Articles in PresS. J Neurophysiol (January 14, 2015). doi:10.1152/jn.00716.2014

1

Acetylcholine excites neocortical

2

pyramidal neurons via nicotinic receptors

3 4

Abbreviated Title: Nicotinic responses of neocortical pyramidal neurons

5

Tristan Hedrick and Jack Waters*

6 7 8

Department of Physiology, Feinberg School of Medicine,

9

Northwestern University, 303 E Chicago Ave, Chicago IL 60611

10 11 12

* current address and address for correspondence:

13

Jack Waters

14

Allen Institute for Brain Science

15

551 N 34th St

16

Seattle WA 98103

17

tel. 206-548-8439

18

e-mail: [email protected]

19 20

Keywords: acetylcholine, nicotinic acetylcholine receptor, neocortical pyramidal neurons

Copyright © 2015 by the American Physiological Society.

21

ABSTRACT

22

The neuromodulator acetylcholine shapes neocortical function during sensory

23

perception, motor control, arousal, attention, learning and memory. Here we investigate

24

the mechanisms by which ACh affects neocortical pyramidal neurons in adult mice.

25

Stimulation of cholinergic axons activated muscarinic and nicotinic ACh receptors on

26

pyramidal neurons in all cortical layers and in multiple cortical areas. Nicotinic receptor

27

activation evoked short-latency, depolarizing postsynaptic potentials in many pyramidal

28

neurons. Nicotinic receptor-mediated postsynaptic potentials promoted spiking of

29

pyramidal neurons. The duration of the increase in spiking was membrane potential-

30

dependent, with nicotinic receptor activation triggering persistent spiking lasting many

31

seconds in neurons close to threshold. Persistent spiking was blocked by intracellular

32

BAPTA, indicating that nAChR activation evoked persistent spiking via a long-lasting

33

calcium-activated depolarizing current. We compared nicotinic PSPs in primary motor,

34

prefrontal and visual cortices. The laminar pattern of nicotinic excitation was not

35

uniform, but was broadly similar across areas, with stronger modulation in deep than

36

superficial layers. Superimposed on this broad pattern were local differences, with

37

nicotinic PSPs being particularly large and common in layer 5 of M1, but not layer 5 of

38

PFC or V1. Hence, in addition to modulating the excitability of pyramidal neurons in all

39

layers via muscarinic receptors, synaptically-released ACh preferentially increases the

40

activity of deep-layer neocortical pyramidal neurons via nicotinic receptors, thereby

41

adding laminar selectivity to the widespread enhancement of excitability mediated by

42

muscarinic ACh receptors.

43

INTRODUCTION

44

The neuromodulator acetylcholine (ACh) shapes neocortical function during sensory

45

perception (Metherate, 2004; Disney et al., 2007), motor control (Berg et al., 2005),

46

arousal (Steriade, 2004; Jones, 2008), attention (Herrero, 2008; Parikh & Sarter,

47

2008), learning (Kilgard, 2003; Ramanathan et al., 2009) and memory (Winkler et al.,

48

1995). A decline in neocortical ACh has been tied to conditions such as depression

49

(Dislaver, 1986), schizophrenia (Raedler & Tandon, 2006), Alzheimer’s disease

50

(Whitehouse et al., 1982) and Parkinson's disease (Whitehouse et al., 1983).

51

Most cholinergic axons in neocortex arise from nucleus basalis and other basal

52

forebrain nuclei such as substantia innominata (Wainer & Mesulam, 1990). The

53

cholinergic projection from basal forebrain plays a central role in shaping neocortical

54

components of arousal, attention, learning, memory, sensory perception and motor

55

control. For example, stimulation of vibrissal M1 evokes whisker movements that are

56

enhanced by activation of basal forebrain (Berg et al., 2005) and basal forebrain lesions

57

impair motor control (Garbawie & Whishaw, 2003) and motor map rearrangement

58

during motor learning (Conner et al., 2003).

59

ACh affects neocortical networks, in part by modulating the activity of pyramidal

60

neurons. Pyramidal neurons express nicotinic and muscarinic ACh receptors (nAChRs

61

and mAChRs) on their plasma membranes (van der Zee et al., 1992; Mrzljak et al., 1993),

62

but ACh is thought to act on pyramidal neurons primarily via mAChRs. Activation of

63

mAChRs evokes an initial hyperpolarization and subsequent slow depolarization of many

64

cortical pyramidal neurons. The hyperpolarization results from activation of an SK-type

65

potassium current, whereas the slow depolarization has been linked to a number of

66

currents, including M-, AHP- and inward rectifier-type potassium currents and a non-

67

specific cation current (Krnjevic, 1971; McCormick & Prince, 1985, 1986; McCormick &

68

Williamson, 1989; Haj-Dahmane & Andrade, 1996; Delmas & Brown, 2005; Gulledge &

69

Stuart, 2005; Carr & Surmeier, 2007; Zhang & Séguéla, 2010). There are reports of ACh

70

activating nAChRs on pyramidal neurons (Roerig et al., 1997; Chu et al., 2000; Kassam

71

et al., 2008; Zolles et al., 2009; Guillem et al., 2011; Poorthuis et al., 2012; but see also

72

Vidal & Changeux, 1993; Gil et al., 1997; Porter et al., 1999). Few authors have studied

73

nicotinic postsynaptic currents in neocortical pyramidal neurons (Roerig et al., 1997;

74

Chu et al., 2000). Hence the functional roles of nAChRs on pyramidal neurons remain

75

obscure.

76

Here we studied how ACh affects pyramidal neurons, focusing on the role of nAChRs.

77

To drive synaptic release of ACh, we expressed the light-activated protein

78

channelrhodopsin-2 (ChR2) in cholinergic neurons in the basal forebrain, allowing us to

79

selectively stimulate cholinergic axons in neocortex (Kalmbach et al., 2012). In contrast

80

to many previous reports, we find that ACh excites pyramidal neurons via both mAChRs

81

and nAChRs. Activation of nAChRs occurs with short latency, consistent with nAChRs

82

being located at synapses between cholinergic axons and pyramidal neurons, and can

83

evoke persistent spiking via a calcium-activated conductance. Direct nAChR-mediated

84

effects occurred in pyramidal neurons in several neocortical areas and in all neocortical

85

layers, indicating that direct excitation of pyramidal neurons via nAChRs can occur

86

across neocortical layers and areas. However, laminar and regional differences in both

87

the incidence and amplitude of the nAChR-mediated depolarization suggest regional

88

differences in the modulation of neocortical networks by nAChRs.

89

METHODS

90

All experiments and procedures were approved by the Northwestern University

91

Institutional Animal Care and Use Committee (IACUC).

92

Two approaches were employed to selectively express ChR2 in cholinergic neurons:

93

(1) stereotaxic injection of a floxed viral vector into the basal forebrain of ChAT-Cre

94

mice, and (2) crossing ChAT-Cre and floxed ChR2 mouse lines.

95

For experiments using virally-delivered ChR2, we used Tg(ChAT-Cre)60Gsat mice

96

(GENSAT), which express Cre-recombinase on a choline-acetyltransferase (ChAT)

97

promoter, resulting in Cre expression in cholinergic neurons throughout the brain. Into

98

this mouse we injected adeno-associated virus with a double-floxed inverse open reading

99

frame (EF1a-DIO-hChR2(H134R)eYFP, Virus Vector Core, University of North

100

Carolina), which drives expression of ChR2-yellow fluorescent protein (ChR2-YFP) in

101

infected neurons containing Cre. 400 nl of virus was injected into the basal forebrain at

102

postnatal day 21 using stereotaxic coordinates (0.2 mm A-P, 1.7 mm M-L, 4.5 mm D-V).

103

Three weeks after injection, ChR2 is expressed almost exclusively in cholinergic neurons

104

and their axons in neocortex, with no adverse effects (Porter et al., 1999).

105

In some experiments, selective expression of ChR2 in cholinergic neurons was

106

achieved without the use of viral vectors by crossing ChAT-Cre (B6;129S6-

107

Chattm1(cre)Lowl/J, Jax 006410) and floxed ChR2 (129S-Gt(ROSA)26Sortm32(CAG-

108

COP4*H134R/EYFP)Hze/J,

109

ChAT-ChR2(Ai32) mice. Cre+ mice were crossed with ChR2+/+ mice to yield Cre+ ChR2+/-

110

offspring. These offspring were crossed to generate Cre+ ChR2+/+ mice, which were used

111

for experiments. Results from ChAT-ChR2(Ai32) mice were similar to those from viral

112

infections and results from these two approaches were pooled, unless otherwise noted.

Jax 012569, Ai32 (Madisen et al., 2012) mouse lines to produce

113

Somatic and axonal labeling with ChR2-YFP was examined in fixed sections. Tissue

114

was fixed by transcardial perfusion with 4% paraformaldehyde in phosphate buffer and

115

100-200 µm thick coronal sections were cut using a vibrating microtome. In some

116

sections, the YFP signal was enhanced with an anti-GFP primary (ab13970, 1:10,000,

117

Abcam) and a fluorescent secondary antibody (613111, 1:750, Invitrogen). Images were

118

acquired by widefield or 2-photon microscopy. Widefield images were acquired using a

119

GFP filter set and a Hamamatsu Orca-285 camera. 2-photon images were acquired with

120

880 nm illumination from a Coherent Chameleon Ultra II Ti:sapphire laser and a 505-

121

545 nm emission filter.

122

Whole-cell recordings were obtained from pyramidal neurons in 300 µm-thick acute

123

parasagittal slices of primary motor cortex and coronal slices of prefrontal and visual

124

cortices from postnatal day 33-61 mice of either sex. In virally-infected mice, recordings

125

were obtained ~3 weeks after viral injection. Slices were prepared in ice-cold artificial

126

cerebro-spinal fluid (ACSF; in mM): 125 NaCl, 2.5 KCl, 1.25 NaH2PO4, 20 NaHCO3, 5

127

HEPES, 25 Glucose, 2 CaCl2, 1 MgCl2, pH 7.3, oxygenated with 95% O2/5% CO2. Whole-

128

cell recording pipettes were 4–8 MΩ when filled with intracellular solution: 135 mM K

129

gluconate, 4 mM KCl, 10 mM HEPES, 10 mM Na2-phosphocreatine, 4 mM Mg-ATP, 0.3

130

mM Na2-GTP, 0.2% (w/v) biocytin, 10 µM alexa 594, pH 7.3.

