14
Neuron Article Membrane Potential Dynamics of Neocortical Projection Neurons Driving Target-Specific Signals Takayuki Yamashita, 1, * Aure ´ lie Pala, 1 Leticia Pedrido, 2 Yves Kremer, 1 Egbert Welker, 2 and Carl C.H. Petersen 1, * 1 Laboratory of Sensory Processing, Brain Mind Institute, Faculty of Life Sciences, E ´ cole Polytechnique Fe ´ de ´ rale de Lausanne (EPFL), 1015 Lausanne, Switzerland 2 Department of Fundamental Neurosciences, University of Lausanne, 1005 Lausanne, Switzerland *Correspondence: takayuki.yamashita@epfl.ch (T.Y.), carl.petersen@epfl.ch (C.C.H.P.) http://dx.doi.org/10.1016/j.neuron.2013.10.059 SUMMARY Primary sensory cortex discriminates incoming sen- sory information and generates multiple processing streams toward other cortical areas. However, the underlying cellular mechanisms remain unknown. Here, by making whole-cell recordings in primary so- matosensory barrel cortex (S1) of behaving mice, we show that S1 neurons projecting to primary motor cor- tex (M1) and those projecting to secondary somato- sensory cortex (S2) have distinct intrinsic membrane properties and exhibit markedly different membrane potential dynamics during behavior. Passive tactile stimulation evoked faster and larger postsynaptic potentials (PSPs) in M1-projecting neurons, rapidly driving phasic action potential firing, well-suited for stimulus detection. Repetitive active touch evoked strongly depressing PSPs and only transient firing in M1-projecting neurons. In contrast, PSP summation allowed S2-projecting neurons to robustly signal sen- sory information accumulated during repetitive touch, useful for encoding object features. Thus, target- specific transformation of sensory-evoked synaptic potentials by S1 projection neurons generates func- tionally distinct output signals for sensorimotor coor- dination and sensory perception. INTRODUCTION Neuronal microcircuits in primary sensory cortex process sen- sory input and subsequently transmit output signals to other distant cortical regions for further processing (Felleman and Van Essen, 1991; Petersen, 2007; Nassi and Callaway, 2009). Distinct output pathways from primary sensory cortex are thought to be critically important for differential processing of sensory information, with different cortical signaling streams facilitating different aspects of sensory perception and sensori- motor coordination (Goodale and Milner, 1992; Mesulam, 1998; Nassi and Callaway, 2009). However, the underlying cellular and network mechanisms are poorly understood. Classical extracellular recordings from putative long-range projection neurons identified by antidromic stimulation in behaving monkeys show that cortical projection neurons with distinct cortical (Movshon and Newsome, 1996; Ferraina et al., 2002) or subcortical (Turner and DeLong, 2000; Ferraina et al., 2002; Hoffmann et al., 2002) axonal targets have differential ac- tivity patterns. Recent in vivo studies using somatic or axonal Ca 2+ imaging in ferrets and mice have confirmed that cortical projection neurons signal different information toward different cortical targets (Sato and Svoboda, 2010; Jarosiewicz et al., 2012; Petreanu et al., 2012; Glickfeld et al., 2013; Chen et al., 2013). Cortical projection neurons might therefore play important roles in generating distinct information streams in the large-scale neocortical neuronal network through target-specific action potential (AP) firing. APs are driven by membrane potential (V m ) depolarization beyond a relatively well-defined threshold as measured at the soma (Azouz and Gray, 2000; Poulet and Pe- tersen, 2008). In order to understand the cellular mechanisms driving AP firing, it is therefore essential to record V m changes, which are largely driven by synaptic inputs. Interestingly, in vitro brain slice studies have revealed differences in the synap- tic connectivity of cortical neurons that have axonal projections to other cortical or subcortical regions (Morishima and Kawagu- chi, 2006; Le Be ´ et al., 2007; Brown and Hestrin, 2009; Mao et al., 2011; Morishima et al., 2011; Otsuka and Kawaguchi, 2011; Kir- itani et al., 2012). However, nothing is currently known about how projection neurons integrate incoming synaptic input to compute functionally tuned output signals during animal behavior. In this study, we combined retrograde fluorescent labeling with in vivo targeted whole-cell recordings in behaving mice to measure V m dynamics of cortical neurons projecting to other cortical areas. We focused on the primary somatosensory barrel cortex (S1) of mice, which contains a remarkable map of the mystacial vibrissae such that each whisker is individually repre- sented in a well-defined cortical barrel column (Petersen, 2007; Diamond et al., 2008; Feldmeyer et al., 2013). For example, the C2 barrel column of S1 is known to process tactile information relating to the contralateral C2 whisker. S1 in rodents has major anatomical (Carvell and Simons, 1987; Welker et al., 1988; Aron- off et al., 2010) and functional (Ferezou et al., 2007; Matyas et al., 2010; Mao et al., 2011; Petreanu et al., 2012; Chen et al., 2013) connections to both primary motor cortex (M1) and secondary somatosensory cortex (S2) of the same hemisphere. Therefore, Neuron 80, 1477–1490, December 18, 2013 ª2013 Elsevier Inc. 1477

Membrane Potential Dynamics of Neocortical Projection ... · Membrane Potential Dynamics of Neocortical Projection Neurons Driving Target-Specific Signals Takayuki Yamashita,1,*

  • Upload
    others

  • View
    11

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Membrane Potential Dynamics of Neocortical Projection ... · Membrane Potential Dynamics of Neocortical Projection Neurons Driving Target-Specific Signals Takayuki Yamashita,1,*

Neuron

Article

Membrane Potential Dynamicsof Neocortical Projection NeuronsDriving Target-Specific SignalsTakayuki Yamashita,1,* Aurelie Pala,1 Leticia Pedrido,2 Yves Kremer,1 Egbert Welker,2 and Carl C.H. Petersen1,*1Laboratory of Sensory Processing, Brain Mind Institute, Faculty of Life Sciences, Ecole Polytechnique Federale de Lausanne (EPFL),

1015 Lausanne, Switzerland2Department of Fundamental Neurosciences, University of Lausanne, 1005 Lausanne, Switzerland*Correspondence: [email protected] (T.Y.), [email protected] (C.C.H.P.)

http://dx.doi.org/10.1016/j.neuron.2013.10.059

SUMMARY

Primary sensory cortex discriminates incoming sen-sory information and generates multiple processingstreams toward other cortical areas. However, theunderlying cellular mechanisms remain unknown.Here, by making whole-cell recordings in primary so-matosensory barrel cortex (S1) of behaving mice, weshowthatS1neuronsprojecting toprimarymotorcor-tex (M1) and those projecting to secondary somato-sensory cortex (S2) have distinct intrinsic membraneproperties and exhibit markedly different membranepotential dynamics during behavior. Passive tactilestimulation evoked faster and larger postsynapticpotentials (PSPs) in M1-projecting neurons, rapidlydriving phasic action potential firing, well-suited forstimulus detection. Repetitive active touch evokedstrongly depressing PSPs and only transient firing inM1-projecting neurons. In contrast, PSP summationallowedS2-projecting neurons to robustly signal sen-sory informationaccumulatedduring repetitive touch,useful for encoding object features. Thus, target-specific transformation of sensory-evoked synapticpotentials by S1 projection neurons generates func-tionally distinct output signals for sensorimotor coor-dination and sensory perception.

INTRODUCTION

Neuronal microcircuits in primary sensory cortex process sen-

sory input and subsequently transmit output signals to other

distant cortical regions for further processing (Felleman and

Van Essen, 1991; Petersen, 2007; Nassi and Callaway, 2009).

