 ment for long pulses (Fig. 10C, blue curves). A similar

567 shift phenomenon occurs when the electrode is above

568

the third node of Ranvier (Fig. 10D): spikes are initiated

569

at the distal end of the thin segment and at a position

clo-570

ser to its center by short and long pulses, respectively.

571 DISCUSSION

572

The time constant of a patch of cell membrane is defined

573

as

s

= R C, with capacity C and resistance R. Chronaxie

500 1000 µm -40 0 40mV short pulse ( Iel=20nA t=50µs 50µs) long pulse (5ms) Iel=0.8nA t=6.18ms t=5.7ms t=5 ms x x x x soma AIS main dendrite thin axon myelinated axon current injection membrane voltage 1 10 100 1000 SOMA AIS DEND NODE OF RANVIER NAKED AXON µA 0.01 0.1 1 10 ms PULSE DURATION ELECTRODE CURRENT 50µs pulse Iel=60µA 5ms pulse I=el13.3µA 500 1000 µm x x x x soma AIS main dendrite thin axon myelinated axon membrane voltage -100 0 50 mV t=4.66ms t=200µs t=50µs t=5ms 5ms pulse I=11.4µAel 50µs pulse Iel=80µA -50 50 mV t=5.14ms t=50µs t=10µs t=300µs t=5ms membrane voltage electrode -250 -150 500 1000 µm x x x x soma AIS main dendrite thin axon myelinated axon

A

B

C

D

C

D

ANODIC STIMULATION ANODIC STIMULATION ABOVE SOMA ANODIC STIMULATION ABOVE N.R. INTRACELLULAR STIMULATION t=1 ms electrode

Fig. 10. Spike initiation sites change for short and long current pulses. Dashed vertical lines indicate electrode position, end of AIS and the beginning of myelination. Membrane voltages profiles are shown for 50ls (green) and 5 ms (blue) pulse stimulation at selected times. (A) Intracellular stimulation of Neuron 2 at the main dendrite. At the end of a 50ls pulse with threshold intensity (20 nA) a dendritic spike (green curve) is generated. Although during a 5 ms pulse just above the 0.79 nA threshold the maximum membrane potential occurs close to the electrode position (thick blue curves), the spike arises at a quite distant position (300lm) at the thin segment of the axon. (B–D) Extracellular anodic stimulation of Neuron 2 for electrode positions as defined inFig. 1A. (B) Electrode positions above the soma (black line) and the third node of Ranvier (blue line) cause combined strength–duration curves as spike initiation site depends on pulse duration in these cases. (C) Stimulus electrode above soma. Thick lines show the situation at the end of the pulses. For better recognition the 10ls response of the short pulse stimulation is shown in black. (D) Stimulus electrode above node of Ranvier.

(11)

574 is about 0.7

s

for such a patch (Ranck, 1975; Reilly, 1992) 575 which is independent of cell geometry as long as

intracel-576 lular axial current flow is negligible. The large variety in

577 membrane resistance is the main reason for specific

578 chronaxie values for different functional unmyelinated cell

579 regions. In contrast to the variance found in membrane

580 resistance, membrane capacity is rather uniform with a

581 value close to 1

l

F/cm2(Cole, 1968). However, myelin-582 ated regions have considerably smaller chronaxies as

583 the total membrane capacity is inversely proportional to 584 the number of myelin sheets.

585 Strength–duration characteristics for external stimula-586 tion of myelinated peripheral axons can be derived with

587 linear methods from the intracellular relationship

588 (Warman et al., 1992). In contrast to peripheral nerve 589 applications, microelectrodes for central nervous system

590 stimulation operate often in or in the vicinity of the soma, 591 where additional effects appear as consequences of the

592 variation of the electrical and geometrical parameters 593 along the neural axis.

594 Here, we have demonstrated a radical switch to short

595 chronaxies when the stimulating electrode position

596 changes from inside to outside the cell. The inside/outside

597 chronaxie ratio with a value close to 6.5 was the largest 598 for the smallest investigated cell–electrode distance of

599 50

l

m (Fig. 8). This ratio depends essentially on the 600 length of the excited zone positioned between

hyperpolar-601 ized regions for the cathodic case. These hyperpolarizing 602 lobes hinder the continuous excitation process seen by

603 intracellular stimulation. Consequently, the inside/outside

604 chronaxie ratio becomes smaller for larger electrode–cell 605 distances. Increasing electrode distance usually will

606 cause the excitation of other functional cell regions with 607 other chronaxic properties. This phenomenon is

espe-608 cially obvious for the electrode position above a soma with

609 highly excitable neighbored axonal regions showing

610 shorter chronaxies (upper part ofFig. 1B), and therefore, 611 the measured chronaxie cannot just reflect the somatic

612 value, but a value influenced by the excited axonal region.