131

Recordings were obtained at 35-37 °C with a feedback circuit and temperature

132

controller (TC-324B; Warner Instruments, Hamden, CT, USA). In experiments using

133

calcium-free ACSF, 2 mM CaCl2 was replaced with 4 mM MgCl2 to give a final MgCl2

134

concentration of 5 mM. ChR2 was activated by wide field illumination through a

135

microscope objective (Olympus, x20/0.95 NA or x40/0.8 NA) using a blue light-emitting

136

diode (LED; Thorlabs LEDC5 and LEDD1 or DC2100 driver). The maximum steady-state

137

intensity was 20 mW/mm2. In some experiments (see figure 4E-G), illumination was

138

restricted by closure or partial closure of the fluorescence field stop. The illumination

139

area was measured offline by bleaching a thin, immobilized film of fluorescence on a

140

microscope slide. Closure of the field stop had no effect on illumination intensity per unit

141

area. In restricted illumination experiments, the soma was positioned in the center of the

142

field of illumination unless noted otherwise.

143

The amplitudes of voltage responses were calculated by subtracting the average

144

baseline membrane potential for ≥1 s before illumination from the peak of the response.

145

Voltage responses which failed to exceed 3 standard deviations of the baseline

146

membrane potential were assigned an amplitude of 0 mV. During brief bursts of nAChR

147

PSPs, the peak amplitude was calculated by subtracting the pre-burst membrane

148

potential from the most depolarized potential during the burst.

149

Local application of ACh was by pressure ejection of 100 µM ACh in ACSF from a

150

glass pipette ~50 µm from the soma, 30 psi, Toohey Spritzer IIe (Toohey Company,

151

Fairfield NJ).

152

Mecamylamine hydrochloride and atropine were obtained from Sigma-Aldrich.

153

Dihydro-β-erythroidine hydrobromide (DHβE), methyllycaconitine citrate,

154

galanthamine hydrobromide and physostigmine hemisulfate were obtained from Tocris

155

Bioscience. NBQX disodium salt, (R)-CPP, gabazine (SR95531), CGP 52432, tetrodotoxin

156

citrate and (R,S)-MCPG were obtained from Ascent Scientific. In some experiments, we

157

used (2R)-amino-5-phosphonopentanoic acid (AP5) to block NMDA receptors, instead of

158

CPP. AP5 was obtained from Tocris.

159 160

Statistical analyses were performed using Graphpad QuickCalcs online tool (http://www.graphpad.com/quickcalcs/) or in Graphpad Instat 3.06 (GraphPad

161

Software Inc., La Jolla, CA). Continuous data (such pharmacology of the nAChR PSP)

162

were analyzed with a two-tailed t-test, Kruskal-Wallis test or Mann-Whitney test and

163

categorical data with Fisher's exact test.

164

The kinetics and latency of nAChR PSPs were measured by fitting a sum of two

165

exponentials to the mean of ten trials. In each trial the PSP was evoked with a single 2

166

ms (or occasionally 5 ms) blue light illumination. The signal-to-noise ratio of our

167

recordings was not sufficient to permit accurate measurement of the timing of the initial

168

rise of the membrane potential of a nAChR PSP, which is smaller and slower than a

169

glutamatergic PSP. To accurately estimate the point of initial rise we therefore measured

170

the time from the start of illumination to 5% of the peak amplitude of the PSP. We would

171

expect this method of latency measurement to overestimate the latency of each PSP by

172

approximately 1 ms and we therefore corrected the latency of each PSP, by back-

173

extrapolating the fit to the resting membrane potential preceding the stimulus.

174

Pyramidal neurons were identified by their somatic shape and spiking pattern and by

175

their large apical dendrites, which were visible by fluorescence microscopy once filled

176

with indicator. Neurons were routinely filled with indicator and their dendrites

177

examined by fluorescence microscopy. Neurons with truncated primary apical dendrites,

178

lacking apical dendrites and with spiking patterns more typical of interneurons (narrow

179

spikes, large after-hyperpolarization) were excluded from further analysis.

180

Pyramidal cells were classified according to the laminar locations of their somata.

181

Laminar borders were based on the Allen Brain Atlas and consistent with laminar

182

variation in opacity of slices, which was visible under brightfield illumination. Distances

183

from the pial surface of the slice were measured perpendicular to the pia. In primary

184

motor cortex, cortical layers were: layer 2/3 150-500 µm, layer 5A 600-750 µm, layer 5B

185

800-1200 µm, layer 6 >1300 µm. In prefrontal cortex, cortical layers were: layer 2/3

186

100-300 µm, layer 5 330-550 µm, layer 6 >550-800 µm (Guillem et al., 2011). In visual

187

cortex, cortical layers were: layer 2/3 100-275 µm, layer 5 450-750 µm, layer 6 >750 µm

188

(Olivas et al., 2012; Petrof et al., 2012).

189

Pyramidal neurons projecting to defined postsynaptic target tissues were labeled

190

using fluorescent microspheres (RetroBeads, Lumafluor). Beads were injected into

191

dorsolateral striatum (0 mm A-P, -2 mm M-L, 2.5 mm D-V), cervical spinal cord

192

(between ~-0.5 mm M-L and the midline, ~1 mm D-V, at the level of the C2 vertebra) or

193

ventral posteromedial thalamus (two sites: 1.6 mm A-P, 1.2 mm M-L, 3.25 mm D-V and

194

1.6 mm A-P, 1.7 mm M-L, 3.5 mm D-V). Recordings were obtained 5-11 days after bead

195

injection from M1 cortex contralateral to striatal and spinal injections and ipsilateral to

196

thalamic injections.

197

RESULTS

198

To stimulate cholinergic axons entering neocortex, we used the blue light-activated

199

membrane protein channelrhodopsin-2 (ChR2; Nagel et al., 2003). We expressed ChR2

200

in cholinergic neurons in the basal forebrain (figure 1A-C), obtaining selectivity for

201

cholinergic neurons with a ChAT-Cre mouse line and floxed virus or by crossing ChAT-

202

Cre and floxed ChR2 mouse lines. We obtained whole-cell recordings from layer 5B

203

pyramidal neurons in acute slices of primary motor cortex containing ChR2-labeled

204

axons (figure 1D). Mean resting potential was -62 ± 1 mV (51 neurons). Widefield

205

illumination of the slice with a blue LED evoked one or more of four voltage responses: a

206

slow depolarization, a hyperpolarization, a medium depolarization, or a fast

207

depolarization (figure 1F). All responses were absent in tetrodotoxin (500 nM TTX, 17

208

neurons) and after removal of extracellular calcium (10 neurons), indicating that all four

209

voltage responses were evoked by spikes, leading to vesicular release (figure 2).

210

We used pharmacology to identify the receptors underlying each of the four voltage

211

responses. The hyperpolarization and slow depolarization were mediated by muscarinic

212

ACh receptors (mAChRs): both hyperpolarization and slow depolarization were

213

eliminated by the mAChR antagonist atropine (1 µM, 140 of 141 neurons, figure 3), but

214

not by the nAChR antagonist mecamylamine (100 µM, 11 of 13 neurons, figure 3) or by

215

glutamate or GABA receptor antagonists (10 µM NBQX and 10 µM CPP; 1 µM gabazine

216

and 3 µM CGP 52432; 14 of 16 neurons).

217

The medium depolarization was mediated by nicotinic ACh receptors (nAChRs). It

218

was eliminated by mecamylamine (100 µM, 12 of 12 neurons, P < 0.05, paired two-tailed

219

t-test), enhanced 73 ± 21% by the nAChR allosteric potentiator and cholinesterase

220

antagonist galanthamine (1 µM, 23 neurons, P < 0.05, paired two-tailed t-test) and 27 ±

221

15% by the ACh esterase inhibitor physostigmine (0.5 µM, 4 neurons, P < 0.05, paired

222

two-tailed t-test), and insensitive to antagonists of mAChRs, GluRs and GABA receptors

223

(5 of 5, 7 of 7 and 8 of 8 neurons, respectively; figure 4). The medium depolarization was

224

unaffected by the α7-nAChR antagonist methyllycaconitine (MLA, 10 nM, 12 of 12

225

neurons) and eliminated by dihydro-β-erythroidine (DHβE, 10 µM, 16 of 16 neurons, P
0.05, Matt-Whitney test, 37 and 22 neurons from virally-infected and

246

ChAT-ChR2(Ai32) mice, respectively). We therefore pooled results from the two

247

techniques.

248

All four postsynaptic responses could occur individually, but most pyramidal

249

neurons exhibited a combination of responses mediated by two or more receptors (figure

250

5B). Hence synaptically-released ACh activates pyramidal neurons via nAChRs and

251

mAChRs, and often via both types of ACh receptor in the same neuron.

252 253

Activation of pyramidal neuron mAChRs by synaptically-released ACh

254

The peak amplitude of the hyperpolarization, evoked by brief illumination (≤10 x 5 ms at

255

20 Hz), was 1.1 ± 0.5 mV and the decay time constant was 774 ± 73 ms (10 neurons). The

256

slow depolarization displayed a peak amplitude of 2.0 ± 0.6 mV (6 neurons) and lasted

257

several seconds. At resting membrane potentials, the hyperpolarization and slow

258

depolarization were rare, occurring in only 7% (8 of 112) and 19% (21 of 112) of layer 5B

259

pyramidal neurons, respectively.

260

Many previous authors have reported activation of pyramidal neurons via mAChRs

261

upon application of ACh to the soma, by pressure ejection from a nearby pipette

262

(McCormick & Prince, 1985; McCormick & Williamson, 1989; Gulledge & Stuart, 2005;

263

Gulledge et al., 2007), with relatively few authors reporting nAChR-mediated

264

depolarization of pyramidal neurons (Roerig et al., 1997; Chu et al., 2000; Kassam et al.,

265

2008; Zolles et al., 2009; Guillem et al., 2011; Poorthuis et al., 2013). Similar to many

266

previous studies, we found that pressure ejection of ACh onto the soma evoked a

267

mAChR-mediated hyperpolarization and slow depolarization (5 of 6 and 6 of 6 neurons,

268

respectively; figure 6A), but no nAChR-mediated depolarization.

269

One possible interpretation of our results might be that synaptically-released ACh

270

activates primarily nAChRs and usually fails to activate mAChRs, whereas pressure

271

ejection onto the soma activates a different population of receptors, primarily mAChRs.

272

mAChR activation modulates primarily potassium conductances (McCormick, 1992) and

273

the reversal potential for potassium is ~-90 mV. mAChR activation may therefore exert

274

little effect on the membrane potential at rest: both mAChR-mediated hyperpolarization

275

and slow depolarization are larger when the neuron is depolarized (McCormick & Prince,

276

1986; Gulledge & Stuart, 2005). To maximize the effects of mAChR activation on the

277

membrane potential and, therefore, the probability that we would observe activation of

278

mAChRs, we depolarized neurons by somatic current injection (figure 6B). Brief stimuli

279

(≤10 x 5 ms illumination at 20 Hz) evoked mAChR-mediated voltage responses in 79% of

280

depolarized layer 5B pyramidal neurons (figure 6C), indicating that synaptically-released

281

ACh activates mAChRs in the majority of layer 5B pyramidal neurons, but that the

282

resulting hyperpolarization or depolarization from rest is often too small to be observed.