Distinct output pathways from primary sensory cortex are

thought to be critically important for differential processing of

sensory information, with different cortical signaling streams

facilitating different aspects of sensory perception and sensori-

motor coordination (Goodale and Milner, 1992; Mesulam,

1998; Nassi and Callaway, 2009). However, the underlying

cellular and network mechanisms are poorly understood.

Ne

Classical extracellular recordings from putative long-range

projection neurons identified by antidromic stimulation in

behaving monkeys show that cortical projection neurons with

distinct cortical (Movshon and Newsome, 1996; Ferraina et al.,

2002) or subcortical (Turner and DeLong, 2000; Ferraina et al.,

2002; Hoffmann et al., 2002) axonal targets have differential ac-

tivity patterns. Recent in vivo studies using somatic or axonal

Ca2+ imaging in ferrets and mice have confirmed that cortical

projection neurons signal different information toward different

cortical targets (Sato and Svoboda, 2010; Jarosiewicz et al.,

2012; Petreanu et al., 2012; Glickfeld et al., 2013; Chen et al.,

2013). Cortical projection neuronsmight therefore play important

roles in generating distinct information streams in the large-scale

neocortical neuronal network through target-specific action

potential (AP) firing. APs are driven by membrane potential (Vm)

depolarization beyond a relatively well-defined threshold as

measured at the soma (Azouz and Gray, 2000; Poulet and Pe-

tersen, 2008). In order to understand the cellular mechanisms

driving AP firing, it is therefore essential to record Vm changes,

which are largely driven by synaptic inputs. Interestingly,

in vitro brain slice studies have revealed differences in the synap-

tic connectivity of cortical neurons that have axonal projections

to other cortical or subcortical regions (Morishima and Kawagu-

chi, 2006; Le Be et al., 2007; Brown andHestrin, 2009;Mao et al.,

2011; Morishima et al., 2011; Otsuka and Kawaguchi, 2011; Kir-

itani et al., 2012). However, nothing is currently known about how

projection neurons integrate incoming synaptic input to compute

functionally tuned output signals during animal behavior.

In this study, we combined retrograde fluorescent labeling

with in vivo targeted whole-cell recordings in behaving mice to

measure Vm dynamics of cortical neurons projecting to other

cortical areas. We focused on the primary somatosensory barrel

cortex (S1) of mice, which contains a remarkable map of the

mystacial vibrissae such that each whisker is individually repre-

sented in a well-defined cortical barrel column (Petersen, 2007;

Diamond et al., 2008; Feldmeyer et al., 2013). For example, the

C2 barrel column of S1 is known to process tactile information

relating to the contralateral C2 whisker. S1 in rodents has major

anatomical (Carvell and Simons, 1987; Welker et al., 1988; Aron-

off et al., 2010) and functional (Ferezou et al., 2007; Matyas et al.,

2010; Mao et al., 2011; Petreanu et al., 2012; Chen et al., 2013)

connections to both primary motor cortex (M1) and secondary

somatosensory cortex (S2) of the same hemisphere. Therefore,

uron 80, 1477–1490, December 18, 2013 ª2013 Elsevier Inc. 1477

Page 2: Membrane Potential Dynamics of Neocortical Projection ... · Membrane Potential Dynamics of Neocortical Projection Neurons Driving Target-Specific Signals Takayuki Yamashita,1,*

Neuron

Target-Specific Membrane Potential Dynamics

the whisker sensorimotor cortex offers unique opportunities for

linking the architecture of specific local microcircuits to large-

scale cortical processing of sensory information. Here, we spe-

cifically measured the Vm dynamics of layer 2/3 neurons in the

C2 barrel column projecting to M1, which we compare to neu-

rons residing in the same layer and barrel column, but projecting

to S2. Our results indicate that M1- and S2-projecting neurons

have distinct intrinsic membrane properties and that they differ-

ently process incoming synaptic input to compute functionally

relevant, target-specific output signals.

RESULTS

Retrograde Labeling and Visualization of S1 ProjectionNeuronsWe labeled S1 neurons that project toM1 andS2 of the ipsilateral

hemisphere using the retrograde tracer choleratoxin subunit

B (CTB) conjugated with fluorescent dyes (Figures 1A–1C)

(Conte et al., 2009). Injection of CTB conjugated with different

fluorescent dyes into M1 and S2 revealed that the spatial loca-

tion of retrogradely labeled neurons was intermingled in layer

2/3 of the C2 barrel column of S1 (Figure 1C), and only very

few cells were double-labeled with both tracers (0.8% ± 0.2%

of total CTB-labeled cells at the subpial depth of 0–300 mm,

n = 4 mice) (Sato and Svoboda, 2010; Chen et al., 2013). By

selectively expressing GFP in CTB-labeled cells using in vivo sin-

gle-cell electroporation (Kitamura et al., 2008), we confirmed that

CTB-labeled neurons in S1 had extensive axonal arborizations in

the CTB injection sites (n = 10 out of 10 cells; Figure 1D; Figures

S1A and S1B available online). Furthermore, using transgenic

mice to label inhibitory neurons, we confirmed that the vast ma-

jority of CTB-labeled projection neurons in S1 were not

GABAergic (Figures S1C and S1D) (Tamamaki and Tomioka,

2010).

Distinct Intrinsic Membrane Properties among S1Projection NeuronsTo measure Vm dynamics of S1 projection neurons, whole-cell

patch-clamp recordings from CTB-labeled neurons in S1 were

obtained under visual control using a two-photon microscope

(Margrie et al., 2003; Gentet et al., 2010, 2012; Mateo et al.,

2011). Recordings were targeted to layer 2/3 neurons in the C2

barrel column of awake head-restrained mice (Crochet and

Petersen, 2006; Poulet and Petersen, 2008; Gentet et al.,

2010). The subpial recording depth was 165 ± 4 mm (n = 71,

range: 110–245 mm) for M1-projecting neurons and 160 ± 4 mm

(n = 63, range: 110–245 mm) for S2-projecting neurons (not

different comparing M1- and S2-projecting neurons, p = 0.40)

(Figure 2A). We first examined basic membrane properties of

the projection neurons during quiet wakefulness, defined as

the periods when the whiskers were not moving. M1- and S2-

projecting neurons exhibited a similar mean Vm (M1-projecting

neurons, –66.5 ± 0.6 mV, n = 42; S2-projecting neurons,

–65.6 ± 0.7 mV, n = 42; p = 0.48) and AP threshold was also

similar (M1-projecting neurons, –42.4 ± 0.6 mV, n = 27; S2-pro-

jecting neurons, –41.8 ± 0.6 mV, n = 29; p = 0.59) (Figure S2).

Spontaneous AP rate was low for both these cell-types (M1-pro-

jecting neurons, 0.10 ± 0.02 Hz, n = 42; S2-projecting neurons,

1478 Neuron 80, 1477–1490, December 18, 2013 ª2013 Elsevier Inc

0.14 ± 0.03 Hz, n = 42; p = 0.72). Whereas these electrophysio-

logical properties ofM1- and S2-projecting neuronswere similar,

we also found striking differences. S2-projecting neurons had

significantly larger input resistance (Rin) (M1-projecting neurons,

35.1 ± 2.0 MU, n = 36; S2-projecting neurons, 54.7 ± 3.8 MU, n =

34; p < 0.0001) (Figure 2B). S2-projecting neurons also had

significantly longer membrane time constants (Tau) (M1-projec-

ting neurons, 6.5 ± 0.5 ms, n = 36; S2-projecting neurons, 9.9 ±

1.0 ms, n = 34; p = 0.0007) (Figure 2B). Presumably as a result of

the higher Rin, S2-projecting neurons were significantly more

excitable than M1-projecting neurons, with S2-projecting neu-

rons showing a left-shifted spike frequency-current relationship

(p < 0.0001, two-way ANOVA; Figures 2C and 2D). Furthermore,

the speed of AP repolarization was significantly slower in S2-pro-

jecting neurons compared to M1-projecting neurons (Figure S2).

