Articles in PresS. J Neurophysiol (October 1, 2014). doi:10.1152/jn.00914.2013

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45

Generation of field potentials and modulation of their dynamics through volume integration of cortical activity Yoshinao Kajikawa1 & Charles E. Schroeder1,2 1. Cognitive Neuroscience and Schizophrenia Program, Nathan S. Kline Institute for Psychiatric Research, Orangeburg, NY. 2. Department of Psychiatry, Columbia University College of Physicians and Surgeons, New York, NY Running Head: Generation

and modulation of cortical field potentials

Corresponding author: Yoshinao Kajikawa Nathan Kline Institute 140 Old Orangeburg Rd Orangeburg, NY10965 Phones: +1 (845) 398-6630 Fax: +1 (845) 398-6545 E-mail: [email protected] Abstract: Field potentials (FPs) recorded within the brain, often called “local field potentials” (LFPs), are useful measures of net synaptic activity in a neuronal ensemble. However, due to volume conduction, FPs spread beyond regions of underlying synaptic activity, and thus, an “LFP” signal may not accurately reflect the temporal patterns of synaptic activity in the immediately surrounding neuron population. To better understand the physiological processes reflected in FPs, we explored the relationship between the FP and its membrane current generators using current source density (CSD) analysis in conjunction with a volume conductor model. The model provides a quantitative description of the spatiotemporal summation of immediate local and more distant membrane currents to produce the FP. By applying the model to FPs in the macaque auditory cortex, we investigate a critical issue that has broad implications for FP research. We show that FP responses in particular cortical layers are differentially susceptible to activity in other layers. Activity in the supragranular layers has the strongest contribution to FPs in other cortical layers, and infragranular FPs were most susceptible to contributions from other layers. In order to define the physiological processes generating FPs recorded in loci of relatively weak synaptic activity, stronger synaptic events in the vicinity have to be taken into account. While outlining limitations and caveats inherent to FP measurements, our results also suggest specific peak and frequency band components of FPs can be related to activity in specific cortical layers. These results may help improving the interpretability of FPs. Keywords: Field potential, Current source density, Volume conduction, LFP.

Copyright © 2014 by the American Physiological Society.

46

Abstract

47 48

Field potentials (FPs) recorded within the brain, often called “local field potentials” (LFPs), are

49

useful measures of net synaptic activity in a neuronal ensemble. However, due to volume

50

conduction, FPs spread beyond regions of underlying synaptic activity, and thus, an “LFP” signal

51

may not accurately reflect the temporal patterns of synaptic activity in the immediately

52

surrounding neuron population. To better understand the physiological processes reflected in

53

FPs, we explored the relationship between the FP and its membrane current generators using

54

current source density (CSD) analysis in conjunction with a volume conductor model. The model

55

provides a quantitative description of the spatiotemporal summation of immediate local and more

56

distant membrane currents to produce the FP. By applying the model to FPs in the macaque

57

auditory cortex, we investigate a critical issue that has broad implications for FP research. We

58

show that FP responses in particular cortical layers are differentially susceptible to activity in

59

other layers. Activity in the supragranular layers has the strongest contribution to FPs in other

60

cortical layers, and infragranular FPs were most susceptible to contributions from other layers.

61

In order to define the physiological processes generating FPs recorded in loci of relatively weak

62

synaptic activity, stronger synaptic events in the vicinity have to be taken into account. While

63

outlining limitations and caveats inherent to FP measurements, our results also suggest specific

64

peak and frequency band components of FPs can be related to activity in specific cortical layers.

65

These results may help improving the interpretability of FPs.

66

67

Introduction

68

The field potential (FP) recorded within active neural tissue is an information-rich measure that

69

arises mainly from synaptically-driven transmembrane currents related to excitability

70

fluctuations in ensembles of neurons (Buzsaki et al. 2012; Mitzdorf 1985). Decoding of FPs can

71

extract information about motor commands (Bansal et al. 2012; Ince et al. 2010), sensory stimuli

72

(Belitski et al. 2010; Kayser et al. 2007) and current behavioral and cognitive state (Fries et al.

73

2001; Lakatos et al. 2009, 2013; Scherberger et al. 2005; Steriade et al. 1993). This is true

74

whether the FP is tightly or loosely phase-locked to an identified stimulus [so-called “evoked”

75

and “induced” FPs (Makeig et al. 2004; Shah et al. 2004)], and it also holds for “spontaneous”

76

activity, whose immediate antecedents are obscure (Fukushima et al. 2012; Lakatos et al. 2005).