283

Previous authors have also reported that ACh, applied by pressure application, more

284

readily evokes postsynaptic mAChR-mediated responses in pyramidal neurons in deep-

285

than in superficial-layers of the neocortex (McCormick & Prince, 1986; McCormick &

286

Williamson, 1989; Gulledge et al., 2007). Similarly, we found that synaptically-released

287

ACh more readily excites pyramidal neurons via mAChRs in deep than superficial layers

288

of M1 cortex, with the proportion of pyramidal neurons that exhibited a mAChR-

289

mediated slow depolarization or hyperpolarization ranging from 53% in layer 2/3 to 91%

290

in layer 6 pyramidal neurons (figure 6C).

291

We conclude that ACh released by a brief burst of cholinergic activity activates

292

mAChRs on the majority of pyramidal neurons throughout primary motor cortex. In

293

comparison to pressure application of ACh, activation of cholinergic synapses with brief

294

bursts of stimuli provides relatively weak activation of mAChRs which often fails to affect

295

the somatic membrane potential at rest.

296 297

ACh release evokes a nicotinic PSP

298

Brief illumination (≤10 x 5 ms at 20 Hz) in mAChR, GluR and GABAR antagonists

299

evoked a nAChR-mediated medium depolarization in almost all (82 of 90) layer 5B

300

pyramidal neurons at rest. We stimulated cholinergic axons with brief bursts of stimuli

301

at 20 Hz, a stimulus designed to mimic the spiking of cholinergic basal forebrain

302

neurons during paradoxical sleep and awake states, when cholinergic basal forebrain

303

neurons spike in bursts, each of ~4-6 spikes at ~25 Hz (Manns et al., 2000; Lee et al.,

304

2005). During such bursts, summation occurred at stimulus frequencies greater than ~5

305

Hz. In our experiments summation may result, in part, from accumulation of calcium

306

entering through ChR2 channels in the presynaptic terminal (Zhang & Oertner, 2007)

307

and may therefore be an artifact of the optogenetic stimulus. Nonetheless, summation

308

permitted us to evoke a substantial depolarization via nAChRs: a burst of stimuli at 20

309

Hz frequently evoked a peak depolarization of 5-10 mV (figure 7A,B).

310

We measured the latency and kinetics of the nAChR-mediated depolarization evoked

311

by a single 2 ms illumination, in 1 µM atropine to inhibit mAChRs and GluR and GABAR

312

antagonists to eliminate any indirect effects. In 15 of 15 neurons, 2 ms illumination

313

evoked reproducible depolarizations (figure 7C), with a mean amplitude of 1.4 ± 0.2 mV

314

(11 neurons). To the mean response we fit a sum of two exponentials (figure 7D) with

315

mean rise and decay time constants of 28 ± 4 ms and 180 ± 23 ms (14 neurons). These

316

relatively slow kinetics are comparable to those of PSPs mediated by α4β2-containing

317

nAChRs in interneurons (Bell et al., 2011), but slower than PSPs mediated by α7-

318

nAChRs (Frazier et al., 1998; Fedorov et al., 2012). These kinetics suggest that the

319

medium depolarization is a PSP mediated by non-α7 nAChRs, consistent with the

320

pharmacology of the medium depolarization, presented above.

321

The latency of the nAChR PSP, measured from the start of illumination, was 5.9 ±

322

0.8 ms (12 neurons). Compared to electrical stimulation, ChR2 drives relatively slow

323

depolarization of the neuronal membrane, typically with rise and decay time constants of

324

at least several milliseconds (Lin et al., 2009; Yizhar et al., 2011), which may be further

325

elongated by the time constant of the membrane. The latency of the fast depolarization

326

was 5.3 ± 0.5 ms (3 neurons). Hence the latency of the nicotinic PSP was only ~0.5 ms

327

longer than that of a monosynaptic glutamatergic PSP, consistent with the medium

328

depolarization being a monosynaptic PSP generated at synapses between cholinergic

329

axons and pyramidal neurons.

330

nAChRs appear to be distributed throughout the dendritic trees of cortical pyramidal

331

neurons (van der Zee et al., 1992; Nakayama et al., 1995) but the locations of cholinergic

332

synapses are unknown. To determine whether nAChR PSPs were evoked primarily by

333

cholinergic synapses in the proximal or distal dendrites of layer 5B pyramidal neurons

334

we measured nAChR PSPs during restricted illumination of the slice (figure 8A,B).

335

Restricting illumination to a radius of less than ~300 µm around the soma was necessary

336

to reduce the amplitude of the nAChR PSP and the amplitude was reduced by 50% when

337

the radius of illumination was ~50 µm (figure 8C, 7 neurons). Illumination of the tuft

338

dendrites failed to evoke a nAChR PSPs at the soma (figure 8B, 3 neurons). Hence the

339

nAChRs which contribute to the somatic depolarization in our experiments are likely to

340

be within 300 µm of the soma and many are probably located in the proximal 50 µm of

341

the apical and basal arbor.

342 343

nAChR activation can evoke persistent spiking

344

We next addressed the functional consequences of nAChR activation. nAChR PSPs

345

increased the spike rate of layer 5B pyramidal neurons in primary motor cortex,

346

depolarized beyond threshold by somatic current injection (figure 9A). The increase was

347

independent of initial spike rate (figure 9B), with 2-5 ms illumination increasing spike

348

rate by 2.2 ± 0.1 Hz (6 neurons) and a brief burst (10 x 5 ms at 20 Hz) increasing spike

349

rate by 3.1 ± 0.2 Hz (4 neurons). Hence, during spiking nAChR activation causes a linear,

350

additive change in spike rate with no change in the slope of the input-output

351

relationship. The elevated spike rate was maintained during repetitive stimulation and

352

declined after cessation of the cholinergic stimulus with a decay time constant of 185 ±

353

21 ms (figure 9C; 2 or 5 ms illumination; 6 neurons), which matches the decay of the

354

underlying nAChR PSP.

355

With the membrane depolarized from rest but subthreshold, cholinergic stimulation

356

evoked persistent spiking. We evoked nAChR PSPs during long step current injections of

357

increasing amplitude until the nAChR PSP evoked one or more spikes. A nAChR PSP

358

reduced rheobase by 11.6 ± 3.7 pA (from 344 ± 26 pA to 337 ± 26 pA, 18 neurons, 5 ms

359

illumination in the presence of atropine). However, under these conditions, a nAChR

360

PSP typically evoked multiple spikes. The spike rate during the first second after

361

stimulus onset was 3.0 ± 0.2 Hz (6 neurons) for a single 2-5 ms illumination and 7.1 ±

362

1.2 Hz for a brief burst (10 x 5 ms at 20 Hz, 7 neurons). Spike rate declined slowly after

363

the last nAChR PSP (figure 10A,B), but spiking typically continued until the holding

364

current was removed, which was up to 4 seconds after the initial spike (figure 10A,B).

365

We defined persistent spiking as spiking that continued for at least 500 ms after the

366

end of the cholinergic stimulus. Using this definition, persistent spiking occurred in

367

every neuron (13 of 13 neurons), but in 6 of 13 neurons persistent spiking occurred on

368

some but not all trials. In trials in which persistent spiking failed to occur, the nAChR

369

PSP evoked only a single spike (mean 1.0 ± 0, 14 trials from 4 neurons, single 2-5 ms

370

illumination), whereas in trials in which persistent spiking occurred the nAChR PSP

371

evoked 7.5 ± 1.6 spikes (10 trials in the same 4 neurons with the same stimulus).

372

Comparing trials with and without persistent spiking, the resting membrane potential at

373

the start of the trial (-62.3 ± 1.9 and -63.2 ± 1.5 mV, respectively), the current injected to

374

depolarize the neuron (209 ± 20 and 268 ± 23 pA, respectively), and the membrane

375

potential immediately before the nAChR PSP (-48.0 ± 2.0 and -48.5 ± 1.8 mV,

376

respectively) and the threshold of the first spike (-35.2 ± 1.9 and -39.1 ± 1.9 mV,

377

respectively) were similar (18 and 14 trials respectively, each P > 0.05, paired 2-tailed t-

378

test). These measurements indicate that trial-to-trial variability in perisomatic

379

membrane potential, input resistance and spike threshold do not account for the

380

variability in persistent spiking, although they do not exclude a role for such changes in

381

the distal dendrite, as a result of ongoing synaptic activity, for example.

382

Persistent spiking required activation of nAChRs on pyramidal neurons, but not

383

glutamate or GABA receptors or mAChRs (figure 11). Persistent spiking occurred in the

384

presence of 10 µM NBQX, 10 µM CPP, 1 µM gabazine, 3 µM CGP and 1-10 µM atropine

385

(13 of 13 neurons). Subsequent addition of 100 µM mecamylamine (in the continued

386

presence of glutamate and GABA receptor antagonists and of atropine) blocked

387

persistent spiking (3 of 3 neurons, figure 11A) but 100 µM MCPG did not (3 of 3

388

neurons).

389

nAChR activation provides more than just the initial depolarization required to

390

initiate persistent spiking since in the absence of cholinergic stimulation, brief

391

depolarization failed to evoke persistent spiking (figure 11C). Furthermore, during

392

persistent spiking evoked by cholinergic stimulation, brief hyperpolarization of the

393

membrane inhibited persistent spiking, only for spiking to resume after the

394

hyperpolarizing pulse (figure 11D). Presumably, an additional depolarizing conductance

395

is activated (or hyperpolarizing conductance deactivated) by nAChR activation or by

396

another receptor co-activated with nAChRs and this additional conductance remains

397

active for several seconds, long after the decay of the nAChR PSP. The spike waveform

398

changed little during persistent spiking (figure 12), suggesting that this additional

399

current is unlikely to arise from one of the sodium or potassium conductances that shape

400

the spike waveform.