Taken together, these results show that M1- and S2-projecting

neurons have distinct intrinsic membrane properties in vivo.

BehavioralModulation ofMembranePotential Dynamicsin S1 Projection NeuronsConsistent with previous Vm measurements from layer 2/3 pyra-

midal neurons of S1 barrel cortex (Petersen et al., 2003; Crochet

and Petersen, 2006; Poulet and Petersen, 2008; Gentet et al.,

2010, 2012; Okun et al., 2010; Crochet et al., 2011; Poulet

et al., 2012; Zagha et al., 2013), we found that both M1- and

S2-projecting neurons showed prominent slow-wave Vm fluctu-

ations during quiet wakefulness (Figure 3A). Interestingly, the

amplitude of such slow Vm dynamics was larger in M1-projecting

neurons compared to S2-projecting neurons. Integrating across

low-frequency components of the Vm fast Fourier transform

(FFT) spectrogram during quiet wakefulness, we found that

M1-projecting neurons had �50% larger amplitude 1–5 Hz Vm

fluctuations compared to S2-projecting neurons (M1-projecting

neurons, 6.1 ± 0.4 mV, n = 10; S2-projecting neurons, 4.2 ±

0.6 mV, n = 10; p = 0.021) (Figures 3A–3C). Furthermore, the

overall standard deviation (SD) of Vm fluctuations during quiet

wakefulness was also significantly larger in M1-projecting neu-

rons compared to S2-projecting neurons (M1-projecting neu-

rons, 4.5 ± 0.3 mV, n = 10; S2-projecting neurons, 3.0 ±

0.4 mV, n = 10; p = 0.010) (Figures 3A and 3C). Given that Rin

of M1-projecting neurons is lower than S2-projecting neurons,

this suggests that M1-projecting neurons receive substantially

stronger synaptic input during quiet wakefulness.

During exploratory rhythmic whisking there is a profound shift

in brain state such that excitatory layer 2/3 pyramidal neurons

reduce slow Vm fluctuations and depolarize (Crochet and Pe-

tersen, 2006; Poulet and Petersen, 2008; Gentet et al., 2010,

2012; Crochet et al., 2011; Poulet et al., 2012). Slow Vm fluctua-

tions were attenuated during whisking in both cell-types, quanti-

fied through the 1-5 Hz integral of the Vm FFT (M1-projecting

neurons, 2.6 ± 0.3mV, n = 10, p = 0.002 compared to quiet wake-

fulness; S2-projecting neurons, 2.3 ± 0.4 mV, n = 10; p = 0.002

compared to quiet wakefulness). Whisking also significantly

reduced Vm standard deviation in both types of projection neu-

rons (M1-projecting neurons, 2.4 ± 0.2 mV, n = 10, p = 0.002

compared to quiet wakefulness; S2-projecting neurons, 2.1 ±

0.3 mV, n = 10; p = 0.004 compared to quiet wakefulness) (Fig-

ures 3A–3C). Both cell-types also depolarized during whisking

.

Page 3: Membrane Potential Dynamics of Neocortical Projection ... · Membrane Potential Dynamics of Neocortical Projection Neurons Driving Target-Specific Signals Takayuki Yamashita,1,*

Figure 1. Retrograde Labeling and Visualization of S1 Neurons Projecting to M1 and S2

(A) CTB conjugated with Alexa-488 (green, 0.5%) and Alexa-594 (red, 0.5%) was injected into S2 (100 nl) and M1 (200 nl) of the left hemisphere.

(B) Epifluorescence images from coronal sections of the injection sites (blue, DAPI).

(C) CTB-labeled layer 2/3 neurons in S1 at the subpial depth of 190 mm imaged in vivo with a two-photon microscope (red, M1-projecting neurons; green,

S2-projecting neurons).

(D) Coronal (left) and horizontal (right) views of 3D-reconstructed projection neurons (M1-p, an M1-projecting neuron; S2-p, an S2-projecting neuron) with

dendrites (dark colors) and axons (light colors). Each cell had an extensive axonal arborization in the CTB-injection site together with local projections to supra-

and infragranular layers within S1 and also a callosal projection toward the other hemisphere, but we were unable to identify callosal targets. The brain outline,

shaded locations of M1 and S2, and the S1 barrel field are schematically indicated.

See also Figure S1.

Neuron

Target-Specific Membrane Potential Dynamics

(Vm during whisking: M1-projecting neurons, –63.7 ± 0.7 mV, n =

40, p < 0.0001 compared to quiet wakefulness; S2-projecting

neurons, –64.1 ± 0.8 mV, n = 41; p < 0.0001 compared to quiet

wakefulness) (Figure 3D). Overall mean AP rate for both cell-

types remained low during whisking (M1-projecting neurons,

0.21 ± 0.10 Hz, n = 40; S2-projecting neurons, 0.15 ± 0.08 Hz,

n = 41) (Figure 3E). However, a small fraction of both M1- and

S2-projecting neurons changed spike rate during whisking,

with a significant fraction of S2-projecting neurons decreasing

Ne

firing rate (p = 0.013, Wilcoxon signed rank test) (Figures 3E

and 3F).

A subset of layer 2/3 pyramidal neurons show rapid Vm dy-

namics phase-locked to the whisking cycle (Crochet and Pe-

tersen, 2006; Poulet and Petersen, 2008; Gentet et al., 2010;

Crochet et al., 2011). We therefore next examined the relation-

ship between whisker movement filmed with a high-speed

camera and rapid Vm dynamics in defined projection neurons

(Figure 4). M1-projecting neurons showed a fast Vm modulation

uron 80, 1477–1490, December 18, 2013 ª2013 Elsevier Inc. 1479

Page 4: Membrane Potential Dynamics of Neocortical Projection ... · Membrane Potential Dynamics of Neocortical Projection Neurons Driving Target-Specific Signals Takayuki Yamashita,1,*

Figure 2. Distinct Intrinsic Membrane Proper-

ties among S1 Projection Neurons

(A) Distribution of the subpial recording depth for

M1-projecting (M1-p, blue; n = 71) and S2-projecting

(S2-p, red; n = 63) neurons.

(B) Input resistance (Rin) and membrane time con-

stant (Tau) were significantly larger in S2-projecting

neurons than in M1-projecting neurons.

(C) Example responses to current injections during

whole-cell Vm recording from an M1-projecting and

an S2-projecting neuron.

(D) Thespike frequency-current relationship indicated

the higher excitability of S2-projecting neurons.

Each open circle in (B) corresponds to an individual

cell. Filled circles with error bars in (B) and (D) indi-

cate mean ± SEM. See also Figure S2.