77

Despite its strengths, the interpretation of FPs is complicated by volume conduction, the

78

spread of electric field generated by current sources in a conductive medium. As FPs are

79

typically recorded with an active recording electrode amplified against a reference signal from a

80

distant electrode, the exact point of origin of the signal is unknown due to volume conduction

81

(Einevoll et al. 2013; Nunez and Srinivasan 2006). The signal could in theory arise anywhere

82

within the conductive medium housing the electrodes. Understanding of the principles of volume

83

conduction allows one to track down the brain sources of electroencephalographic (EEG) and

84

event related potential (ERP) components (Godlove et al. 2011; Schroeder et al. 1995). This

85

same understanding also prompts the use of techniques that isolate relatively local activity like

86

current source density (CSD) (Einevoll et al. 2013; Kajikawa and Schroeder 2011; Mitzdorf

87

1985; Nicholson and Freeman 1975, Pettersen et al. 2006) or the surface Laplacian, a 2-

88

dimensional approximation of the second derivative of potential, used to derive scalp current

89

density (Nunez and Srinivasan 2006; Tenke and Kayser 2012). CSD analysis estimates the

90

laminar profile and magnitude of current sources (net local outward currents) and sinks (net local

91

inward currents) that produce fluctuation of voltage in the conductive medium of the

92

extracellular space (Nicholson and Freeman 1975), and is in most experimental cases explored

93

by analyzing the one-dimensional spatial pattern of FPs (but see Riera et al. 2012). Previous

94

ERP/EEG studies have shown that ERP peak components and EEG frequency components

95

(bands) have distinct scalp topographies and source localizations. The same conditions apply to

96

the laminar-spatial patterns of FP and CSD signals within a cortical area. This is a critical

97

problem for LFP studies, as it limits their ability to attribute experimental effects to specific cell

98

populations (e.g., supragranular pyramidal cells) or input types (e.g., top-down vs. bottom-up)

99

Effects of volume conduction at a macro (>1 mm) scale are recognized as essential to the

100

formation of the spatiotemporal profile of electromagnetic signals that can be measured in ERP

101

and EEG recordings at the scalp (Nunez 1998; Nunez and Srinivasan 2006; Pascual-Marqui

102

1999), but these effects at a micro (< 1 mm) scale are poorly understood (Bédard and Destexhe

103

2011, Buzsaki et al. 2012, Einevoll et al. 2013, Riera et al. 2012). Recent papers on the related

104

topic of the “spatial spread of the LFP” have outlined several variables that likely affect volume

105

conduction beyond the boundaries of active (generator) tissue. Some have emphasized spatial

106

variables, such as the size and shape of the generator substrate that is the population of neurons

107

whose transmembrane currents are reflected in an FP, and the magnitude of the activation

108

(Kajikawa and Schroeder 2011), while others have emphasized temporal variables, such as the

109

synchrony of local cellular activity (Lindén et al. 2011; Reimann et al. 2013). Critically, in most

110

circumstances, synchrony and strength of activation are confounded. On one hand, a given

111

ensemble of neurons is expected to produce a stronger net response, as reflected by a larger FP

112

signal magnitude, when they are activated simultaneously than when they are activated

113

asynchronously (e.g. Reimann et al. 2013). On the other hand, we cannot directly relate the

114

magnitude of FP signal, to the degree of synchronous firing in a given neuron population

115

because, as discussed above, synchrony is not a sole factor that determines spatial spread of the

116

FP. In the present study, we deal with the spread of FP activity within hot spots of highly

117

synchronized activation; i.e., we examine local neuronal responses to preferred stimuli that

118

excite massive and synchronous activation of local neuron ensembles. In this circumstance, high

119

synchrony can be assumed. Interestingly, such strong, synchronous, responses, likely

120

encompassing multiple cortical layers, are reported in many papers. However, investigation of

121

how local (i.e., sub-millimeter scale), well-correlated (i.e. sensory-evoked response) activity

122

patterns of neuron populations in different layers of cerebral cortex, contribute to the

123

construction of a local FP began only recently.

124

A simulation study investigated the contributions of electrical activity of neurons and

125

their compartments to LFP at such scales and suggested the potential role of the active

126

membrane conductance that generates action potentials in shaping the spatiotemporal patterns of

127

FP (Reimann et al. 2013). Regardless of types of membrane conductance and difference in

128

patterns of constituent neurons, transmembrane currents of neuronal population can be translated

129

to CSD features. In sensory systems, sensory stimuli effective in driving strong neuronal firing

130

responses also evoked larger CSD responses (Fishman and Steinschneider 2006, Kajikawa and

131

Schroeder 2011) and larger postsynaptic currents (Tan et al. 2004), that corroborate the close

132

correspondence between CSD components and transmembrane currents.