401

Persistent spiking was eliminated by 10 mM intracellular BAPTA. With or without

402

intracellular BAPTA, nAChR PSPs evoked 1 or more spikes (figure 11E), but in BAPTA

403

spiking did not continue after the decay of the underlying PSP. Without BAPTA spiking

404

continued until the holding current was removed, with the last spike 2155 ± 261 ms after

405

the start of a single 2-5 ms illumination and 3357 ± 169 ms after a burst of stimuli (10 x 5

406

ms at 20 Hz; 13 neurons). With BAPTA, spiking ended 185 ± 38 ms after a single

407

illumination and 445 ± 115 ms after a burst (3 neurons; single illumination and burst

408

each P < 0.05, unpaired 2-tailed t-test). As a result, cholinergic stimuli evoked fewer

409

spikes with BAPTA (single 5 ms illumination, 1.8 ± 0.1 spikes, 2 neurons; burst 3.3 ± 1.7

410

spikes, 3 neurons) than without BAPTA (single 5 ms illumination, 6.6 ± 0.8 spikes, 7

411

neurons; burst 21.7 ± 0.6 spikes, 7 neurons; single illumination and burst each P < 0.05,

412

unpaired 2-tailed t-test). Hence activation of cholinergic axons evokes persistent spiking

413

that requires activation of nAChRs and a calcium-activated current that does not affect

414

the spike waveform.

415

Our results indicate that ACh has both brief, additive and prolonged, non-linear

416

effects on the spiking of layer 5B pyramidal neurons, with neurons being particularly

417

sensitive to cholinergic activity when their membrane potentials are within ~10 mV of

418

spike threshold, such that a nAChR-mediated increase in spiking can be short-lived or

419

can be more dramatic, evoking spiking that persists for many seconds.

420 421

Cholinergic responses by projection target

422

In primary sensory neocortices, nAChRs are expressed by selected sub-populations of

423

presynaptic terminals. For example, ACh can enhance the transmission of sensory

424

information to neocortex via the activation of nAChRs on thalamocortical terminals in

425

primary somatosensory and visual cortices (Metherate, 2004; Disney et al., 2007; Gil et

426

al., 1997). Neuromodulators can also act selectively on different projection pathways out

427

of neocortex (Gaspar et al., 1995; Beique et al., 2007; Sheets et al., 2011; Avesar &

428

Gulledge, 2012; Gee et al., 2012; Seong & Carter, 2012). For example, in medial

429

prefrontal cortex mAChR activation by ACh has a greater effect on the excitability of

430

layer 5 pyramidal neurons that project to the pons than on neurons that project to

431

contralateral cortex (Dembrow et al., 2010) and nAChR activation evokes larger-

432

amplitude currents from corticothalamic layer 6 pyramidal neurons than from layer 6

433

pyramidal neurons that do not project to thalamus (Kassam et al., 2008). Might ACh

434

differentially modulate the output of motor cortex via expression of nACh receptors in

435

pyramidal neurons that project to some sub-cortical targets, but not pyramidal neurons

436

that project to other target tissues?

437

In motor cortex, descending axons of layer 5 pyramidal neurons project into the

438

pyramidal tract or to the contralateral striatum and these two pathways are mutually

439

exclusive (Shepherd, 2013). Within layer 6, the primary subcortical output is to thalamus

440

and layer 6 pyramidal neurons may therefore be divided into corticothalamic and non-

441

corticothalamic, or intracortical, neurons. We compared the incidence of nAChR- and

442

mAChR-mediated potentials in each of these subpopulations in motor cortex after

443

retrograde labeling of pyramidal neurons by injection of fluorescent beads into spinal

444

cord, contralateral striatum or ipsilateral thalamus (figure 13A). In slices, we identified

445

neurons with different projection targets by the somatic accumulation of fluorescent

446

beads (figure 13B).

447

Synaptically-released ACh frequently evoked nAChR PSPs in all four subpopulations of

448

deep layer M1 pyramidal neurons: corticospinal layer 5 pyramidal neurons;

449

corticostriatal layer 5 pyramidal neurons; corticothalamic layer 6 pyramidal neurons;

450

non-corticothalamic (intracortical) layer 6 pyramidal neurons (figure 13C). The

451

incidence of nAChR PSPs was not different between the two populations of layer 5

452

neurons nor between the two populations of layer 6 neurons (nAChR PSPs in 5 of 10

453

corticospinal layer 5 pyramidal neurons, 8 of 11 corticostriatal layer 5 pyramidal

454

neurons, P > 0.05, Fisher's exact test; nAChR PSPs in 8 of 11 corticothalamic layer 6

455

pyramidal neurons, 6 of 12 non-corticothalamic layer 6 pyramidal neurons, P > 0.05,

456

Fisher's exact test). Hence our experiments revealed no evidence for different

457

probabilities of nAChR PSPs in sub-populations of deep-layer pyramidal neurons with

458

different projection targets. However, the amplitudes of nAChR PSPs were greater in

459

layer 6 pyramidal neurons that projected to thalamus (9.5 ± 2.6 mV, 7 neurons) than in

460

layer 6 neurons that did not project to thalamus (6.0 ± 2.3 mV, 6 neurons; P < 0.05,

461

paired 2-tailed t-test). nAChR-mediated currents are larger in corticothalamic than non-

462

corticothalamic pyramidal neurons in prefrontal cortex (Kassam et al., 2008) and our

463

results suggest that this enhancement of nAChR responses in corticothalamic layer 6

464

pyramidal neurons extends to primary motor cortex.

465

The mAChR-mediated slow depolarization was also common in neurons from all four

466

projection-based populations of deep-layer pyramidal neurons (figure 13C; no difference

467

in incidence of slow depolarization between layer 5 or layer 6 subpopulations, P > 0.05,

468

Fisher's exact test). In contrast, the hyperpolarization displayed differential expression

469

by projection target (figure 13C), occurring often in both layer 5 projection-based

470

populations and in non-corticothalamic layer 6 pyramidal neurons (no difference in

471

incidence of hyperpolarization between layer 5 subpopulations, P > 0.05, Fisher's exact

472

test), but being completely absent from corticothalamic layer 6 pyramidal neurons

473

(different incidence of slow depolarization between layer 6 subpopulations, P < 0.05,

474

Fisher's exact test).

475 476

nAChR PSPs responses across layers and cortical areas

477

In primary motor cortex, nAChRs are expressed by pyramidal neurons throughout the

478

layers of neocortex (van der Zee et al., 1992; Nakayama et al., 1995; Duffy et al., 2009)

479

and cholinergic axons ramify through all layers (Wainer & Mesulam, 1990; Lysakowski et

480

al., 1989). Hence nAChR PSPs might be expected in pyramidal neurons in all layers. To

481

test this hypothesis, we determined the frequency with which synaptically-released ACh

482

evoked nAChR PSPs in pyramidal neurons in layers 2/3, 5, and 6 of primary motor,

483

prefrontal, and visual cortices.

484

In primary motor cortical slices with abundant ChR2-labeled axons in all neocortical

485

layers (figure 14A) and nAChR PSPs in layer 5B pyramidal neurons, cholinergic stimuli

486

(10 x 5 ms at 20 Hz) rarely evoked nAChR PSPs in layer 2/3 pyramidal neurons (4 of 21

487

layer 2/3 neurons; figure 14B; lower probability in layer 2/3 than in layer 5A, layer 5B, or

488

layer 6, P < 0.05 for each, Fisher's exact test). In layers 5A and 5B, nAChR PSPs were

489

common (6 of 8 layer 5A neurons, 82 of 90 layer 5B neurons; figure 14B; no difference in

490

probability between layers 5A and 5B, Fisher's exact test) and in layer 6, nAChR PSPs

491

were evoked in approximately half of neurons (16 of 32 neurons; figure 14B; greater

492

probability in layer 6 than in layer 2/3 and lower than in layer 5B, P < 0.05 for each,

493

Fisher's exact test). Hence in primary motor cortex, nAChR PSPs occur almost

494

exclusively in deep-layer pyramidal neurons.

495

In prefrontal and primary visual cortices (PFC and V1), the laminar pattern of nAChR

496

PSPs was different from that in primary motor cortex. In prefrontal cortex, cholinergic

497

stimuli (10 x 5 ms at 20 Hz) evoked nAChR PSPs in 2 of 6 layer 2/3 pyramidal neurons, 2

498

of 13 layer 5 pyramidal neurons and 5 of 8 layer 6 pyramidal neurons (figure 14B). Hence

499

nAChR PSPs were less common in all three layers of PFC than in layer 5B neurons in M1

500

(P < 0.05 for each layer, Fisher's exact test). Within PFC, nAChR PSPs were more

501

common in layer 6 than in more superficial layers (P < 0.05, Fisher's exact test). As

502

expected from previous studies (Kassam et al., 2008; Poorthuis et al., 2012; Bailey et al.,

503

2010), nAChR PSPs in layer 6 of prefrontal cortex arose from activation of non-α7

504

nAChRs, being unaffected by MLA and eliminated by DHβE (4 of 4 neurons). In visual

505

cortex, cholinergic stimuli (10 x 5 ms at 20 Hz) commonly evoked nAChR PSPs in

506

pyramidal neurons in all cortical layers (5 of 6 layer 2/3 neurons, 9 of 11 layer 5

507

pyramidal neurons, 8 of 9 layer 6 pyramidal neurons; figure 14B; no differences in

508

probability, Fisher's exact test).

509

The amplitudes of nAChR PSPs also differed across cortical layers and areas (figure

510

14C). The largest responses were observed in layer 5B of primary motor cortex

511

(maximum 18.3 mV), but the mean nAChR PSP amplitude was greatest in layer 6

512

pyramidal neurons (M1 6.06 ± 1.21 mV, maximum 16.2 mV; PFC 5.99 ± 2.86 mV,

513

maximum 15.9 mV; V1 2.76 ± 0.74 mV, maximum 5.8 mV; for M1, P < 0.05, Kruskal-

514

Wallis test; L5A different from L5B, L5B different from L6, each P < 0.05, Mann-

515

Whitney test). In all areas, mean peak amplitudes in layers 2-5 were between 1 and 2 mV,

516

with the exception of motor cortex (figure 14C), where responses in layer 5B were larger

517

than in layer 5A (P < 0.05, Mann-Whitney test) and larger than in layer 5 of prefrontal or

518

visual cortices (L5B of M1 14.21 ± 1.14 mV; amplitude larger in M1 L5B than PFC L5 and

519

V1 L5; P < 0.05, Kruskal-Wallis test).

520

To compare the effects of nAChR activation on pyramidal neurons in different layers

521

and areas, we multiplied the probability and amplitude of nAChR PSPs for each layer

522

(figure 14D). This analysis provides a measure of the overall effect of nAChR PSPs on

523

laminar excitability and reveals that, in all three cortical areas, the effects of nAChR

524

activation are greater in deep layers than in superficial layers. To summarize the average

525

effect of ACh via nAChRs across cortical areas, we plot the mean effect by layer by

526

averaging the effects in M1, PFC and V1 (figure 14E). Hence in these cortical areas there

527

is a general pattern of increased effectiveness of nAChR activation in deep layers, on

528

which is superimposed area-specific variations in laminar sensitivity to ACh.