613 This chronaxie mix disturbs the original trend toward 614 chronaxie values for intracellular stimulation as marked

615 by gray arrows in the lower part ofFig. 8.

616 The presented theory predicts the experimentally

617 observed trend to larger chronaxies for increased dis-618 tances (West and Wolstencroft, 1983). An analytical

ap-619 proach is available for non-myelinated nerve and muscle

620 fibers that predicts different chronaxies for intra- and 621 extracellular stimulation and the increase of chronaxie

622 with electrode–fiber distance (Suarez-Antola, 2005).

623 Examples of combined strength–duration curves have

624 been presented where short and long pulses cause spike 625 initiation in different cell regions. Thus, fitting strength–

626 duration data with a single curve according to the classical 627 approaches of Weiss or Lapique may be equivocal. Some

628 combined strength–duration curves resulting from retinal

629 ganglion cell recordings can be found in Fig. 9 ofGerhardt 630 et al. (2011).

631 Our results help explain trends of strength–duration

632 characteristics. For example, Neuron 1 and Neuron 2

633 have comparable strength–duration curves with similar 634 excitability sequences (node-ais-soma-dend) both for

635

intra- and extracellular stimulation (Fig. 1B). However,

636

unexpected differences between both neurons occur,

637

e.g., only Neuron 2 shows a bimodal intracellular

638

strength–duration curve (red thick curve in the lower part

639

ofFig. 1B).

640

It is important to notice that many variables like

synap-641

tic activity (Spruston, 2008; Sjo¨stro¨m et al., 2008),

refrac-642

tory behavior (Miocinovic and Grill, 2004), inhomogeneity

643

in ion channel density (Migliore and Shepherd, 2002;

644

Keren et al., 2009), branching (Manita and Ross, 2009),

645

curvature of axons (Rattay et al., 2000; Iles, 2005), pulse

646

shape (Wongsarnpigoon et al., 2010), electrode

configu-647

rations (Smith and Finley, 1997), the inhomogeneity and

648

anisotropy in the tissue (Roth, 1995), pulse trains and

649

neuromodulation (De Vries et al., 2007; Minassian et al.,

650

2007) can influence recruitment and strength–duration

651

relationships. Nevertheless, by applying careful analysis,

652

the large variance in chronaxie, even within one cell

653

should be helpful for selective stimulation, especially

654

when microelectrodes are used to activate a specific

655

region like the AIS, a part of the dendritic tree or the

mye-656

linated axon. As an example, short pulse widths

selec-657

tively activate cells with their somas close to the

658

electrode in epiretinal stimulation (Behrend et al., 2011).

659

One should be aware that chronaxie is also the time

660

where stimulus pulse duration needs a minimum of

661

energy (Geddes, 2004), an important fact for neural

pros-662

theses. Moreover, charge injection (electrode current

663

times pulse duration) from electrode surface limits neural

664

prosthetic applications. The left (quite) linear part of

dou-665

ble logarithmic strength–duration curves, e.g. inFig. 1B,

666

predicts (nearly) constant charges when thresholds of

667

very short pulses are compared (half pulse duration

668

needs double threshold current). Minimum charge occurs

669

with infinitely short pulse duration. With the approach of

670

Weiss (Fig. 3) the factor for the additional charge costs

671

are pulse duration divided by chronaxie.

672 CONCLUSIONS

673

A recent review and several often cited articles sustain

674

the wrong dogma that chronaxie is the same for

intra-675

and extracelluar stimulation (Borchers et al., 2012; Nowak

676

and Bullier, 1998; Geddes, 2004). On the contrary,

evalu-677

ations of our compartment model of a cortical pyramidal

678

cell demonstrate up to 20 times longer chronaxies for

679

intracellular stimulation than for extracellular stimulation

680

(Fig. 1B). This fact as well as the occurrence of combined

681

strength–duration curves are supported by theoretical

682 investigations. 683 UNCITED REFERENCE 684 Butson et al. (2007). Q6 685 Acknowledgments—We thank Dr. Katarina Leao for meaningful

686 comments in earlier versions of this manuscript. This work was

687 supported by the Austrian Science Fund , Grant No.