Neuron

Target-Specific Membrane Potential Dynamics

phase-locked to the whisking cycle, which was not obvious in

S2-projecting neurons (Figure 4A). Averaging Vm segments

aligned at the peak of protraction (Figure 4B) revealed that M1-

projecting neurons had a significantly larger Vm modulation

than S2-projecting neurons (M1-projecting neurons, 1.51 ±

0.33 mV, n = 14; S2-projecting neurons, 0.83 ± 0.08 mV, n =

13; p = 0.038) (Figure 4C). Averaging Vm at random timing during

whisking periods (shuffling) revealed that fast Vm modulation

phase-locked to whisker protraction was significantly above

chance level in M1-projecting neurons, but not in S2-projecting

neurons (shuffled Vm modulation: M1-projecting neurons,

0.76 ± 0.05mV, n = 14, p = 0.0001 shuffled compared to protrac-

tion-triggered Vm modulation of M1-projecting neurons; S2-pro-

jecting neurons, 0.59 ± 0.07 mV, n = 13; p = 0.17 shuffled

compared to protraction-triggered Vm modulation of S2-projec-

ting neurons) (Figure 4C). Furthermore, during whisking, the Vm

was more highly correlated with whisker position in M1-projec-

ting neurons compared to S2-projecting neurons (absolute

peak value of cross-correlation: M1-projecting neurons, 0.33 ±

0.02, n = 12; S2-projecting neurons, 0.26 ± 0.02, n = 11; p =

0.012) (Figures 4D and 4E). Averaging the Vm aligned to whisking

phase, through the Hilbert transform, revealed stronger phase-

locked Vm modulation in M1-projecting neurons (the phase-

locked Vm amplitude: M1-projecting neurons, 2.15 ± 0.33 mV,

n = 14; S2-projecting neurons, 1.18 ± 0.14 mV, n = 13; p =

0.023) (Figures 4F and 4G). The most depolarized phase of the

Vm relative to the whisker movement varied across cells, but it

was mostly in more retracted phases in M1-projecting neurons

(Figure 4H). Taken together, these results suggest that fast Vm

1480 Neuron 80, 1477–1490, December 18, 2013 ª2013 Elsevier Inc.

modulation phase-locked to whisker

movement ismore prominent inM1-projec-

ting neurons compared to S2-projecting

neurons. Thus, M1-projecting neurons are

more specialized for encoding whisker

position on the millisecond timescale.

Target-Specific Responses of S1Projection Neurons to PassiveTactile SensationTo investigate how sensory information is

transmitted from S1 to M1 and S2, we

next examined Vm changes evoked by tactile stimulation. In

response to a brief (1 ms) passive deflection of the contralateral

C2 whisker during quiet wakefulness, only a subset of the pro-

jection neurons in the C2 barrel column fired APs with a proba-

bility of more than 10% within 0.1 s after stimulation (4 out of 15

cells among M1-projecting neurons, 2 out of 15 among S2-pro-

jecting neurons) or within 0.3 s after stimulation (5 out of 15 cells

among M1-projecting neurons, 4 out of 15 among S2-projecting

neurons). Averaged across all recordings, AP rates evoked by

passive C2 whisker deflection were generally low (evoked AP

rates within 0.1 s after stimulation, with spontaneous AP rates

subtracted: M1-projecting neurons, 0.87 ± 0.32 Hz, n = 15;

S2-projecting neurons, 0.68 ± 0.49 Hz, n = 15; p = 0.36; evoked

AP rates within 0.3 s after stimulation, with spontaneous AP

rates subtracted: M1-projecting neurons, 0.32 ± 0.12 Hz, n =

15; S2-projecting neurons, 0.45 ± 0.21 Hz, n = 15; p = 0.89).

These results suggest a sparse coding of passive tactile sensa-

tion in both M1-projecting and S2-projecting layer 2/3 pyramidal

neurons.

Although the total number of evoked APs quantified over 0.1 s

or 0.3 s after passive whisker stimulation was similar in the two

types of projection neurons, we found striking differences in

the timing of evoked APs. M1-projecting neurons transiently

fired APs with short latency, whereas S2-projecting neurons

fired APs after a longer delay but for a longer duration (Figures

5A and 5B). Quantified within the first 20 ms after stimulation

and with spontaneous AP firing rates subtracted, deflection

of the C2 whisker evoked AP firing at 1.65 ± 0.69 Hz (n = 15)

in M1-projecting neurons and at 0.05 ± 0.13 Hz (n = 15) in

Page 5: Membrane Potential Dynamics of Neocortical Projection ... · Membrane Potential Dynamics of Neocortical Projection Neurons Driving Target-Specific Signals Takayuki Yamashita,1,*

Figure 3. Behavioral Modulation of Vm Dynamics in M1- and S2-Projecting Neurons

(A) Whisker angle (green) was monitored during whole-cell recordings of Vm from an M1-projecting (M1-p, blue) and an S2-projecting (S2-p, red) neuron.

(B) Grand average fast Fourier transform (FFT) spectra of Vm, showing a reduced amplitude in low-frequency Vm fluctuations during whisking.

(C) Standard deviation (SD) of Vm and 1–5 Hz area of the Vm FFT during quiet wakefulness (Q) and whisking (W). M1-projecting neurons have larger slow Vm

fluctuations during quiet wakefulness compared to S2-projecting neurons.

(D) Mean Vm in both cell-types depolarized during whisking.

(E) Spontaneous AP rate of M1- and S2-projecting neurons.

(F) Changes of AP rate induced by whisking. In a subset of cells, AP rate was changed by more than 0.1 Hz by whisking, indicated as Up (green, W > Q), Down

(purple, W < Q) or Unchanged (gray, < 0.1 Hz change). A significant fraction of S2-projecting neurons reduced firing during whisking.

Lightly colored lines in (C), (D), and (E) correspond to individual cells. Bold lines and filled circles with error bars show mean ± SEM.

Neuron

Target-Specific Membrane Potential Dynamics

S2-projecting neurons, which was significantly different (p =

0.007) (Figures 5B and 5C). In contrast to the immediate phasic

firing of M1-projecting neurons, a single brief whisker deflection

evoked delayed firing in S2-projecting neurons that remained

elevated for hundreds of milliseconds (evoked change in AP

rates in the 0.1–0.3 s after the stimulation: M1-projecting neu-

rons, 0.05 ± 0.04 Hz, n = 15; S2-projecting neurons, 0.32 ±

0.10 Hz, n = 15; p = 0.019) (Figures 5B, 5C, and S3A–S3E).

Ne

To examine the mechanisms underlying such a temporal dif-

ference in firing patterns, we measured the subthreshold post-

synaptic potentials (PSPs) evoked by stimulation. Prestimulus

Vm did not differ between M1- and S2-projecting neurons

(M1-projecting neurons, –66.0 ± 0.9 mV, n = 15; S2-projecting

neurons, –64.8 ± 1.0 mV, n = 15; p = 0.54). Compared to S2-pro-

jecting neurons, the PSPs in M1-projecting neurons had a

shorter onset latency (M1-projecting neurons, 6.9 ± 0.4 ms,

uron 80, 1477–1490, December 18, 2013 ª2013 Elsevier Inc. 1481

Page 6: Membrane Potential Dynamics of Neocortical Projection ... · Membrane Potential Dynamics of Neocortical Projection Neurons Driving Target-Specific Signals Takayuki Yamashita,1,*

Figure 4. Fast Vm Modulation Phase-Locked to Whisker Movement in M1-Projecting Neurons

(A) Whisker angle (green) wasmonitored at 500 Hz during whole-cell recordings of Vm from anM1-projecting (M1-p, blue) and an S2-projecting (S2-p, red) neuron

during whisking periods.

(B) Whisker protraction-triggered average Vm traces (Normal) and average Vm traces aligned at random timings during whisking (Shuffled, superimposed) in M1-

and S2-projecting neurons (same neurons as shown in A). The timing of peak whisker protraction is indicated by a dashed line.