529

Hence our results indicate that ACh, acting via nAChRs, can directly excite pyramidal

530

neurons in many cortical areas and layers. Our experiments reveal differences in the

531

nAChR-mediated responsiveness of pyramidal neurons between cortical areas and

532

neocortical layers. However, these local variations appear to operate within a more

533

general framework which is common to neocortical areas, in which ACh exerts greater

534

nAChR-mediated effects on deep-layer pyramidal neurons.

535

536

DISCUSSION

537

nAChRs on neocortical pyramidal neurons have proven difficult to activate in brain slice

538

preparations, limiting the study of nAChRs. We have overcome this barrier by expressing

539

channelrhodopsin in cholinergic axons and evoking ACh release in neocortical slices.

540

Our results indicate that ACh activates nAChRs on pyramidal neurons in multiple layers

541

and three cortical regions, probably via synapses between cholinergic axons and

542

pyramidal neurons. Hence cholinergic activation of pyramidal neurons via nAChRs is

543

common across neocortex.

544 545

Effects of ACh on pyramidal neurons

546

Several authors have reported nAChR-mediated responses from pyramidal neurons

547

(Roerig et al., 1997; Chu et al., 2000; Kassam et al., 2008; Zolles et al., 2009; Guillem et

548

al., 2011; Poorthuis et al., 2012), but in other studies no such responses were observed

549

(Vidal & Changeux, 1993; Gil et al., 1997; Porter et al., 1999) and in many the actions of

550

ACh were mediated by mAChRs (Krnjevic, 1971; McCormick & Prince, 1986; Haj-

551

Dahmane & Andrade, 1996; Gulledge & Stuart, 2005; Gulledge et al., 2007; McCormick,

552

1992; Schwindt et al., 1988; Giessel & Sabatini, 2010). The absence of nAChR responses

553

is puzzling as nAChRs are expressed in the dendrites of pyramidal neurons (van der Zee

554

et al., 1992; Nakayama et al., 1995; Duffy et al., 2009; Lubin et al., 1999; Levy & Aoki,

555

2002).

556

In most studies, ACh was applied by bulk perfusion or from a nearby pipette, which

557

results in a relatively slow, widespread increase in concentration. nAChR desensitization

558

during ACh application might be significant, reducing nAChR-mediated currents.

559

Furthermore, many mAChRs are located extrasynaptically (Mrzljak et al., 1998;

560

Yamasaki et al., 2010) and might be more strongly activated by applied ACh than by ACh

561

released from cholinergic axons. Our results suggest that ACh application favors

562

mAChR- over nAChR-mediated currents in pyramidal neurons and this weighting may

563

account for the paucity of nAChR-mediated responses in the literature.

564

ACh can act on interneurons in layer 1 of sensorimotor cortex via α7 nAChR-

565

mediated synaptic and non-α7 nAChR-mediated diffuse mechanisms (Bennett et al.,

566

2012) and there is a wider debate on whether ACh acts in neocortex primarily via

567

synaptic contacts or volume transmission (Sarter et al., 2009; Arroyo et al., 2014). The

568

short latency of nAChR-mediated responses in our experiments suggests that ACh forms

569

cholinergic synapses with pyramidal neurons, but via non-α7 nAChRs. We found no

570

evidence for an α7 nAChR-mediated effect or nAChR-mediated actions via volume

571

transmission, but our results do not exclude such responses in pyramidal neurons.

572

Although our recordings were from somata, there are nAChRs throughout the dendritic

573

trees of pyramidal neurons (van der Zee et al., 1992), and it therefore seems likely that

574

there are additional effects of ACh on the dendrites of pyramidal neurons.

575 576

Persistent spiking evoked by nAChRs

577

Our results indicate that nAChR activation can evoke persistent spiking when paired

578

with additional depolarization. Persistent spiking was prevented by intracellular BAPTA,

579

suggesting that a rise in intracellular calcium concentration is also required. Calcium

580

might enter through nAChRs or arise from a secondary source, such as voltage-activated

581

calcium channels or a transmitter co-released by cholinergic axons. Cholinergic axons

582

may co-release glutamate (Manns et al., 2001; Allen et al., 2006; Gritti et al., 2006;

583

Henny & Jones, 2008), but in our experiments persistent spiking was unaffected by

584

glutamate receptor antagonists. Other potential co-transmitters in the basal forebrain

585

include neurotensin, somatostatin, neuropeptide Y and galanin (Koliatsos et al., 1990).

586

nAChR-dependent persistent spiking has been reported in dopaminergic neurons in

587

the substantia nigra pars compacta and subthalamic nucleus (Yamashita & Isa, 2003a,

588

2003b), but not cortical neurons. In entorhinal, perirhinal, cingulate and somatosensory

589

cortices, persistent spiking can be evoked in excitatory neurons, but via mAChRs and a

590

rise in intracellular calcium concentration (Zhang & Séguéla, 2010; Egorov et al., 2002;

591

Navaroli et al., 2011; Rahman & Berger, 2011). Hence nAChR-dependent persistent

592

spiking shares common mechanistic elements with mAChR-dependent persistent

593

spiking, but is initiated by activation of nAChRs, not mAChRs.

594

In spiking neurons, nAChR PSPs evoked a brief and modest increase in spike rate. Why

595

was the effect during ongoing spiking not more prolonged and why was the resulting

596

spike rate typically lower than during nAChR-evoked persistent spiking? Presumably

597

ongoing spiking suppresses the current that underlies persistent spiking or the resulting

598

depolarization, perhaps by shunting the membrane.

599

Previous studies have revealed other mechanisms of prolonged spiking in pyramidal

600

neurons, particularly deep-layer pyramidal neurons. For example, subthreshold DC

601

current injection into the trunk of the apical dendrite can facilitate propagation of a

602

dendritic spike from the distal apical dendrite to the soma, resulting in a burst of spikes

603

(Larkum et al., 2001). Similarly, activation of glutamate receptors in the basal dendrites

604

can evoke a burst of spikes (Milojkovic et al., 2004; Milojkovic et al., 2007). Presumably

605

nAChR-persistent spiking and other mechanisms that can evoke prolonged spiking share

606

some common mechanistic elements and differ important ways. Further experiments

607

will be required to investigate these similarities and differences, but our results add

608

nAChR activation to the collection of identified mechanisms that can evoke prolonged

609

spiking from cortical pyramidal neurons.

610

In summary, nAChR activation increases the excitability of pyramidal neurons. The

611

increase in spiking can be modest and transient or profound and persistent, depending

612

on the membrane potential of the neuron. Hence ongoing synaptic drive to the

613

pyramidal neuron determines the strength and duration of the increase in spiking

614

evoked by ACh.

615 616

Laminar and regional variation in nAChR PSPs

617

α3, α4, α7 and β2 nAChR subunits are located on neocortical pyramidal neuron somata,

618

dendrites and spines (Disney et al., 2007; Nakayama et al., 1995; Duffy et al., 2009;

619

Lubin et al., 1999; Levy & Aoki, 2002; Wevers et al., 1994) and several studies describe

620

responses mediated by α4β2-, α7- and α5-containing nAChRs, the latter probably in

621

heteromeric assembly with α4 and β2 subunits (Kassam et al., 2008; Zolles et al., 2009;

622

Poorthuis et al., 2012).

623

Our results are generally consistent with these previous studies, but there are

624

contrasts. Layer 6 contains high expression of non-α7 nAChRs (Tribollet et al., 2004),

625

including α5 (Wada et al., 1990; Proulx et al., 2013) and α4 (Lein et al., 2007) subunits,

626

consistent with the high incidence of nAChR-mediated responses in layer 6 of PFC in

627

previous studies (Kassam et al., 2008; Poorthuis et al., 2012; Mailey et al., 2010). We

628

found that nAChR PSPs in layer 6 are common and of large amplitude in all three areas

629

studied, suggesting that the presence of α4/α5-mediated PSPs is a feature of layer 6

630

pyramidal neurons across cortical regions.

631

In layer 5, we found that nAChR PSPs are common in M1 and V1 and rare in PFC. In

632

M1, nAChR PSPs were mediated by non-α7 nAChRs. In contrast, Poorthuis et al., 2012

633

observed α7 nAChR-mediated responses in layer 5 pyramidal neurons in PFC. nAChR

634

PSPs that we observed originated from nAChRs in the proximal dendrites. Hence one

635

explanation for the lack of α7-mediated PSPs in our results might be that α7 nAChRs are

636

located primarily in the distal dendrites and that α7-mediated depolarization of the

637

distal dendrite failed to evoke depolarization at the soma in our experiments.

638

M1 is unusual in that the pyramidal neurons in layer 5B of M1 commonly display large-

639

amplitude nAChR PSPs, unlike layer 5 pyramidal neurons in PFC and V1. The α5 subunit

640

is not present in layer 5 (Wada et al., 1990). There is evidence for greater expression of

641

α4 subunits in deep than in superficial layers (Lein et al., 2007), but see also (Tribollet et

642

al., 2004), but this expression pattern is not unique to M1. Hence it is unclear which

643

nAChR subunit(s) underlie the unusually large nAChR PSPs in layer 5 of M1, but α5

644

subunits are unlikely to be involved.

645

In superficial layers, we found that pyramidal neurons in M1 and PFC rarely displayed

646

nAChR PSPs, consistent with results from PFC (Poorthuis et al., 2012), but that nAChR

647

PSPs are common in superficial pyramidal neurons in V1. Layer 2/3 contains dense non-

648

α7 labeling (Tribollet et al., 2004). Our results suggest that this dense labeling is from

649

interneurons and perhaps in the dendrites of deep-layer pyramidal neurons.

650

Our results extend our knowledge of pyramidal neuron nAChRs from PFC into M1 and

651

V1, revealing variation in laminar responsiveness to ACh between cortical regions;

652

presumably the responsiveness of pyramidal neurons is tuned to the unique demands of

653

each area. However, our results also reveal a general tendency for ACh to exert stronger

654

effects in deep than superficial layers, suggesting that preferential modulation of deep

655

layer pyramidal neurons via nAChRs is a general property of the actions of ACh in

656

neocortex.