133

Detailed understanding of the contribution of local activity to local FP is particularly

134

important at this time, as studies are beginning to investigate laminar and sublaminar

135

distributions of FPs in different frequency bands in order to help to sort out influences of

136

feedforward and feedback projections (Buffalo et al. 2011; Spaak et al 2012; Sundberg et al.

137

2012). Importantly in this regard, distinct frequency bands can be dissociated by their differential

138

behavior in both event-related responses (Fries et al. 2011; Khayat et al. 2010; Ray and Maunsell

139

2011; Wilke et al. 2006) and ongoing activity (Lakatos et al. 2005; Magri et al. 2012). However,

140

the spatiotemporal summation by which these microprocesses combine to form a local FP has

141

not been directly investigated in physiological data. In an effort to advance the understanding

142

and physiological interpretation of FP measures this study examined the contribution of volume

143

conduction to the generation of FPs at a sub-millimeter scale. We used a volume conductor

144

model constructed on physical principles to spatially integrate “observed” CSD activity

145

distributed across cortical depths, in order to calculate a “predicted” spatiotemporal pattern of

146

intracortical FPs, and evaluated the similarity between the originally observed and model-derived

147

FP patterns. Our results indicate that volume conduction integrates distributed activity across

148

layers to shape the temporal pattern of an FP in any given layer. Our results also suggest that

149

particularly in the infragranular layers where local FP generation is relatively weak, the form and

150

frequency content of the locally-recorded FP may largely reflect activity in other cortical layers.

151 152

Materials and Methods

153 154

All experimental procedures were approved by the Institutional Animal Care and Use Committee

155

of the Nathan Kline Institute.

156

Subjects. Six macaque monkeys were implanted with a headpost and one or two recording

157

chambers during aseptic surgeries. The chambers were oriented to make penetrations

158

perpendicular to the lateral sulcus.

159

Stimuli. Calibrated binaural tone stimuli were delivered through two free field speakers directed

160

to ears (TDT). Tones ranged from 353.55 Hz to 32 kHz with 0.5 octave intervals (14

161

frequencies) were presented at 60 dB SPL (duration: 100 ms, SOA: 625 ms).

162

Recordings. All recordings were done while monkeys were awake. We used electrodes of linear-

163

arrays of 23 electrical contacts spaced either 100 or 200 m apart (0.3-0.5M at 1.0 kHz) to

164

record signals at different cortical depths. FP (0.1Hz-500Hz) were recorded from all contacts

165

simultaneously and sampled at 2 kHz. As a reference (or indifferent) electrode, a metal pin

166

immersed in saline filling the recording chamber was used. The best frequency (BF) of multiunit

167

activity responses to tones was identified in each recording site (Kajikawa and Schroeder 2011).

168

In the present study, responses to BF tones were analyzed.

169

Current source density. CSD was calculated from FPs recorded from 3 adjacent electrode

170

contacts by the second order finite differences:

171

CSDz   FPz  dz   2FPz   FPz  dz  dz 2 ,

(Eq. 1)

172

as the one dimensional approximation of the second order spatial derivative of the FP (Mitzdorf

173

1985), with an assumption of homogeneous conductivity (see Discussion):

174

2 2 2     2         r    2r     2  2  2     y z  z 2  x 

,

(Eq. 2)

175

The generality of this method of CSD approximation (Eq. 1) has been questioned recently

176

because of generator “edge” effects in rodent barrel cortex (Einevoll et al. 2013), with regard to

177

traditional assumption of “trans-positional invariance” to validate Eq. 1 performed only in

178

vertical dimension across cortex (Mitzdorf 1985). However, the solution proposed by Einevoll

179

and colleagues “iCSD,” (discussed below) was devised for the unique structure of rodent barrel

180

cortex, and does not generalize to the neocortex as a whole. Here, to support the last equality in

181

Eq. (2) as an approximation, “trans-positional invariance” is only a special case of generally

V V V  , z x y

182

required conditions that satisfy the inequality:

183

electrodes (see Discussion). As Figure 2D will show, the condition is usually met when

184

suprathreshold stimulation excites extensive regions of cortex (see also Reimann et al. 2013).

185

Thereby, it is appropriate for a wide range of neocortical applications, including the present case.

186

In the present study we used this standard approximation for compatibility with the vast majority

187

of prior studies.