(legend continued on next page)

Neuron

Target-Specific Membrane Potential Dynamics

1482 Neuron 80, 1477–1490, December 18, 2013 ª2013 Elsevier Inc.

Page 7: Membrane Potential Dynamics of Neocortical Projection ... · Membrane Potential Dynamics of Neocortical Projection Neurons Driving Target-Specific Signals Takayuki Yamashita,1,*

Neuron

Target-Specific Membrane Potential Dynamics

n = 15; S2-projecting neurons, 8.4 ± 0.4 ms, n = 15; p = 0.022), a

larger slope in their rising phase (M1-projecting neurons, 1.34 ±

0.19 V/s, n = 15; S2-projecting neurons, 0.76 ± 0.12 V/s, n = 15;

p = 0.021) and larger amplitude (M1-projecting neurons, 11.2 ±

0.9 mV, n = 15; S2-projecting neurons, 8.1 ± 0.8 mV, n = 15;

p = 0.018) (Figures 5D and 5E). Thus, the short-latency AP firing

responses in M1-projecting neurons are driven by faster and

larger postsynaptic depolarization. M1-projecting neurons

therefore appear to be specialized for delivering rapid transient

output signals evoked by passive stimulation.

Because S2-projecting neurons fired more APs than M1-pro-

jecting neurons at late times (0.1–0.3 s) after whisker deflection,

we also analyzed Vm over this period. In addition to the early de-

polarizing PSP evoked by whisker deflection, the Vm of both M1-

and S2-projecting neurons also had a long-lasting depolarizing

component to the sensory response. However, we did not find

any difference comparing M1- and S2-projecting neurons in

the average Vm quantified 0.1–0.3 s after whisker deflection

(M1-projecting neurons, –64.2 ± 0.8 mV, n = 15; S2-projecting

neurons, –63.8 ± 1.1 mV, n = 15; p = 0.95) (Figures S3A and

S3B). The late Vm depolarization in both M1- and S2-projecting

neurons was enhanced on trials in which the mouse actively

began whisking after the whisker stimulus compared to trials in

which themouse did not move its whiskers after stimulation (Fig-

ures S3C and S3E) (Sachidhanandam et al., 2013). However, late

AP firing was not significantly modulated by stimulus-evoked

whisking (Figure S3D). Specifically, the late AP firing in S2-pro-

jecting neurons remained significantly higher than forM1-projec-

ting neurons on the subset trials in which the mouse did not

actively move its whiskers following whisker stimulus (M1-pro-

jecting neurons, 0.02 ± 0.05 Hz, n = 15; S2-projecting neurons,

0.25 ± 0.09 Hz, n = 15; p = 0.0025) (Figure S3D). Thus, although

late Vm depolarization depended upon stimulus-evoked whisk-

ing, the late AP firing in S2-projecting neurons did not. Therefore,

the average Vm does not provide a simple explanation for the

enhanced late AP firing in S2-projecting neurons.

In some experiments, we applied passivewhisker stimuli when

themicewere actively whisking in air (Figure 5F). Duringwhisking

periods, the amplitude of PSPs evoked by passive stimulation in

M1-projecting neurons was reduced to 66.2% ± 8.4% of the

response during quiet wakefulness (n = 8, p = 0.016), but whisk-

ing did not impact PSP amplitude in S2-projecting neurons

(PSP amplitude: 99.2% ± 18.7%, n = 8, p = 0.55) (Figures 5F,

5G, and S3F). PSPs evoked during whisking periods showed

prominently slower rising slopes in M1-projecting neurons

(0.31 ± 0.08 V/s, n = 8, p = 0.008 compared to quiet wakefulness),

but there was no significant slowing of PSP rise for S2-projecting

neurons (0.43 ± 0.11 V/s, n = 8, p = 0.075 compared to quiet

(C) M1-projecting neurons had significant Vm modulations quantified at ± 50 ms

S2-projecting neurons did not. Lightly colored lines correspond to individual cell

(D) Cross-correlograms between whisker angle and Vm during whisking (same n

(E) The Vm of M1-projecting neurons had significantly larger absolute peak corre

(F) Averaged Vm traces in function of the phase of whisking cycles (same neuron

(G) M1-projecting neurons had larger amplitudes of phase-locked Vm fluctuation

(H) The amplitude and most depolarized phase of Vm relative to the whisker mov

neurons). Most of M1-projecting neurons depolarized at a retracted phase.

Each open circle in (E) and (G) corresponds to a single cell. Filled circles with err

Ne

wakefulness) (Figures 5F and S3F). These results suggest that

internal brain states and whisking behavior differently affect

sensory responses among M1- and S2-projecting neurons,

with whisking strongly suppressing whisker-deflection-evoked

signaling to M1.

Target-Specific Responses of S1 Projection Neurons toActive TouchSensory responses to passively applied whisker deflections

were thus suppressed during active whisking periods in M1-pro-

jecting, but not in S2-projecting neurons, suggesting that S2-

projecting neurons might preferentially respond to tactile input

during active exploratory whisking. To examine this possibility

in detail, we recorded Vm changes evoked by active touch,

when the whisking mouse repetitively palpates an object

(Crochet and Petersen, 2006; Crochet et al., 2011). Similarly to

passive stimulation, active touch evoked APs with more than

10% probability in only a fraction of recorded cells (2 out of 19

cells among M1-projecting neurons, 7 out of 20 among S2-pro-

jecting neurons), suggesting sparse coding of active touch in

projection neurons. Touch-evoked AP firing occurred either dur-

ing the whisker-object contact (duration 24.8 ± 1.5 ms, n = 39

cells) or shortly after detachment of the whisker from the object.

There were prominent differences in the timing of AP firing

comparing M1-projecting and S2-projecting neurons during a

bout of whisking with active touch. All of the M1-projecting neu-

rons that showed touch-evoked APs (n = 9) responded best to

the first whisker-object contact, and AP firing was then sup-

pressed during subsequent touches (Figures 6A and 6B). In

contrast, all of the S2-projecting neurons that showed touch-

evoked APs with more than 10% probability (n = 7) responded

robustly to each whisker-object contact (Figures 6A and 6B).

Across all recorded cells, the number of APs within 50 ms after

whisker-object contact in M1-projecting neurons was pro-

foundly decreased at short intercontact intervals (ICIs, defined

as the time from the end of the preceding touch to the touch

onset) (AP number per touch: 0.15 ± 0.11 with ICIs > 0.2 s;

0.02 ± 0.01 with ICIs < 0.08 s; n = 19; p = 0.008), whereas active

touch-evoked AP firing in S2-projecting neurons was resistant to

shortening of ICIs (AP number per touch: 0.14 ± 0.05 with ICIs >

0.2 s; 0.17 ± 0.06 with ICIs < 0.08 s; n = 20; p = 0.33) (Figures 6C

and 6D).

Consistent with the strong depression of AP firing in M1-pro-

jecting neurons during repetitive active touch, touch-evoked

PSPs in M1-projecting neurons were also strongly depressed

in a frequency-dependent manner (Figures 6E, 6F, and S4) and

the peak Vm hyperpolarized at short ICIs (Figures 6E, 6G, and

S4). The frequency-dependent depression of touch-evoked

around the peak of whisker protraction (N) compared to shuffled data (S), but

s. Bold lines and filled circles with error bars show mean ± SEM.

eurons as shown in A).

lations with whisker angle compared to S2-projecting neurons.

s as shown in A).

than S2-projecting neurons.

ement for individual neurons (blue, M1-projecting neurons; red, S2-projecting

or bars in (E) and (G) indicate mean ± SEM.

uron 80, 1477–1490, December 18, 2013 ª2013 Elsevier Inc. 1483

Page 8: Membrane Potential Dynamics of Neocortical Projection ... · Membrane Potential Dynamics of Neocortical Projection Neurons Driving Target-Specific Signals Takayuki Yamashita,1,*

Figure 5. Distinct Vm Changes in M1- and S2-Projecting Neurons Evoked by Passively Applied Stimuli

(A) Brief (1 ms) whisker stimuli (dashed lines) were delivered during recording from an M1- or an S2-projecting neuron. Example trials showing Vm around the

stimulus time (superimposed, above) and corresponding AP raster plots (below).