21848-688 N13. R.N.L. is supported by a Kjell and Ma¨rta Beijers Foundation

689 Grant.

(12)

690 REFERENCES

691 Behrend MR, Ahuja AK, Humayun MS, Chow RH, Weiland JD (2011) 692 Resolution of the epiretinal prosthesis is not limited by electrode 693 size. IEEE Trans Neural Syst Rehabil Eng 19:436–442. 694 Blair HA (1932) On the measure of excitability. J Gen Physiol 695 16:165–175.

696 Boinagrov D, Loudin J, Palanker D (2010) Strength–duration 697 relationship for extracellular neural stimulation: numerical and 698 analytical models. J Neurophysiol 104:2236–2248.

699 Borchers S, Himmelbach M, Logothetis N, Karnath H (2012) Direct 700 electrical stimulation of human cortex – the gold standard for 701 mapping brain functions? Nature Rev Neurosci 13:63–70. 702 Brindley GS (1955) The site of electrical excitation of the human eye. 703 J Physiol 127:189–200.

704 Butson CR, Cooper SE, Henderson JM, McIntyre CC (2007) Patient-705 specific analysis of the volume of tissue activated during deep 706 brain stimulation. NeuroImage 34:661–670.

707 Carnevale NT, Hines ML The NEURON book. Cambridge, 708 UK: Cambridge University Press; 2006.

709 Cohen ED (2007) Prosthetic interfaces with the visual system: 710 biological issues. J Neural Eng 4:R14–R31.

711 Cole KS Membranes, ions and impulses. Berkley: University of 712 California Press; 1968.

713 De Vries J, De Jongste MJ, Spincemaille G, Staal MJ (2007) Spinal 714 cord stimulation for ischemic heart disease and peripheral 715 vascular disease. Adv Tech Stand Neurosurg 32:63–89. 716 Dowling J (2005) Artificial human vision. Expert Rev Med Devices 717 2:73–85.

718 Ferna´ndez E, Pelayo F, Romero S, Bongard M, Marin C, Alfaro A, 719 Merabet L (2005) Development of a cortical visual 720 neuroprosthesis for the blind: the relevance of neuroplasticity. J 721 Neural Eng 2:R1–R12.

722 Fried SI, Hsueh HA, Werblin FS (2006) A method for generating 723 precise temporal patterns of retinal spiking using prosthetic 724 stimulation. J Neurophysiol 95:970–978.

725 Geddes LA (2004) Accuracy limitations of chronaxie values. IEEE 726 Trans Biomed Eng 51:176–181.

727 Gerhardt M, Groeger G, MacCarthy N (2011) Monopolar vs. bipolar 728 subretinal stimulation – an in vitro study. J Neurosci Methods, in 729Q7 press<http://dx.doi.org/10.1016/j.jneumeth.2011.04.017>. 730 Greenberg RJ, Velte TJ, Humayun MS, Scarlatis GN, Juan E (1999) 731 A computational model of electrical stimulation of the retinal 732 ganglion cell. IEEE Trans Biomed Eng 46:505–514.

733 Hines ML, Morse T, Migliore M, Carnevale NT, Shepherd GM (2004) 734 ModelDB: a database to support computational neuroscience. J 735 Comput Neurosci 17(1):7–11.

736 Hodgkin AL, Huxley AF, Katz B (1952) Measurement of current– 737 voltage relations in the membrane of the giant axon of Loligo. J 738 Physiol 115:424–448.

739 Holsheimer J, Buitenweg JR, Das J, de Sutter P, Manola L, Nuttin B 740 (2011) The effect of pulse width and contact configuration on 741 paresthesia coverage in spinal cord stimulation. Neurosurgery 742 68:1452–1461.

743 Horsager A, Greenwald SH, Weiland JD, Humayun MS, Greenberg 744 RJ, McMahon MJ, Boynton GM, Fine I (2009) Predicting visual 745 sensitivity in retinal prosthesis patients. Invest Ophthalmol Vis Sci 746 50:1483–1491.

747 Hu W, Tian C, Li T, Yang M, Hou H, Shu Y (2009) Distinct 748 contributions of Nav1.6 and Nav1.2 in action potential initiation 749 and backpropagation. Nat Neurosci 12:996–1002.

750 Iles JF (2005) Simple models of stimulation of neurones in the brain 751 by electric fields. Prog Biophys Mol Biol 87:17–31.