188

Spectral analyses of FP and CSD. To derive the spectrotemporal patterns and frequency bands of

189

those FP and CSD ( FP(t ) and CSD(t ) respectively), we used the complex Morlet wavelet

190 191

 t '  4

1



e

i0t '



, along the vertical path of array

t' 2

e , 0  6 ,

to calculate the wavelet transforms of those signals (Torrence and Compo 1998):

192

WTLFP (t , f ) 

193

WT CSD(t , f ) 

194

f 

t

FP(t 't )  s

t

*

 t 't   dt ' ,  s 

CSD(t 't )  s

*

 t 't   dt ' ,  s 

(Eq. 3),

0  2  02 , 4s

195

in which t is the sampling interval. The complex Morlet wavelets were scaled to conserve the

196

energy relationship between frequency components. The translation step was 1 ms. Center

197

frequencies of the wavelet stepped from 1 to 256 Hz with 0.1 octave intervals. This dense

198

frequency sampling leaves high redundancy between neighboring frequency bands of the

199

wavelet transform and allows reconstruction of the original signal as (Farge 1992)

200

256 FP(t )  Re   WTFP ( f , t )df  . 1 

201

Power and phase spectrotemporal distributions were derived as the absolute values and phase

202

angles of WTFP (t , f ) and WTCSD (t , f ) .

203

Frequency band signals of FP and CSD were calculated as the sums of 16 center

204

frequencies (delta: 1.0~2.9 Hz, theta: 3.0~8.8 Hz, gamma1: 27.7~81 Hz, gamma2: 83.9~256 Hz)

205

or 8 center frequencies (alpha: 9.1~15.9 Hz, beta: 17.0~25.8 Hz,) of WTFP (t , f ) and WTCSD (t , f )

206

along the frequency axis (Kajikawa and Schroeder 2011). Derived signals were approximately

207

1.6 or 0.8 octaves wide with small overlaps between frequency bands, and were complex-valued,

208

same as WT (t , f ) . The real part of signals (e.g. WT (t, f ) ) was identical to the transform with

209

real part of wavelet (e.g. replacing  t ' in Eq. 2 with  t ' ). As the original signals were of

210

real-valued, real parts of derived signals corresponded to the filtered signals and were used as

211

frequency band-limited signals (Farge 1992). Absolute amplitudes of the complex transforms

212

were used as the power of signals.

213

Volume conductor model. The volume conductor model assumed that CSD at different depths

214

contributed to the FP by volume conduction. Under the assumption of homogeneous tissue

215

conductivity, volume conducted signals decay in proportion to the inverse of distance with no

216

assumption of frequency dependent gain change; this predicted decay is matched by empirical

217

findings (Schroeder et al. 1992; 1998; Kajikawa and Schroeder 2011). We calculated FP at a

218

certain depth d k that could be generated by a given spatial distribution of CSD at every moment

219

as

220

vcFP k , t   A j

CSDd j , t  h  d j  dk 2

2

, h  rh 2d  d

(Eq. 4),

221

in which d  d k 1  d k  is the spatial distance between neighboring contacts of array electrode.

222

The parameter h represents a displacement distance of the center of mass of CSD from the array

223

electrode assuming all CSD components were vertically aligned across cortical depths. The idea

224

behind the displacement is as follows.

225

In general, there are many possible solutions of sources that could generate topographic

226

or depth patterns of FP in solving Eq. (2). In the present case, the only boundary was the line of

227

the array electrodes with depth-distributed signals with surrounding open space. It was not

228

definitive in finding particular CSD solutions as there are many possible configurations of CSD.

229

However, to be consistent with the columnar structure of cortex, one constraint we could

230

presume was that CSD distributed vertically across cortical layers. That reduced the problem to

231

one of determination of the finite horizontal spread. However, again, any horizontal spread of

232

CSD could be replaced by a variety of other CSD to create the same spatial pattern of FP. Thus,

233

it would be an ideal approach to reduce the arbitrariness of horizontal spread and model it as

234

simply as possible, since at this time, we cannot realize more complex (parametric) models in

235

detail or justify them.

236

Responses to suprathreshold pure tones in auditory cortex must distribute as medial-

237

lateral stripes defined by iso-frequency contours. The shapes of stripes were kept similar across

238

cortical layers and horizontally finite, constrained by the anterior-posterior extent of an iso-

239

frequency domain. Any such spatial distributions of sources may be substituted by one or several

240

charge densities located at a horizontal point somewhere inside of responsive zones to generate

241

same field potential patterns. That is equivalent to what is done at a different spatial scale in

242

source estimation for EEG/MEG. The anchoring point around which spread is estimated is single

243

point at the center of mass in a horizontal distribution of sources. Its exact position may remain

244

unknown. Only its distance from electrodes is needed in deriving FP at the electrodes. Thus, we

245

included the horizontal displacement parameter h to represent the horizontally delimited

246

distribution of activity in the simplest way. While such a simplified model can introduce error

247

(uncertainty), it is preferable to a model with more unknown parameters, unless they can be

248

empirically validated.