(B) Grand average peristimulus time histogram (PSTH) showing spike-timing differences in the response of M1- and S2-projecting neurons.

(C) M1-projecting neurons fired more APs in initial phases of responses (0–20 ms after whisker deflection), whereas at later times (0.1–0.3 s after whisker

deflection) S2-projecting neurons fired more. The spontaneous AP rate for each cell was subtracted to give the evoked AP rate.

(D) Grand average of subthreshold responses from M1- (blue) and S2- (red) projecting neurons (superimposed with the baseline Vm subtracted).

(E) Subthreshold responses in M1-projecting neurons had shorter latencies, larger rising slopes and larger amplitudes than those in S2-projecting neurons.

(F) Grand average of subthreshold responses during quiet andwhisking periods (superimposed). Data were included only whenmore than 10 traces for both quiet

and whisking periods were obtained in each cell.

(legend continued on next page)

Neuron

Target-Specific Membrane Potential Dynamics

1484 Neuron 80, 1477–1490, December 18, 2013 ª2013 Elsevier Inc.

Page 9: Membrane Potential Dynamics of Neocortical Projection ... · Membrane Potential Dynamics of Neocortical Projection Neurons Driving Target-Specific Signals Takayuki Yamashita,1,*

Neuron

Target-Specific Membrane Potential Dynamics

PSPs was smaller in S2-projecting neurons (magnitude of PSP

depression with ICIs < 0.08 s: 59% ± 4% for S2-projecting neu-

rons; 70% ± 3% for M1-projecting neurons; p = 0.045). At short

ICIs the peak Vm depolarized in S2-projecting neurons, whereas

it hyperpolarized in M1-projecting neurons (normalized peak Vm

amplitude with ICIs < 0.08 s: 1.32 ± 0.14 for S2-projecting neu-

rons; 0.84 ± 0.05 for M1-projecting neurons; p = 0.0002) (Figures

6E, 6G, and S4). Thus, strong frequency-dependent depression

of PSPs preferentially shut down AP generation in M1-projecting

neurons during repetitive touch, whereas summation of PSPs

exhibiting less frequency-dependent depression in S2-projec-

ting neurons allowed sustained depolarization and firing during

active touch.

DISCUSSION

Through whole-cell recordings from cortical projection neurons

in layer 2/3 of the C2 barrel column in S1 of behaving mice, we

have obtained insights into Vm dynamics of projection neurons

with different cortical targets. Although a previous Ca2+ imaging

study suggested that training of mice for texture discrimination

or object localization is needed for the generation of functionally

distinct information streams in mouse barrel cortex (Chen et al.,

2013), our Vm measurements with high temporal resolution sug-

gest that reward-associated training is not required to establish

differential activity patterns of M1- and S2-projecting neurons.

Rather, their distinct intrinsic membrane properties and their dif-

ferential integration of synaptic input cause temporal and rate

differences in their firing patterns in naive mice. M1- and S2-pro-

jecting neurons could therefore belong to different genetically

defined cell-types, a question that should be further investigated

in future detailed anatomical and molecular studies. Given the

functional specificity of sensory information conveyed by M1-

and S2-projecting neurons, our results suggest an interesting

analogy to the classical dorsal (‘‘where’’) and ventral (‘‘what’’)

streams of the visual system (Goodale and Milner, 1992; Nassi

and Callaway, 2009) and the auditory system (Rauschecker

and Tian, 2000). As further discussed below, we propose a

similar large-scale organizing principle for the somatosensory

system of mice with a ‘‘dorsal’’ S1 to M1 projection signaling

detection and localization of objects, acting together with a

‘‘ventral’’ S1 to S2 projection encoding object features (Figure 7).

Cellular Mechanisms Underlying Target-SpecificSensory CodingIn general, the early phase of firing responses ofM1- and S2-pro-

jecting neurons after passive whisker stimulation or active touch

was well-correlated with their subthreshold Vm dynamics. The

short-latency firing of M1-projecting neurons evoked by passive

stimulation was driven by fast and large PSPs (Figure 5). The

frequency-dependent suppression of touch-evoked APs in M1-

projecting neurons was associated with frequency-dependent

depression of PSPs and frequency-dependent hyperpolarization

(G) PSP amplitudes were normalized to those obtained during quiet wakefulness

neurons.

Open circles and lightly colored lines in (C), (E), and (G) correspond to individual

Figure S3.

Ne

of peak Vm (Figure 6). In contrast, longer latency firing of S2-

projecting neurons after passive stimuli was driven by longer

latency, more slowly rising PSPs (Figure 5). Furthermore, fre-

quency-dependent summation of PSPs in S2-projecting neu-

rons caused their sustained depolarization and rhythmic firing

during repetitive touch (Figure 6). The longer membrane time

constants of S2-projecting neurons (Figure 2) might contribute

to the enhanced PSP summation in S2-projecting neurons

during high-frequency repetitive touch, whereas the shorter

membrane time constants in M1-projecting neurons will reduce

temporal summation of PSPs. Thus, processing of sensory-

evoked synaptic input by these projection neurons contributes

to the generation of their target-specific sensory coding. Our

findings that M1- and S2-projecting neurons receive different

spontaneous synaptic input (Figure 3) and different sensory

inputs (Figures 5 and 6) suggest that they are embedded in

different neuronal networks. In future studies, it will therefore

be of great interest to examine the differential synaptic connec-

tivity of M1- and S2-projecting neurons with respect to the

surrounding neocortical microcircuit and their long-range inputs

from other cortical and subcortical brain regions.

Unlike the early component of evoked responses, the later

phase of subthreshold Vm changes after passive stimulation

did not account for the target-specific differences in AP firing.

Although S2-projecting neurons have a significantly higher firing

rate than M1-projecting neurons at 0.1–0.3 s after whisker

deflection (Figures 5A–5C), the mean subthreshold Vm in this

late period was not significantly different among these projection

neurons (Figure S3). APs in excitatory layer 2/3 neurons are typi-

cally driven by large, rapid, and cell-specific depolarizations

(Poulet and Petersen, 2008; Gentet et al., 2010), which might

not always significantly impact the average somatic Vm due to

their asynchronous occurrence and short duration. Large rapid

depolarizations driving APs in S2-projecting neurons during the

late response period could result from large synaptic inputs spe-

cifically converging on S2-projecting neurons or they could be

driven by active processes in dendrites (Larkum et al., 2009;

Branco and Hausser, 2010; Xu et al., 2012; Petersen and

Crochet, 2013). Interestingly, it has recently been suggested

that AP generation can be suppressed by dendritic inhibition

without changes in subthreshold Vm recorded from the soma

(Palmer et al., 2012). Differences in dendritic excitability or in

the spatiotemporal pattern of excitatory and inhibitory synaptic

input across the dendritic arborizations ofM1- and S2-projecting

neurons might therefore underlie the differences in late spiking

evoked by passive whisker stimulation.