752 Jensen RJ, Ziv OR, Rizzo 3rd JF (2005) Thresholds for activation of 753 rabbit retinal ganglion cells with relatively large, extracellular 754 microelectrodes. Invest Ophthalmol Vis Sci 46:1486–1496. 755 Keren N, Bar-Yehuda D, Korngreen A (2009) Experimentally guided 756 modelling of dendritic excitability in rat neocortical pyramidal 757 neurones. J Physiol 587:1413–1437.

758 Lapicque L (1907) Recherches quantitatives sur l’excitation

759 e´lectrique des nerfs traite´e comme une polarisation. J Physiol

760 Pathol Gen 49:620–635.

761 Lapicque L (1929) The chronaxic switching in the nervous system.

762 Science 70:151–154.

763 Leao RN, Leao FN, Walmsley B (2005) Non-random nature of

764 spontaneous mIPSCs in mouse auditory brainstem neurons

765 revealed by recurrence quantification analysis. Proc Biol Sci

766 272:2551–2559.

767 Leao RN, Tan HM, Fisahn A (2009) Kv7/KCNQ channels control

768 action potential phasing of pyramidal neurons during hippocampal

769 gamma oscillations in vitro. J Neurosci 29:13353. Q8

770 Mainen ZF, Joerges J, Hugenard JR, Sejnowski TJ (1995) A model of

771 spike initiation in neocortical pyramidal neurons. Neuron

772 15:1427–1439.

773 Manita S, Ross WN (2009) Synaptic activation and membrane

774 potential changes modulate the frequency of spontaneous

775 elementary Ca2+ release events in the dendrites of pyramidal

776 neurons. J Neurosci 29:7833–7845.

777 Migliore M, Shepherd GM (2002) Emerging rules for the distributions

778 of active dendritic conductances. Nat Rev Neurosci 3:362–370.

779 Minassian K, Persy I, Rattay F, Pinter MM, Kern H, Dimitrijevic MR

780 (2007) Human lumbar cord circuitries can be activated by extrinsic

781 tonic input to generate locomotor-like activity. Hum Mov Sci

782 26:275–295.

783 Miocinovic S, Grill WM (2004) Sensitivity of temporal excitation

784 properties to the neuronal element activated by extracellular

785 stimulation. J Neurosci Methods 132:91–99.

786 Nowak LG, Bullier J (1998) Axons, but not cell bodies, are activated

787 by electrical stimulation in cortical gray matter. I: Evidence from

788 chronaxie measurements. Exp Brain Res 118:477–488.

789 Palmer LM, Stuart GJ (2006) Site of action potential initiation in layer

790 5 pyramidal neurons. J Neurosci 26:1854–1863.

791 Pezaris JS, Reid RC (2009) Simulations of electrode placement for a

792 thalamic visual prosthesis. IEEE Trans Biomed Eng 56:172–178.

793 Ranck JB (1975) Which elements are excited in electrical stimulation

794 of mammalian central nervous system: a review. Brain Res

795 98:417–440.

796 Rattay F (1986) Analysis of models for external stimulation of axons.

797 IEEE Trans Biomed Eng 33:974–977.

798 Rattay F (1987) Ways to approximate current–distance relations for

799 electrically stimulated fibers. J Theor Biol 125:339–349.

800 Rattay F Electrical nerve stimulation: theory, experiments and

801 applications. Vienna: Springer; 1990.

802 Rattay F (1999) The basic mechanism for the electrical stimulation of

803 the nervous system. Neuroscience 89:335–346.

804 Rattay F, Minassian K, Dimitrijevic MR (2000) Epidural electrical

805 stimulation of posterior structures of the human lumbosacral cord.

806 2: Quantitative analysis by computer modeling. Spinal Cord

807 38:473–489.

808 Rattay F, Resatz S (2004) Effective electrode configuration for

809 selective stimulation with inner eye prostheses. IEEE Trans

810 Biomed Eng 51:1659–1664.

811 Rattay F, Wenger C (2010) Which elements of the mammalian

812 central nervous system are excited by low current stimulation with

813 microelectrodes? Neuroscience 170:399–407.

814 Reilly JP Electrical stimulation and electropathology. Cambridge Q9

815 University Press; 1992.

816 Roth BJ (1995) A mathematical model of make and break electrical

817 stimulation of cardiac tissue by a unipolar anode or cathode. IEEE

818 Trans Biomed Eng 42:1174–1184.

819 Rushton WAH (1927) The effect upon the threshold for nervous

820 excitation of the length of nerve exposed, and the angle between

821 current and nerve. J Physiol 63:357–377.