249

The output of the resultant model becomes equivalent to that generated by a dipole-like

250

substrate (i.e., having a vertical distribution), horizontally displaced at an average distance h

251

from the recording electrode array across depths (Kajikawa and Schroeder 2011; Somogyvari et

252

al. 2012). In Eq. 3, rh 2 d was the sole free parameter that could influence the shape of vcFP (k , t ) .

253

Its values were chosen to maximize the similarity (see below) of spatiotemporal patterns between

254

the calculated and the observed LFP. The parameter h  rh 2d  d may be considered to be the

255

distance to the center of horizontally distributed sources generating FP from the electrodes. Note

256

that, due to the linearity of both wavelet decomposition and volume conduction, the order of

257

calculating them is interchangeable. However, once the power or phase of wavelet-decomposed

258

signals is derived, these values cannot be entered into the volume conductor model due to the

259

non-linearity of power/phase calculations.

260

Within a “column,” vertical profiles of CSD that contributed to the FP at a given depth

261

were derived as follows. The spatiotemporal CSD profiles were weighted along the depth axis by

262

their distances from the origins of CSD with an additional displacement h . Wavelet-derived

263

frequency band signals were derived in a similar manner. The spatiotemporal profiles of CSD

264

power were derived after distance-weighing.

265

Depth distributions of the power of CSD contributions to granular FP responses and

266

phase differences of CSD from granular FP responses were derived as follows. The mean of the

267

CSD power over a150 ms period post-onset of tone was calculated for each CSD and subtracted

268

by mean power during a 150 ms period prior to the tone onset. For data sampled with 200 m

269

intervals, missing data points were linearly interpolated. For each frequency band, such

270

distribution was normalized by the power at the depth of the granular sink. Since depths of

271

cortical layers relative to electrode positions differed between penetration sites, depths of

272

granular sinks were aligned to zero. Median and confidence intervals at each depth were derived

273

after alignment.

274

While the field generated by dipoles decays with the inverse of the square of distances, 2

275

1 r , the volume conductor model treated individual poles of multi-poles as separate monopoles

276

whose field decayed with the inverse of the distance, 1 r , not squared. However, it should be

277

noted that superimposition of electric fields generated by two monopoles of opposite signs

278

decaying with 1 r becomes same as an electric field that decays with 1 r

279

relatively larger than the dimension of the dipole. Therefore, while the model treats CSD features

280

as spatially distributed monopoles, it can generate the field potentials generated by dipoles as

281

well.

282

This implementation was preferred, because recordings in the present study were done within the

283

zones of active generators or dipoles, where the dipole field approximation’s relevance is

284

questionable.

285

Similarity score. To quantify the similarity between temporal patterns of FP and CSD, and the

286

similarity between the spatiotemporal patterns of the observed FP and CSD or model-derived FP,

287

we calculated the similarity score as described below. Temporal patterns of FP, CSD, and model-

288

   derived FP at individual recording channels were expressed as vectors, vFP , vCSD and v vcFP of nT

289

independent samples, where nT = sample duration x sampling rate. Spatiotemporal profiles were

290

expressed as matrices, M FP , M CSD , and M vcFP , whose number of rows corresponded to the

2

at distances

291

number of recording contacts of electrodes - 23 and 21 for FP and CSD, respectively.

292

Spectrotemporal patterns of signals at each channel were expressed similarly in matrices

293

WTFP , WTCSD , and WTvcFP ; whose number of rows corresponded to the number of wavelet

294

center frequencies. Before calculation of the similarity, these vectors and matrices were

295

 normalized by the root of mean squared values (e.g. v

296

compensate for their difference in the signal power and to analyze shapes in isolation. For

297

matrices, the mean was calculated across both rows and columns. Similarity between those

298

patterns was calculated as the inner product between vectors or the Frobenius inner products of

299

matrices after normalization. The similarity value ranged from –1 for completely mirrored

300

shapes of opposite polarities to 1 for identical shapes of same polarities regardless of

301

magnitudes. While similarity could depend on the periods of FP, CSD, or their wavelet

302

transforms, values were derived using signals between -30 and 170 ms from the onset of sound.

303

However, the similarity score is model-free in that it can be evaluated for any models or analysis

304

that creates spatiotemporal patterns as long as it keeps dimensions (the number of depth channels

305

and time bins) constant.

306

Statistics. All statistical tests were done non-parametrically with a criterion level of p = 0.01,

307

except for ad hoc comparisons at p = 0.05. 95% Confidence intervals were derived by bootstrap

308

using the Matlab function, bootstrp, with 1000 re-sampling (Figs. 4, 6A and B).