Dorsal S1 to M1 Processing Pathway for SensorimotorCoordinationSensory signals evoked by whisker deflection are transmitted

from S1 barrel columns to whisker M1 within milliseconds (Fere-

zou et al., 2007), forming a dorsal stream of sensory information

(Figure 7). Stimulation of whisker M1 causes rhythmic whisker

. Whisking induced a significant reduction in PSP amplitude in M1-projecting

cells. Bold lines and filled circles with error bars show mean ± SEM. See also

uron 80, 1477–1490, December 18, 2013 ª2013 Elsevier Inc. 1485

Page 10: Membrane Potential Dynamics of Neocortical Projection ... · Membrane Potential Dynamics of Neocortical Projection Neurons Driving Target-Specific Signals Takayuki Yamashita,1,*

Figure 6. Active Touch Responses of M1- and S2-Projecting Neurons

(A) An object was presented to an awakemouse during simultaneous whole-cell recording (Vm) and high-speedwhisker filming. Themouse could actively contact

the object by whisking toward it (gray bars, onset and duration of contact). During repetitive touch, M1-projecting neurons (M1-p) exhibited only transient AP

firing, whereas S2-projecting (S2-p) neurons responded equally to each touch.

(B) Additional examples of active touch responses. Example Vm and whisker angle traces during repetitive active touch from another M1-projecting and another

S2-projecting neuron (both these neurons are different to those shown in A).

(C) M1-projecting neurons, but not S2-projecting neurons, showed strong depression in touch-evoked AP responses at short intercontact intervals (ICIs). Data

are shown as mean ± SEM.

(legend continued on next page)

Neuron

Target-Specific Membrane Potential Dynamics

1486 Neuron 80, 1477–1490, December 18, 2013 ª2013 Elsevier Inc.

Page 11: Membrane Potential Dynamics of Neocortical Projection ... · Membrane Potential Dynamics of Neocortical Projection Neurons Driving Target-Specific Signals Takayuki Yamashita,1,*

Figure 7. Schematic Summary of Two

Distinct Neocortical Processing Streams in

the Whisker Somatosensory System

(A) Schematic diagram of dorsal (S1 to M1) and

ventral (S1 to S2) signaling pathways.

(B) Summary of possible functional roles for S1 to

M1 and S1 to S2 pathways. Passive tactile

sensation rapidly activates S1 to M1 signal flow,

possibly leading mice to initiate whisking. During

palpation of objects, the S1 to M1 pathway is

transiently activated, and this might help object

detection, object localization, and cause motor

commands for optimization of whisker move-

ments. Repetitive active touch sensation contin-

uously activates the S1 to S2 pathway, which

might be essential for object feature recognition.

Neuron

Target-Specific Membrane Potential Dynamics

protraction (Matyas et al., 2010). Therefore, excitatory signal

output from S1 to M1 through M1-projecting neurons may be

important for the initiation of whisking, so that, after a passive

stimulus is perceived, the mouse can actively explore to uncover

further sensory information relating to the stimulus. The short-

latency, transient firing of M1-projecting neurons we observed

after passive whisker deflection (Figure 5) is well-suited for

such a function of this pathway, signaling the detection of a sen-

sory stimulus and enabling a rapid motor response if required.

On the other hand, the Vm dynamics and sensory responses of

M1-projecting neurons were strongly affected by internal brain

states and whisking behavior. Firing activities in a subset of

M1-projecting neurons were modulated by whisking (Figures

3E and 3F) and whisking phase-locked Vm modulation was

prominent in M1-projecting neurons (Figure 4). Passive whisker

deflection evoked smaller and slower PSPs when the mice

were actively whisking (Figures 5F, 5G, and S3F), which is in a

good agreement with the attenuated S1 to M1 signal flow during

whisking imaged with voltage-sensitive dyes (Ferezou et al.,

2007). Whisking phase-locked Vm modulation and reduced sen-

sory responses during active states might reflect state-depen-

dent recruitment of different types of GABAergic neurons (Gentet

et al., 2010, 2012; Mateo et al., 2011), state-dependent differ-

ences in ongoing thalamic AP firing rates (Poulet et al., 2012)

and differences in neuromodulation (Lee and Dan, 2012). Most

strikingly, during repetitive active touch, M1-projecting neurons

show only transient firing with frequency-dependent suppres-

sion of APs (Figure 6). These observations suggest that M1-pro-

jecting neurons are specialized for detecting stimulus onset,

rather than encoding object quality during active states. Further-

more, whisking phase-locked Vm fluctuations in M1-projecting

neurons (Figure 4) together with their strong responses to the

initial whisker-object contact during active touch (Figures 6A–

6C) might enable these neurons to compute the location of an

object by combining whisker position information with the timing

of whisker-object contact (Curtis and Kleinfeld, 2009). In agree-

ment with such a hypothesis, a recent calcium imaging study in

(D) Average PSTHs for long (>0.2 s) and short (<0.08 s) ICIs.

(E) Grand average of touch-evoked PSPs in M1- and S2-projecting neurons for lon

(F and G) PSP amplitude (F) and peak DVm (G) show stronger frequency-depen

normalized to long ICI).

Filled circles with error bars show mean ± SEM. See also Figure S4.

Ne

trained mice suggested that activity of M1-projecting neurons

discriminated object locations better than S2-projecting

neurons (Chen et al., 2013). Taken together, M1-projecting neu-

rons appear well-suited to play a key role in spatial attention and

sensorimotor coordination by generating a dorsal ‘‘where’’ pro-

cessing stream related to stimulus detection and possibly object

location (Figure 7).

Ventral S1 to S2 Processing Pathway for SensoryPerceptionSensory responses of S2-projecting neurons in S1 were quite

different from M1-projecting neurons. PSPs evoked by passive

whisker stimulation and recorded in S2-projecting neurons had

smaller amplitudes with longer latency and slower rate of rise,

driving APswith longer peak-latency compared toM1-projecting

neurons (Figure 5). The internal brain state andwhisking behavior

did not affect the amplitude of evoked PSPs in S2-projecting

neurons, unlike the state-dependent modulation observed for

M1-projecting neurons. The PSP response to passive whisker

stimulation in S2-projecting neurons was equal in amplitude dur-

ing both quiet wakefulness and active whisking (Figures 5F and

5G). Furthermore, during repetitive active touch, S2-projecting

neurons robustly responded to each touch irrespective of

intercontact interval (Figure 6). This is in clear contrast to the

response properties of M1-projecting neurons that only tran-

siently signal the onset of repetitive touch. The robust firing of

S2-projecting neurons during active touch would be useful for

encoding information about object quality accumulated over

time across repetitive whisker-object contacts. In agreement

with this hypothesis, a recent calcium imaging study suggested

that the activity of S2-projecting neurons discriminated textures

during whisking better than M1-projecting neurons in trained

mice (Chen et al., 2013). The S1 to S2 signaling pathway

may thus represent a ventral ‘‘what’’ information stream, impor-

tant for mice to recognize the surface features of objects

gathered during repetitive whisker palpation (Figure 7). How-

ever, in considering such a ‘‘ventral’’ stream of information, it is

g (>0.2 s) and short (<0.08 s) ICIs. The touch onset is indicated by a gray arrow.

dent depression of postsynaptic responses in M1-projecting neurons (values

uron 80, 1477–1490, December 18, 2013 ª2013 Elsevier Inc. 1487

Page 12: Membrane Potential Dynamics of Neocortical Projection ... · Membrane Potential Dynamics of Neocortical Projection Neurons Driving Target-Specific Signals Takayuki Yamashita,1,*

Neuron

Target-Specific Membrane Potential Dynamics

important to note that in addition to inputs from S1, the rodent S2

region receives direct thalamocortical input presumably related

to whisker behavior (Spreafico et al., 1987; Lee and Sherman,

2008). In future studies it will therefore be important to examine

how sensory information is integrated and processed by

neuronal microcircuits in S2.