657 658

nAChR-mediated modulation of neocortical circuits

659

How might the layer-selective effect of ACh influence the flow of excitation through

660

neocortex? The principal ascending excitatory drive to cortex is thalamocortical axons,

661

which contact pyramidal neurons primarily in layers 5B and 4 (5B and 3 in primary

662

motor cortex, which lacks layer 4 (Shepherd, 2009; Hooks et al., 2013). From layer 4,

663

information is passed by excitatory connections through layer 2/3 to layer 5 and to the

664

sub-cortical projection targets of neocortex. Hence there are two primary excitatory

665

pathways through neocortex: a short loop which connects thalamus with target

666

structures through layer 5 and a longer loop which includes layers 4 and 2/3

667

(Armstrong-James et al., 1992; de Kock et al., 2007; Petreanu et al., 2009;

668

Constantinople & Bruno, 2013).

669

ACh enhances activation of neocortical pyramidal neurons by ascending thalamic

670

drive. This enhancement arises from mAChR-mediated depolarization of pyramidal

671

neurons and enhanced glutamate release from thalamocortical terminals in layer 4

672

(Metherate, 2004; Disney et al., 2007; Gil et al., 1997). In addition, ACh activates non-

673

fast spiking, non-parvalbumin-expressing interneurons in layers 1 and 2/3 via non-α7

674

nAChRs (Porter et al., 1999; Gulledge et al., 2007; Letzkus et al., 2011; Arroyo et al.,

675

2012; Brombas et al., 2014). These interneurons inhibit parvalbumin- and somatostatin-

676

expressing interneurons that target the somata and dendrites of pyramidal neurons.

677

Hence nAChR-mediated inhibition of superficial interneurons reduces inhibition of

678

superficial pyramidal neurons (Letzkus et al., 2011; Brombas et al., 2014). Our results

679

indicate that ACh, again acting at nAChRs, directly promotes the spiking of deep-layer

680

pyramidal neurons. Hence ACh modulates cortical output by at least three different

681

nAChR-dependent mechanisms which enhance the responsiveness of neocortex to

682

incoming sensory drive: by increasing the release of glutamate in layer 4, by indirectly

683

enhancing the excitability of superficial pyramidal neurons and by directly enhancing the

684

excitability of deep-layer pyramidal neurons.

685

Why employ several mechanisms to modulate the excitability of pyramidal neurons

686

in cortex? Multiple mechanisms may offer circuit specificity and semi-independent

687

control. Multiple mechanisms of network modulation might allow ACh to independently

688

modulate the excitability of deep and superficial layer pyramidal neurons, thereby gating

689

the balance of sensory information processing in different cortical layers and controlling

690

the balance of information flowing through the long and short loops from thalamus to

691

the output structures of neocortex.

692

Acknowledgements:

693

We thank Becky Imhoff and Lauren Sybert for technical assistance. TH and JW

694

designed and performed the experiments, analyzed the results and wrote the manuscript.

695 696

Grants:

697

This work was supported by the National Institute of Mental Health (5T32MH067564

698

and 5R21MH085117), the National Institute of Neurological Disorders and Stroke

699

(1R01NS078067) and the Brain Research Foundation (BRF SG 2010-13).

700 701

Disclosures:

702

The authors declare no competing financial interests.

703

References

704

Allen TGJ, Abogadie FC, Brown DA. Simultaneous release of glutamate and

705

acetylcholine from single magnocellular “cholinergic” basal forebrain neurons. J

706

Neurosci 26: 1588–1595, 2006.

707 708 709 710 711

Armstrong-James M, Fox K, Das-Gupta A. Flow of excitation within rat barrel cortex on striking a single vibrissa. J Neurophysiol 68: 1345-1358, 1992. Arroyo S, Bennett C, Hestrin S. Nicotinic modulation of cortical circuits. Front Neural Circuits 8: 1-6, 2014. Arroyo S, Bennett C, Aziz D, Brown SP, Hestrin S. Prolonged disynaptic inhibition in the

712

cortex mediated by slow, non-α7 nicotinic excitation of a specific subset of cortical

713

interneurons. J Neurosci 32(11): 3859 –3864, 2012.

714 715 716

Avesar D, Gulledge AT. Selective serotonergic excitation of callosal projection neurons. Front in Neural Circuits 6: 1-11, 2012. Bailey CDC, De Biasi M, Fletcher PJ, Lambe EK. The nicotinic acetylcholine receptor α5

717

subunit plays a key role in attention circuitry and accuracy. J Neurosci 30: 9241–9252,

718

2010.

719

Beique J-C, Imad M, Mladenovic L, Gingrich JA, Andrade R. Mechanism of the 5-

720

hydroxytryptamine 2A receptor-mediated facilitation of synaptic activity in prefrontal

721

cortex. PNAS 104: 9870-9875, 2007.

722

Bell KS, Shim H, Chen C-K, McQuiston AR. Nicotinic excitatory postsynaptic potentials

723

in hippocampal CA1 interneurons are predominantly mediated by nicotinic receptors

724

that contain α4 and β2 subunits. Neuropharm 61: 1379-1388, 2011.

725

Bennett C, Arroyo S, Berns D, Hestrin S. Mechanisms generating dual-component

726

nicotinic EPSCs in cortical interneurons. J Neurosci 32(48): 17287-17296, 2012.

727 728 729

Berg RW, Friedman B, Schroeder LF, Kleinfeld D. Activation of nucleus basalis facilitates cortical control of a brain stem motor program. J Neurophysiol 94: 699–711, 2005. Brombas A, Fletcher LN, Williams SR. Activity-dependent modulation of layer 1

730

inhibitory neocortical circuits by acetylcholine. J Neurosci 34(5): 1932–1941, 2014.

731

Carr DB, Surmeier DJ. M1 muscarinic receptor modulation of Kir2 channels enhances

732

temporal summation of excitatory synaptic potentials in prefrontal cortex pyramidal

733

neurons. J Neurophysiol 97: 3432–3438, 2007.

734 735

Chu ZG, Zhou FM, Hablitz JJ. Nicotinic acetylcholine receptor-mediated synaptic potentials in rat neocortex. Brain Res 887: 399–405, 2000.

736

Conner JM, Culberson A, Packowski C, Chiba AA, Tuszynski MH. Lesions of the basal

737

forebrain cholinergic system impair task acquisition and abolish cortical plasticity

738

associated with motor skill learning. Neuron 38: 819–829, 2003.

739 740 741

Constantinople CM, Bruno RM. Deep cortical layers are activated directly by thalamus. Science 340(6140): 1591-1594, 2013. de Kock CP, Bruno RM, Spors H, Sakmann B. Layer- and cell-type-specific

742

suprathreshold stimulus representation in rat primary somatosensory cortex. J Physiol

743

581: 139-154, 2007.

744 745 746 747

Delmas P, Brown DA. Pathways modulating neural KCNQ/M(Kv7) potassium channels. Nature Rev Neurosci 6: 850-862, 2005. Dembrow NC, Chitwood RA, Johnston D. Projection-specific neuromodulation of medial prefrontal cortex neurons. J Neurosci 30: 16922–16937, 2010.

748

Dilsaver S. Cholinergic Mechanisms in Depression. Brain Res 396: 285-316, 1986.

749

Disney AA, Aoki C, Hawken MJ. Gain modulation by nicotine in macaque V1. Neuron 56:

750

701-713, 2007.

751

Duffy AM, Zhou P, Milner TA, Pickel VM. Spatial and intracellular relationships between

752

the α7 nicotinic acetylcholine receptor and the vesicular acetylcholine transporter in

753

the prefrontal cortex of rat and mouse. Neuroscience 161: 1091–1103, 2009.

754 755 756

Egorov AV, Hamam BN, Fransen E, Hasselmo ME, Alonso AA. Graded persistent activity in entorhinal cortex neurons. Nature 420: 173-178, 2002. Fedorov N, Benson L, Graef JD, Hyman J, Sollenberger J, Perrefsson F, Lippiello PM,

757

Bencherif M. A method for bidirectional solution exchange—“liquid bullet”

758

applications of acetylcholine to alpha7 nicotinic receptors. J Neurosci Methods 206(1):

759

23-33, 2012.

760

Frazier CJ, Buhler AV, Weiner JL, Dunwiddie TV. Synaptic potentials mediated via

761

alpha-bungarotoxin-sensitive nicotinic acetylcholine receptors in rat hippocampal

762

interneurons. J Neurosci 18(20): 8228-8235, 1998.

763

Gaspar P, Bloch B, Le Moine C. D1 and D2 receptor gene expression in the rat fronal

764

cortex: cellular localization in different classes of efferent neurons. Eur J Neurosci 7:

765

1050-1063, 1995.

766

Gee S, Ellwood I, Patel T, Luongo F, Deisseroth K, Sohal VS. Synaptic activity unmasks

767

dopamine D2 receptor modulation of a specific class of layer V pyramidal neurons in

768

prefrontal cortex. J Neurosci 32: 4959-4971, 2012.

769

Gharbawie OA, Whishaw IQ. Cholinergic and serotonergic neocortical projection lesions

770

given singly or in combination cause only mild impairments on tests of skilled

771

movement in rats: evaluation of a model of dementia. Brain Res 970: 97-109, 2003.

772

Giessel & Sabatini. M1 muscarinic receptors boost synaptic potentials and calcium influx

773

in dendritic spines by inhibiting postsynaptic SK channels. Neuron 68(5): 936-947,

774

2010.

775 776 777

Gil Z, Connors BW, Amitai Y. Differential regulation of neocortical synapses by neuromodulators and activity. Neuron 19: 679–686, 1997. Gritti I, Henny P, Galloni F, Mainville L, Mariotti M, Jones BE. Stereological estimates of

778

the basal forebrain cell population in the rat, including neurons containing choline

779

acetyltransferase, glutamic acid decarboxylase or phosphate-activated glutaminase and

780

colocalizing vesicular glutamate transporters. Neuroscience 143: 1051–1064, 2006.

781

Guillem K, Bloem B, Poorthuis RB, Loos M, Smit AB, Maskos U, Spijker S, Mansvelder

782

HD. Nicotinic Acetylcholine Receptor β2 Subunits in the Medial Prefrontal Cortex

783

Control Attention. Science 333 (6044): 888-891, 2011.

784 785 786 787 788 789 790

Gulledge AT, Park SB, Kawaguchi Y, Stuart GJ. Heterogeneity of phasic cholinergic signaling in neocortical neurons. J Neurophysiol 97: 2215–2229, 2007. Gulledge AT, Stuart GJ. Cholinergic inhibition of neocortical pyramidal neurons. J Neurosci, 25: 10308-10320, 2005. Haj-Dahmane S, Andrade R. Muscarinic Activation of a Voltage-Dependent Cation Nonselective Current in Rat Association Cortex. J Neurosci 16(12): 3848-3861, 1996. Henny P, Jones BE. Projections from basal forebrain to prefrontal cortex comprise

791

cholinergic, GABAergic and glutamatergic inputs to pyramidal cells or interneurons.