822 Sekirnjak C, Hottowy P, Sher A, Dabrowski W, Litke AM, Chichilnisky

823 EJ (2008) High-resolution electrical stimulation of primate retina

824 for epiretinal implant design. J Neurosci 28:4446–4456.

825 Shepherd RK, Hardie NA, Baxi JH (2001) Electrical stimulation of the

826 auditory nerve: single neuron strength–duration functions in

827 deafened animals. Ann Biomed Eng 29:195–201.

(13)

828 Sjo¨stro¨m PJ, Rancz EA, Roth A, Ha¨usser M (2008) Dendritic 829 excitability and synaptic plasticity. Physiol Rev 88:769–840. 830 Smith DW, Finley CC (1997) Effects of electrode configuration on 831 psychophysical strength–duration functions for single biphasic 832 electrical stimuli in cats. J Acoust Soc Am 102:2228–2237. 833 Spruston N (2008) Pyramidal neurons: dendritic structure and 834 synaptic integration. Nat Rev Neurosci 9:206–221.

835 Stuart G, Spruston N, Sakmann B, Ha¨usser M (1997) Action potential 836 initiation and backpropagation in neurons of the mammalian CNS. 837 Trends Neurosci 20:125–131.

838 Suarez-Antola RE (2005) The time constants for the electric 839 stimulation of nerve and muscle fibers by point electrodes. In: 840 Proceeding of the 2nd IEEE EMBS conference on neural 841Q10 engineering, pp. 652–655.

842 Tehovnik EJ, Slocum WM, Smirnakis SM, Tolias AS (2009) 843 Microstimulation of visual cortex to restore vision. Prog Brain 844 Res 175:347–375.

845 Voorhees CR, Voorhees 3rd WD, Geddes LA, Bourland JD, Hinds M 846 (1992) The chronaxie for myocardium and motor nerve in the dog

847 with chest-surface electrodes. IEEE Trans Biomed Eng

848 39:624–628.

849 Warman EN, Grill WM, Durand D (1992) Modeling the effects of

850 electric fields on nerve fibers: determination of excitation

851 thresholds. IEEE Trans Biomed Eng 39:1244–1254.

852 Weiss G (1901) Sur la possibilite´ de rendre comparables entre eux

853 les appareils servant a l’excitation e´lectrique. Arch Ital Biol

854 35:413–446.

855 West DC, Wolstencroft JH (1983) Strength–duration characteristics

856 of myelinated and non-myelinated bulbospinal axons in the cat

857 spinal cord. J Physiol 337:37–50.

858 Wongsarnpigoon A, Woock JP, Grill WM (2010) Efficiency analysis of

859 waveform shape for electrical excitation of nerve fibers. IEEE

860 Trans Neural Syst Rehabil Eng 18:319–328.

861 Yu Y, Shu Y, McCormick DA (2008) Cortical action potential

862 backpropagation explains spike threshold variability and

rapid-863 onset kinetics. J Neurosci 28:7260–7272.

864 Zrenner E (2002) Will retinal implants restore vision? Science

865 295:1022–1025.

866

867 (Accepted 5 April 2012)

Referências

Documentos relacionados

The first tests using the optical fiber setup – Laser #1 (described above on section 2.4.1) - show that both the control solution and the gold nanoparticle so- lution are affected

In 1900 he attends the Unversal Exhibition that took place in Paris After a brief passage as chief of a section of Coimbra’s Public Works Direction, he is

Desta forma, o presente estudo objetivou analisar quali-quantitativamente os resíduos sólidos gerados pelas empresas de galvanoplastia do município, bem como avaliar as

Dentre os elencados, dois deles versam sobre mecanismos incentivados pelo Ministério do Meio Ambiente, como a Agenda Ambiental da Administração Pública (A3P) e a

A discussão e o entendimento dos resultados foram realizados através de levantamento bibliográfico de dados sobre a implementação da Política Nacional de Resíduos Sólidos

Com base no que já foi mencionado a respeito de como a empresa em estudo realiza seu fluxo logístico reverso, é possível destacar algumas vantagens proporcionadas pela

Para isso, este trabalho objetiva verificação da gestão dos órgãos municipais quanto aos resíduos sólidos com ênfase nos resíduos pneumáticos, mensurar o quantitativo

Por sua vez, ROCHA FILHO (2012), em Arcoverde-PE, avaliou o efeito de genótipos de palma forrageira resistentes à cochonilha do carmim (miúda, IPA Sertânia e orelha de