 

2 mean v or M

 ) to

mean M

2

309 310

Results

311 312

Sensory-evoked FP and CSD signals at a given location in the brain often differ in temporal

313

pattern even though both signals are considered to reflect local synaptic activity. We first

314

elaborate the difference between these signals in both temporal and spectral domains. We then

315

describe how the volume conductor model bridges the gap between the two signals. We further

316

show that the FP integrates local and distant activity in individual frequency bands independently

317

of the spatial distributions of the CSD activity of other frequency bands.

318 319

Differences between co-located FP and CSD signals.

320

First, we show the difference in temporal patterns between FP and CSD. Figure 1 shows

321

examples of FP and co-located CSD responses recorded simultaneously from (i) supragranular

322

(Sg), (ii) granular (Gr), and (iii) infragranular (Ig) layers in an A1 penetration, along with

323

spectrotemporal patterns of signal power (C,D) and phase difference (E). FPs in the different

324

layers share similar patterns starting with onset negativity (Fig. 1A), but after this point

325

waveforms diverge between Sg and other layers. This is not surprising because FP patterns are

326

expected to change systematically as recordings traverse an active FP generator region (Mitzdorf

327

1985, Schroeder et al. 1995) and the strongest CSD activity was found in the Sg layer. The point

328

of interest here, however, is the contrast between co-located FP and CSD signals as it reveals

329

effects of volume conduction. The FP - CSD contrast is described systematically by comparison

330

of their temporal patterns (B), low frequency power plots (C vs. D) and phase differences (E). In

331

Sg layers, both FP and CSD signals maintained negative deflections after the onset until 100 ms.

332

Low frequency power had similar peaks. In Gr layers, though temporal patterns looked similar,

333

the onset negativity occurred earlier in CSD than FP and had a nearly 90 degrees phase

334

difference in high frequency bands. In Ig layer, the larger positive/negative fluctuations in FP

335

were more rapid than the source/sink fluctuations in the CSD signal. Thus, the peak FP power

336

was higher in a higher frequency band than peak CSD power.

337

To assess the dissimilarity between FP and CSD across penetration sites, we quantified

338

the similarity score as an inner product of vectors representing temporal patterns of 2 signals in 3

339

cortical layers (Fig. 1E). The score changes from 1 when FP and CSD had identical temporal

340

patterns to -1 when two patterns were of opposite polarity and mirror images of one another. The

341

scores for 3 pairs in Fig. 1B were 0.85 (i), 0.44 (ii), and -0.08 (iii), respectively. The distributions

342

of the score were skewed towards 1 in all cortical layers. Median scores were 0.81, 0.75, and

343

0.33 for Sg, Gr, and Ig layers respectively (n=130), with significant difference (Friedman’s non-

344

parametric repeated measures ANOVA, 2(df=2, n=116) = 30.4, p < 0.01, excluding penetration

345

sites missing Ig layer). This difference was attributable to the lower scores of Ig layer when

346

compared to other layers (Tukey's HSD test, p < 0.05). Thus, temporal patterns of FP and CSD were

347

more dissimilar in Ig layers than Gr and Sg layers. However, even in Gr and Sg layers, the scores

348

distributed over wide ranges. These results suggested the generality of the differences in

349

spectrotemporal patterns between co-located FP and CSD signals. Below, we show how the

350

discrepancies between the two signals are explained by volume conduction, and evaluated the

351

effect by the similarity score.

352 353 354

Volume conductor model Here, we explain the volume conductor model first, and then fill the gap between FP and

355

CSD by applying the model to experimental data. CSD signals reflect transmembrane current

356

flow patterns that are localized to the spatial extent of neuronal processes (1). The local 1-

357

dimensional CSD profile is customarily derived from a laminar FP profile using Eq. 1 by

358

approximating  2 (r )  qr  , in which  is the tissue conductivity (assumed to be spatially

359

   uniform), (r ) is the FP, and qr  is CSD. FP reflects spatially distributed events qr  through









360

volume conduction. When the boundary condition of ()  0 is met,  2 (r )  qr  can be

361

solved as    1 qr  r   ( r )    dr  . 4  r  r 

362

(Eq. 4)

363

Since auditory cortical FP responses to best frequency tones are several hundreds of microvolts

364

and scalp ERPs centimeters away from the auditory cortex are generally on the order of 1

365

microvolts, we used this solution as an approximate model of a local volume conductor. There is

366

 no constraint on the spatial distribution of qr  within a body. The volume conductor model

367

summates spatially distributed CSD signals after weighing them with the inverse of their

368

distances from the point of measurement as in Eq. (4), which weakens distant CSD to some

369

degree.