792

Eur J Neurosci 27: 654-670, 2008.

793

Herrero JL, Roberts MJ, Delicato LS, Gieselmann MA, Dayan P, Thiele A. Acetylcholine

794

contributes through muscarinic receptors to attentional modulation in V1. Nature 454:

795

1110-1114, 2008.

796

Hooks BM, Mao T, Gutnisky DA, Yamawaki N, Svoboda K, Shepherd GM. Organization

797

of cortical and thalamic input to pyramidal neurons in mouse motor cortex. J Neurosci

798

33: 748-760, 2013.

799 800 801 802 803

Jones BE. Modulation of cortical activation and behavioral arousal by cholinergic and orexinergic systems. Ann N Y Acad Sci 1129: 26-34, 2008. Kalmbach A, Hedrick T, Waters J. Selective optogenetic stimulation of cholinergic axons in neocortex. J Neurophysiol 107: 2008-2019, 2012. Kassam SM, Herman PM, Goodfellow NM, Alves NC, Lambe EK. Developmental

804

excitation of corticothalamic neurons by nicotinic acetylcholine receptors. J Neurosci

805

28: 8756-8764, 2008.

806 807

Kilgard M. Cholinergic modulation of skill learning and plasticity. Neuron 38(5): 678680, 2003.

808

Koliatsos VE, Martin LJ, Price DL. Efferent organization of the mammalian basal

809

forebrain. In: Brain Cholinergic Systems. Steriade M, Biesold D, Eds. Oxford

810

University Press. ISBN 0-19-854266-6, 1990.

811 812

Krnjevic K. The mechanism of excitation by acetylcholine in the cerebral cortex. J Physiol 215(1): 247-268, 1971.

813

Larkum ME, Zhu JJ, Sakmann B. Dendritic mechanisms underlying the coupling of the

814

dendritic with the axonal action potential initiation zone of adult rat layer 5 pyramidal

815

neurons. J Physiol 533.2: 447–466, 2001.

816 817

Lee MG, Hassani OK, Alonso A, Jones BE. Cholinergic basal forebrain neurons burst with theta during waking and paradoxical sleep. J Neurosci 25(17):4365–4369, 2005.

818

Lein ES, Hawrylycz JM, Ao N, Ayres M, Bensinger A, Bernard A, Boe AF, Boguski MS,

819

Brockway KS, Byrnes EJ, Chen L, Chen L, Chen TM, Chin MC, Chong J, Crook BE,

820

Czaplinska A, Dang CN, Datta S, Dee NR, Desaki AL, Desta T, Diep E, Dolbeare TA,

821

Donelan MJ, Dong HW, Dougherty JG, Duncan BJ, Ebbert AJ, Eichele G, Estin LK,

822

Faber C, Facer BA, Fields R, Fischer SR, Fliss TP, Frensley C, Gates SN, Glattfelder KJ,

823

Halverson KR, Hart MR, Hohmann JG, Howell MP, Jeung DP, Johnson RA, Karr PT,

824

Kawal R, Kidney JM, Knapik RH, Kuan CL, Lake JH, Laramee AR, Larsen KD, Lau C,

825

Lemon TA, Liang AJ, Liu Y, Luong LT, Michaels J, Morgan JJ, Morgan RJ, Mortrud

826

MT, Mosqueda NF, Ng LL, Ng R, Orta GJ, Overly CC, Pak TH, Parry SE, Pathak SD,

827

Pearson OC, Puchalski RB, Riley ZL, Rockett HR, Rowland SA, Royall JJ, Ruiz MJ,

828

Sarno NR, Schaffnit K, Shapovalova NV, Sivisay T, Slaughterbeck CR, Smith SC, Smith

829

KA, Smith BI, Sodt AJ, Stewart NN, Stumpf KR, Sunkin SM, Sutram M, Tam A,

830

Teemer CD, Thaller C, Thompson CL, Varnam LR, Visel A, Whitlock RM, Wohnoutka

831

PE, Wolkey CK, Wong VY, Wood M, Yaylaoglu MB, Young RC, Youngstrom BL, Yuan

832

XF, Zhang B, Zwingman TA, Jones AR. Genome-wide atlas of gene expression in the

833

adult mouse brain. Nature 445: 168-176, 2007.

834

Letzkus JL, Wolff SBE, Meyer EMM, Tovote P, Courtin J, Herry C, Luthi A. A

835

disinhibitory microcircuit for associative fear learning in the auditory cortex. Nature

836

480: 331-335, 2011.

837

Levy RB, Aoki C. α7 nicotinic acetylcholine receptors occur at postsynaptic densities of

838

AMPA receptor-positive and -negative excitatory synapses in rat sensory cortex. J

839

Neurosci 22: 5001–5015, 2002.

840

Lin JY, Lin MZ, Steinbach P, Tsien RY. Characterization of engineered channelrhodopsin

841

variants with improved properties and kinetics. Biophys J 96: 1803–1814, 2009.

842

Lubin M, Erisir A, Aoki C. Ultrastructural immunolocalization of the alpha 7 nAChR

843

subunit in guinea pig medial prefrontal cortex. ANYAS 868: 628-632, 1999.

844

Lysakowski A, Wainer BH, Bruce G, Hersh LB. An atlas of the regional and laminar

845

distribution of choline acetyltransferase immunoreactivity in rat cerebral cortex.

846

Neuroscience 28(2): 291-336, 1989.

847

Madisen L, Mao T, Koch H, Zhuo JM, Berenyi A, Fujisawa S, Hsu YW, Garcia AL 3rd, Gu

848

X, Zanella S, Kidney J, Gu H, Mao Y, Hooks BM, Boyden ES, Buzsaki G, Ramirez JM,

849

Jones AR, Svoboda K, Han X, Turner EE, Zeng H. A toolbox of Cre-dependent

850

optogenetic transgenic mice for light-induced activation and silencing. Nat Neurosci

851

15: 793-802, 2012.

852

Manns ID, Alonso A, Jones BE. Discharge properties of juxtacellularly labeled and

853

immunohistochemically identified cholinergic basal forebrain neurons recorded in

854

association with the electroencephalogram in anesthetized rats. J Neurosci 20(4):

855

1505–1518, 2000.

856

Manns ID, Mainville L, Jones BE. Evidence for glutamate, in addition to acetylcholine

857

and GABA, neurotransmitter synthesis in basal forebrain neurons projecting to the

858

entorhinal cortex. Neuroscience 107: 249-263, 2001.

859 860 861 862 863 864 865 866 867 868

McCormick DA. Neurotransmitter actions in the thalamus and cerebral cortex. J Clin Neurophysiol 9(2): 212-223, 1992. McCormick DA, Prince DA. Two types of muscarinic response to acetylcholine in mammalian cortical neurons. PNAS 82: 6344-6348, 1985. McCormick DA, Prince DA. Mechanisms of action of acetylcholine in the guinea-pig cerebral cortex in vitro. J Physiol 375: 169-194, 1986. McCormick DA, Williamson A. Convergence and divergence of neurotransmitter action in human cerebral cortex. PNAS USA 86: 8098–8102, 1989. Metherate R. Nicotinic acetylcholine receptors in sensory cortex. Learn Mem 11(1): 5059, 2004.

869

Milojkovic BA, Radojicic MS, Antic SD. A strict correlation between dendritic and

870

somatic plateau depolarizations in the rat prefrontal cortex pyramidal neurons. J

871

Neurosci 25(15):3940 –3951, 2005.

872 873 874

Milojkovic BA, Zhou W-L, Antic SD. Voltage and calcium transients in basal dendrites of the rat prefrontal cortex. J Physiol 585.2: 447–468, 2007. Mrzljak L, Levey AI, Belcher S, Goldman-Rakic PS. Localization of the m2 muscarinic

875

acetylcholine receptor protein and mRNA in cortical neurons of the normal and

876

cholinergically deafferented rhesus monkey. J Comp Neurol 390: 112–132, 1998.

877

Mrzljak L, Levey AI, Goldman-Rakic PS. Association of m1 and m2 muscarinic receptor

878

proteins with asymmetric synapses in the primate cerebral cortex: morphological

879

evidence for cholinergic modulation of excitatory neurotransmission. PNAS 90: 5194-

880

5198, 1993.

881

Nagel G, Szellas T, Huhn W, Kateriya S, Adeishvili N, Berthold P, Ollig D, Hegemann P,

882

Bamberg E. Channelrhodopsin-2, a directly light-gated cation-selective membrane

883

channel. Proc Natl Acad Sci USA 100: 13940–13945, 2003.

884

Nakayama H, Shioda S, Okuda H, Nakashima T, Nakai Y. Immunocytochemical

885

localization of nicotinic acetylcholine receptor in rat cerebral cortex. Mol Brain Res 32

886

(1995): 321-328, 1995.

887

Navaroli VL, Zhao Y, Boguszewski P, Brown TH. Muscarinic receptor activation enables

888

persistent firing in pyramidal neurons from superficial layers of dorsal perirhinal

889

cortex. Hippocampus 22: 1392-1404, 2011.

890 891

Olivas ND, Quintanar-Zilinskas V, Nenadic Z, Xu X. Laminar circuit organization and response modulation in mouse visual cortex. Front Neural Circuits 6: 70, 2012.

892 893 894 895 896

Parikh V, Sarter M. Cholinergic mediation of attention: contributions of phasic and tonic increases in prefrontal cholinergic activity. ANYAS 1129: 225–235, 2008. Petreanu L, Mao T, Sternson SM, Svoboda K. The subcellular organization of neocortical excitatory connections. Nature 457: 1142-1146, 2009. Petrof I, Viaene AN, Sherman SM. Two populations of corticothalamic and interareal

897

corticocortical cells in the subgranular layers of the mouse primary sensory cortices. J

898

Comp Neurol 520(8): 1678-1686, 2012.

899

Poorthuis RB, Bloem B, Schak B, Wester J, de Kock CPJ, Mansvelder HD. Layer-specific

900

modulation of the prefrontal cortex by nicotinic acetylcholine receptors. Cereb Cortex

901

23(1): 148-161, 2013.

902

Poorthuis RB, Bloem B, Verhoog MB, Mansvelder HD. Layer-specific interference with

903

cholinergic signaling in the prefrontal cortex by smoking concentrations of nicotine. J

904

Neurosci 33: 4843– 4853, 2013.

905

Porter JT, Cauli B, Tsuzuki K, Lambolez B, Rossier J, Audinat E. Selective excitation of

906

subtypes of neocortical interneurons by nicotinic receptors. J Neurosci 19: 5228–5235,

907

1999.