370

Figure 2 shows how Eq. (4) governs volume conduction when there are two different FP

371

generators, each consisting of sink and source, operating in isolation (one strong source/sink -

372

Fig. 2A; one weak sink/source - Fig. 2B), or together (Fig, 2C). Sink sand sources modeled here

373

have round shapes delimited to two parallel flat planes spread horizontally (xy, Fig. 2Ai and Bi).

374

FP profiles were calculated based on Eq. 4. Due to symmetry of sink and source shapes around

375

their centers, FP profiles formed in planes that contain x-y centers of sink and source are

376

identical. When there is one FP generator, inversion of FP polarity occurs at the mid z position

377

between sink and source (Fig. 2Aii and Bii). Above and below the inversion, the FP spreads with

378

constant polarity with amplitude peak positions aligned vertically to sink and source (Fig. 2Aiii

379

and Biii). However, when two FP generators co-exist, FP distributions from single generators

380

superimpose, and an FP distribution reflecting an interaction of the two generators emerges.

381

Ventral spread of the negative FP below the inversion of the single strong generator (Fig. 2A) is

382

reduced in the combined case (Fig. 2C) and is interrupted by small positivity due to the lower

383

generator’s source. Compared to the positive peak below the inversion in Fig. 2B, the positive

384

peak at the corresponding z-position is smaller in Fig. 2C. Thus, within the cortex where activity

385

spreads over cortical layers through volume conduction, the FP in any one layer is

386

‘contaminated’ by signals of activity generated in other layers. We utilized the volume conductor

387

model to specify this effect quantitatively.

388

Figure 3 shows how the volume conductor model links sensory-evoked FP and CSD

389

responses. At all penetration sites, we simultaneously recorded FP at 23 depths spanning the

390

superficial to deep layers of auditory cortex, as illustrated by the representative FP profile in Fig.

391

3C, from which the spatiotemporal CSD profile in Fig. 3A was derived. The volume conductor

392

model-derived FP profile for these signals (described below) is shown in Fig. 3B. Fig. 3D

393

illustrates the gradual change in the FP produced by the systematically increasing spatial range of

394

the CSD profile that is incorporated into the volume conductor model to derive Gr layer FP

395

waveform. This is shown in increasing order from top to bottom, starting from a summation of

396

one channel that corresponds to Gr layer CSD itself. While temporal patterns of the observed FP

397

and CSD (bottom red and top green traces in Fig. 3D) differed in their temporal patterns, the FP

398

generated by the model (traces between the top and bottom in Fig. 3D) became successively

399

closer approximations of the observed FP as the spatial range of CSD estimates (incorporated by

400

the model) is increased. In this example, the model broadened the onset negativity and delayed

401

the timing of the following positive peak (asterisks in Fig. 3D) from 38 to 62 ms. This effect

402

indicates that the Gr layer FP was influenced by current generators (indexed by CSD) in other

403

layers. By performing the same model calculation over all depths of FP recording, we calculated

404

a model-derived spatiotemporal FP profile (Fig. 3B), which resembles the profile of the observed

405

FP (Fig. 3C) more closely than it does the CSD profile. Accordingly, the spectrotemporal pattern

406

of the model-derived FP for the Gr layer also bears greater similarity to that of the observed FP

407

than that of the CSD signal (Figs. 3E-G).

408

For the example shown in Fig. 3, the similarity score of the observed FP with the co-

409

located CSD signal was 0.38 and increased to 0.95 when computed using the model-derived FP.

410

Also, across penetration sites (n=130), the similarity over the 150 ms long temporal patterns

411

between the observed FP in Gr layer and the model-derived FP increased as a function of the

412

number of proximal CSD signals (Fig. 4B). Improvement of the score through use of the model

413

waveform was also observed for FP in Sg and Ig layers (Fig. 4A and C).

414

In all layers, the score reached asymptote by integrating signals within less than 1 mm.

415

This suggested that FP responses to the best frequency tones within the auditory cortex are

416

shaped by activity of variable temporal patterns occurring in the gray matter over the range of a

417

millimeter where strong activation is spread. It may be also noted that the model-derived FP in

418

Sg layer, where local CSD activity was stronger than other layers (Fig. 6C), still increased its

419

similarity to the observed FP by integrating activity over several hundreds of micrometers. These

420

results suggested that even at loci where the CSD signals were strong, the effect of volume

421

conduction would not be negligible due to activity in their vicinity.