908

Proulx E, Piva M, Tian MK, Bailey CD, Lambe EK. Nicotinic acetylcholine receptors in

909

attention circuitry: the role of layer VI neurons of prefrontal cortex. Cell Mol Life Sci

910

71(7): 1225-1244, 2013.

911 912 913 914

Raedler TJ, Tandon R. Cholinergic mechanisms in schizophrenia: current concepts. Current Psychosis and Therapeutics Reports 1: 20-26, 2006. Rahman J, Berger T. Persistent activity in layer 5 pyramidal neurons following cholinergic activation of mouse primary cortices. Eur J Neurosci 34: 22–30, 2011.

915

Roerig B, Nelson DA, Katz LC. Fast synaptic signaling by nicotinic acetylcholine and

916

serotonin 5-HT3 receptors in developing visual cortex. J Neurosci 17: 8353–8362,

917

1997.

918

Ramanathan D, Tuszynski MH, Conner JM. The basal forebrain cholinergic system is

919

required specifically for behaviorally mediated cortical map plasticity. J Neurosci 29:

920

5992–6000, 2009.

921 922

Sarter M, Parikh V, Howe WM. Phasic acetylcholine release and the volume transmission hypothesis: time to move on. Nature Rev Neurosci 10: 383-390, 2009.

923

Schwindt PC, Spain WJ, Foehring RC, Chubb MC, Crill WE. Slow conductances in

924

neurons from cat sensorimotor cortex in vitro and their role in slow excitability

925

changes. J Neurophysiol 59(2): 450-467, 1988.

926

Seong HJ, Carter AG . D1 receptor modulation of action potential firing in a

927

subpopulation of layer 5 pyramidal neurons in the prefrontal cortex. J Neurosci 32:

928

10516-10521, 2012.

929

Sheets PL, Suter BA, Kiritani T, Chan CS, Surmeier DJ, Shepherd GMG. Corticospinal-

930

specific HCN expression in mouse motor cortex: Ih-dependent synaptic integration as

931

a candidate microcircuit mechanism involved in motor control. J Neurophysiol 106:

932

2216-2231, 2011.

933 934 935 936 937 938

Shepherd GM. Intracortical cartography in an agranular area. Front Neurosci 3(3): 337343, 2009. Shepherd GM. Corticostriatal connectivity and its role in disease. Nat Rev Neurosci 14(4): 278-291, 2013. Steriade M. Aceylcholine systems and rhythmic activities during the waking-sleep cycle. Prog Brain Res 145: 179-196, 2004.

939

Tribollet E, Bertrand D, Marguerat A, Raggenbass M. Comparative distribution of

940

nicotinic receptor subtypes during development, adulthood and aging: an

941

autoradiographic study in the rat brain. Neuroscience 12: 405–420, 2004.

942

van der Zee EA, Streefland C, Strosberg AD, Schröder H, Luiten PG. Visualization of

943

cholinoceptive neurons in the rat neocortex: colocalization of muscarinic and nicotinic

944

acetylcholine receptors. Mol Brain Res 14(4): 326-36, 1992.

945

Vidal C, Changeux JP. Nicotinic and muscarinic modulations of excitatory synaptic

946

transmission in the rat prefrontal cortex in vitro. Neuroscience 56: 23–32, 1993.

947

Wada E, McKinnon D, Heinemann S, Patrick J, Swanson LW. The distribution of mRNA

948

encoded by a new member of the neuronal nicotinic acetylcholine receptor gene family

949

(α5) in the rat central nervous system. Brain Res 526: 45-53, 1990.

950

Wainer BH, Mesulam MM. Ascending cholinergic pathways in the rat brain. In: Brain

951

Cholinergic Systems. Steriade M, Biesold D, Eds. Oxford University Press. ISBN 0-19-

952

854266-6, 1990.

953

Wevers A, Jeske A, Lobron C, Birtsch C, Heinemann S, Maelicke A, Schroder R, Schroder

954

H. Cellular distribution of nicotinic acetylcholine receptor subunit mRNAs in the

955

human cerebral cortex as revealed by non-isotopic in situ hybridization. Mol Brain Res

956

25: 122-128, 1994.

957 958 959

Whitehouse PJ, Hedreen JC, White CL, Price DL. Basal forebrain neurons in the dementia of Parkinson disease. Ann Neurol 13: 243-248, 1983. Whitehouse PJ, Price D, Struble RG, Clarke AW, Coyle JT, Delong MR. Alzheimer’s

960

disease and senile dementia: loss of neurons in the basal forebrain. Science 215: 1237-

961

1239, 1982.

962 963

Winkler J, Suhr ST, Gage FH, Thal LJ, Fisher LJ. Essential role of neocortical acetylcholine in spatial memory. Nature 375: 484-487, 1995.

964

Yamasaki M, Matsui M, Watanabe M. Preferential localization of muscarinic M1 receptor

965

on dendritic shaft and spine of cortical pyramidal cells and its anatomical evidence for

966

volume transmission. J Neurosci 30(12): 4408–4418, 2010.

967

Yamashita T, Isa T. Flufenamic acid sensitive, Ca(2+)-dependent inward current induced

968

by nicotinic acetylcholine receptors in dopamine neurons. Neurosci Res 46: 463-473,

969

2003a.

970

Yamashita T, Isa T . Ca2+-dependent inward current induced by nicotinic receptor

971

activation depends on Ca2+/calmodulin-CaMKII pathway in dopamine neurons.

972

Neurosci Res 47: 225-232, 2003b.

973 974 975 976 977 978

Yizhar O, Fenno LE, Davidson TJ, Mogri M, Deisseroth K. Optogenetics in Neural Systems. Neuron 71: 9-34, 2011. Zhang YP, Oertner TG. Optical induction of synaptic plasticity using a light-sensitive channel. Nat Meth 4: 139-141, 2007. Zhang Z, Séguéla P. Metabotropic induction of persistent activity in layers II/III of anterior cingulate cortex. Cereb Cortex 20: 2948-2957, 2010.

979

Zolles G, Wagner E, Lampert A, Sutor B. Functional expression of nicotinic acetylcholine

980

receptors in rat neocortical layer 5 pyramidal cells. Cereb Cortex 19: 1079-1091, 2009.

981

Figure legends

982

Figure 1. Stimulation of cholinergic axons evokes four responses in

983

pyramidal neurons

984

(A) Schematic of a coronal slice illustrating the expression of ChR2-YFP in cholinergic

985

neurons in the basal forebrainand their axonal projections to neocortex. CP: caudate

986

putamen, LV: lateral ventricle. (B) Image of ChR2-YFP labeled neurons in nucleus

987

basalis in a ChAT-ChR2(Ai32) mouse. Maximum intensity projection from a 2-photon z-

988

stack through a fixed section. (C) Image of ChR2-YFP labeled axons in primary motor

989

cortex in a ChAT-ChR2(Ai32) mouse. Maximum intensity projection from a 2-photon z-

990

stack through a fixed section. (D,E) Pyramidal neuron in layer 5B of primary motor

991

cortex, 3 weeks after virus injection. Neuron was filled with Alexa 594 during whole-cell

992

recording (D) and surrounding cholinergic axons expressing ChR2-YFP (E). Both images

993

are maximum intensity projections from 2-photon z-stacks. (F) Examples of voltage

994

recordings from four layer 5B pyramidal neurons at rest. Each voltage response evoked

995

by stimulation of cholinergic axons by widefield illumination of the slice with a blue LED

996

(stimulus: 10 x 5 ms at 20 Hz; blue bars). Example of slow depolarization includes a

997

preceding hyperpolarization. Examples of slow depolarization and fast hyperpolarization

998

were obtained from virus-injected mice. Examples of medium depolarization and fast

999

depolarization were obtained from ChAT-ChR2(Ai32) mice.

1000 1001

Figure 2. Light-evoked responses are eliminated by TTX and by removal of

1002

extracellular calcium

1003

Examples of the slow depolarization and hyperpolarization (A, neuron depolarized by DC

1004

current injection), medium depolarization (B, neuron at rest), and fast depolarization (C,

1005

neuron at rest) from three neurons. All responses were reversibly eliminated by removal

1006

of external calcium, and by addition of 500 nM TTX. Stimulus: 10 x 5 ms illumination at

1007

20 Hz. All examples were obtained from ChAT-ChR2(Ai32) mice.

1008 1009

Figure 3. Pharmacology of the hyperpolarization and slow depolarization

1010

(A) Example of sequential addition of 100 µM mecamylamine and 1 µM atropine to a

1011

neuron displaying a light-evoked hyperpolarization. Stimulus: 10 x 5 ms illumination at

1012

20 Hz. Neuron held close to threshold by DC current injection. Recording from a virus-

1013

injected mouse. (B) Example of sequential addition of 100 µM mecamylamine and 1 µM

1014

atropine to a neuron displaying a light-evoked slow depolarization. Stimulus: 10 x 5 ms

1015

illumination at 20 Hz. Neuron held close to threshold by DC current injection. Recording

1016

from a ChAT-ChR2(Ai32) mouse. (C) Probability of evoking a mAChR-mediated

1017

potential during constant depolarization by DC current injection. Control, 38 neurons;

1018

mec, 100 µM mecamylamine, 13 neurons; atr, 1 µM atropine, 141 neurons.

1019 1020

Figure 4. Pharmacology of the medium depolarization

1021

Example recordings from six neurons. Stimulus: 2-3 x 5 ms illumination at 20 Hz.

1022

Physo, 0.5 µM physostigmine; atr, 1 µM atropine; mec, 100 µM mecamylamine; gal, 1 µM

1023

galanthamine; NBQX/CPP, 10 µM NBQX and 10 µM CPP; GBZ / CGP, 1 µM gabazine

1024

and 3 µM CGP 52432; MLA, 10 nM methylylcaconitine; DHβE, 10 µM dihydro-β-

1025

erythroidine hydrobromide. All examples at resting membrane potential and from virus-

1026

injected mice. Summary: mean ± SEM, normalized to pre-drug amplitude. TTX, 7

1027

neurons; 0 Ca, 7 neurons; NBQX/CPP, 7 neurons; GBZ / CGP, 8 neurons;

1028

physostigmine, 4 neurons; atropine, 5 neurons; mecamylamine, 12 neurons;

1029

galanthamine, 23 neurons; MLA, 12 neurons; DHβE, 16 neurons. Asterisks denote

1030

significant effects (P

Acetylcholine excites neocortical pyramidal neurons via nicotinic receptors.

The neuromodulator acetylcholine (ACh) shapes neocortical function during sensory perception, motor control, arousal, attention, learning, and memory...
4MB Sizes 0 Downloads 17 Views