422

Notably, the FPs in Ig layers (which are farthest from Sg layers) had the lowest similarity

423

scores. FPs in these layers required the widening of spatial coverage in the model to include CSD

424

in the Sg layers in order for the similarity score to asymptote. We estimated the channel

425

increments needed for the scores to reach 90% of the net changes, for all 3 layers in each

426

penetration site (Fig 4D). Those channel increments differed significantly between layers

427

(Friedman’s nonparametric repeated-measures ANOVA, 2(df=2, n=116) = 38.2, p < 0.01,

428

excluding sites missing Ig layer data). This difference was attributable to the channel increments

429

needed significantly more in Ig layer than other layers (Tukey's HSD test, p < 0.05). Thus, in

430

order to define the physiological processes generating FPs recorded in loci of relatively weak

431

synaptic activity (e.g., the Ig layers), stronger synaptic events in the vicinity (e.g., those in the Sg

432

layers) have to be taken into account.

433 434 435

Contributions of different frequency bands to volume conducted signals Below, we expand the consideration of volume conduction to the relationships between

436

components of FP and CSD. Sensory cortical responses are not only distributed across layers but

437

 also transient and of broad spectral content. A CSD signal qr , t  can be considered as a linear

438

summation of the CSD frequency bands q f r , t  as q r , t    q f r , t  . By incorporating this into







f

439

440

Eq. 4, FP signals decomposed into frequency bands in the volume conductor model is  1 (r , t )  4 

   q r  r , t  f

f

  r  r

 dr  .

441

Due to linearity of both volume conduction for spatially distributed CSD signals and spectral

442

transforms for frequency bands in each CSD signal, those operations are interchangeable. Thus,

443

the FP due to volume conduction can be derived separately for different frequency bands as

444

   1 q f r  r , t    f (r , t )  dr  .   4  r  r 

445

Figure 5 illustrates how spatiotemporal patterns of CSD activity in the different

446

frequency bands contribute to the infragranular FP response. Fig. 5A and B show the

447

spatiotemporal profiles of CSD responses and their power over a wide frequency range (,

448

1~256 Hz) and in five 1.6 octave- (, ,  and ) and 0.8 octave-wide ( and ) frequency

449

bands. Due to division by the distance from the point of interest, the volume conductor model

450

weakens distant CSD before summation. Therefore, actual signals in the Sg layer CSD

451

components were stronger than what is plotted in Fig. 5A and B, regardless of frequency band.

452

In each frequency band, summation of CSD responses weighted inversely by the distance from

453

the Ig layer along depth reflects the Ig layer FP response derived by the volume conductor model

454

(orange traces in Fig. 5C).

455

In the example shown in Fig. 5, the Ig layer CSD response (green in Fig. 5C, ) was

456

negative from 10 to 100 ms after the onset of sound and had a negative peak at 50 ms followed

457

by a positive peak at 130 ms. The FP response (red in Fig. 5C, ) had a positive peak at 50 ms

458

followed by a negative peak around 130 ms. Temporal pattern differences between CSD and FP

459

signals were reflected in opposite polarities essentially throughout the response time course

460

across frequency bands (, ,  and ). Beta band CSD and FP signals also differed in

461

frequency.

462

These temporal pattern discrepancies between CSD and FP signals across frequency

463

bands were due to volume conduction. In most frequency bands, powers of signals in Gr and Sg

464

layers were comparable to or even stronger than the power in the Ig layer (Fig. 5B, , , ,  and

465

). In each band, by spatially summating the CSD signals across depths, the volume conductor

466

model reproduced an FP response that greatly resembled the observed FP, but with a polarity

467

opposite to that of the co-located Ig layer CSD signal (orange in Fig. 5C). Thus, the temporal

468

pattern of the Ig layer FP response reflected not the local generator currents, but rather those in

469

other cortical layers via volume conduction.

470

We quantified the relative contributions of CSD signals across depths to the granular FP

471

for all experiments (penetration sites). In each penetration site, the depth distributions of CSD

472

response power, weighted by the inverse of distances from the Gr layer sink were calculated

473

separately for the same 6 frequency bands as done for Fig. 5. In general, sensory evoked CSD

474

signals are distributed across cortical layers in sensory cortices and have broad frequency content

475

(Kajikawa and Schroeder 2011; Lakatos et al. 2005, Maier et al. 2011).

476

Figure 6A illustrates the laminar distribution of CSD power by frequency band that

477

contributed to the Gr layer FP response. As the power of the Gr layer CSD responses differed

478

significantly between frequency bands (median values from low to high frequency bands: 0.49,

479

0.41, 0.076, 0.032, 0.012 and 0.86x10-3 mV/mm2, Friedman's test, 2(df=5, n=130) = 621, p

Generation of field potentials and modulation of their dynamics through volume integration of cortical activity.

Field potentials (FPs) recorded within the brain, often called "local field potentials" (LFPs), are useful measures of net synaptic activity in a neur...
847KB Sizes 0 Downloads 5 Views