ConclusionsOur results reveal that projection neurons in layer 2/3 of mouse

barrel cortex exhibit important target-specific differences with

respect to intrinsic membrane properties, synaptic inputs

(PSPs), and AP firing patterns. We find that the interplay between

frequency-dependent depression and summation of PSPs is a

key determinant for S1 projection neurons to segregate

incoming sensory information. Thus, processing of synaptic

input by projection neurons dynamically controls large-scale

connectivity between distinct brain areas. We propose a dorsal

‘‘where’’ S1 to M1 pathway driving rapid but transient signaling,

well-suited for object detection and localization, whereas a

ventral ‘‘what’’ S1 to S2 pathway integrates over longer time-

scales helping to identify object features during repetitive touch.

Our data therefore begin to define cellular mechanisms for

target-specific corticocortical signaling, but future experiments

must uncover the differential functional synaptic connectivity

of these different types of projection neurons and the roles

of target-specific signaling during behavioral tasks probing

different aspects of tactile sensory perception.

EXPERIMENTAL PROCEDURES

All experimental procedures were approved by the Swiss Federal Veterinary

Office.

Animal Preparation and Surgery

Male 4- to 8-week-old C57BL6J mice, GAD67-GFP mice (Tamamaki et al.,

2003) and Gad2-Cre mice (Taniguchi et al., 2011) crossed to lox-stop-lox

tdTomato reporter mice (Madisen et al., 2010) were implanted with a light-

weight metal head-holder and a recording chamber under isoflurane anes-

thesia. The locations of the left S1-C2 barrel column and the left whisker-S2

region were functionally identified through intrinsic optical imaging under light

isoflurane anesthesia (Ferezou et al., 2007). The bone over S1 and S2 regions

was gently scraped in order to improve intrinsic optical imaging and protected

by a thick layer of superglue after imaging. A small craniotomy (<0.5 mm in

diameter) was opened over left M1 (1 mm anterior, 1 mm lateral from Bregma)

and/or left S2 and then CTB conjugated with Alexa-Fluor 488 or 594 (0.5%,

weight/volume, Invitrogen) was injected using a glass pipette (tip inner diam-

eter = 20–30 mm). Injection volume was 100 nl for M1 and 50 nl for S2 at the

depths of 300 and 800 mm, giving a total volume of 200 nl for M1 and 100 nl

for S2.Micewere analyzed 5–15days afterCTB injection.Micewerehabituated

to head restraint (three to five sessions, one session per day) before recording.

Electrophysiology and Quantification of Whisker Behavior in Awake

Head-Restrained Mice

Whole-cell patch-clamp recordings were targeted to cell bodies of CTB-

labeled neurons in the center of the C2 barrel column (as identified with

intrinsic optical signals) of C57BL6J mice under visual control, essentially as

previously described (Gentet et al., 2010, 2012). We used a custom built

two-photon microscope under the control of Helioscan software (Langer

et al., 2013). Recordings were made at the subpial depth of 110–245 mm,

and the recording depth for M1- and S2-projecting neurons was similar for

each data set that we analyzed. The recording pipettes had resistances of

5–7 MU and were filled with a solution containing (in mM): 135 potassium glu-

1488 Neuron 80, 1477–1490, December 18, 2013 ª2013 Elsevier Inc

conate, 4 KCl, 10 HEPES, 10 sodium phosphocreatine, 4 MgATP, 0.3 Na3GTP

(adjusted to pH 7.3 with KOH). For targeting CTB-labeled neurons, Alexa 488

or 594 (1–20 mM) was added to the pipette solution, depending on the color of

the targeted cells. The Vm was not corrected for liquid junction potential. Cur-

rent injection experiments shown in Figure 2 were performed in the initial

periods of recordings (within 5 min of break-in) after series resistance was

compensated by the bridge-balance function of the patch-clamp amplifier

(MultiClamp 700B, Molecular Devices).

All whiskers except for C2 were trimmed before the recording session. Short

(20 s) sweeps were recorded while the whisker behavior of the mouse was

simultaneously filmed using a high-speed camera (MotionPro, Redlake) oper-

ating at 500 frames per second. The behavioral images were synchronized to

the electrophysiological recording through TTL pulses. Whisker movements

and whisker-object contacts were quantified off-line (Crochet and Petersen,

2006; Crochet et al., 2011). To detect contacts of the whisker during active

touch, a piezo sensor was used for the object and the onset of whisker con-

tacts was taken from the piezo sensor signals, which preceded the initial

deflection of the whisker identified in the high-speed whisker filming by

�2 ms (Crochet et al., 2011). The offset of whisker-object contact was identi-

fied through visual inspection of the behavioral images. For passive whisker

stimulation, we used a brief (1 ms) magnetic pulse to elicit a small and rapid

single deflection of the right C2 whisker transmitted by a small metal particle

glued on the whisker (Crochet and Petersen, 2006). Each pulse was delivered

every 2 s in each 20 s sweep, and the intersweep interval was more than 2min.

The animal was placed over an electromagnetic coil (10 cm inside diameter)

and the right C2 whisker was positioned above the center of the coil.

Data Analysis

Input resistances, AP threshold, AP half-width, and peak/trough ratio of AP

waveforms were measured as previously described (Gentet et al., 2010).

Mean Vm and average spontaneous AP rates were computed as the mean

of periods totaling over 20 s for quiet periods and over 2 s for whisking periods

for each cell. Membrane time constants were estimated by fitting a single

exponential function to the initial onset of averaged Vm responses to hyperpo-

larizing current pulses (–100 pA), starting 1 ms after the onset of the current

pulse. For analysis of AP threshold and waveforms, APs were only included

if the time from the preceding spike wasmore than 50ms, andmore than three

overshooting APs obtained with zero current injection were averaged for each

cell. FFTs were computed as magnitudes in IgorPro (Wavemetrics) for 2 s seg-

ments of the recordings. The amplitude of low-frequency (1–5 Hz) Vm fluctua-

tions was calculated by integrating the computed FFTs from 1 to 5 Hz.

For Figures 4B and 4C, every recorded whisking cycle (that had a protraction

amplitude of more than 5� and peak rates of protraction and retraction >500�/s)was aligned at the peak of protraction and averaged, revealing both the mean

whiskermovement during awhisking cycle and themeanchange in Vm (Crochet

and Petersen, 2006; Poulet and Petersen, 2008). The modulation depth was

defined as the difference in the average Vm between the most positive and the

most negative peakwithin ±50ms from thepeak ofwhisker protraction. For Fig-

ures 4D and 4E, the cross-correlogram between whisker angle and Vm was

calculated for 1 s segments of the recordings during whisking and the absolute

peak values of the cross-correlationwithin ± 50mswere averaged for each cell.

For Figures4F–4H, thephaseof awhiskingcyclewasextractedusing theHilbert

transform (Hill et al., 2011) and corresponding Vm traces in function of the phase

were averaged for each cell. Phase-locked Vm amplitude was defined as the

difference in the average Vm trace between the most positive and the most

negative peak during the whole whisking cycle.

APs evoked by passive whisker stimulation were estimated by subtracting

spontaneous AP rate from the AP rate measured in the early (0–0.02 s) or

late (0.1–0.3 s) periods after the stimulation for each cell. Peristimulus time his-

tograms (PSTHs) were computed by counting AP number in each 10ms bin for

each cell and averaging the number across cells recorded, with the AP rate in

these grand average PSTHs shown in Hz. Intercontact interval (ICI) for active

touch was defined as the time difference between the end of the preceding

whisker-object contact and the onset of contact. The subthreshold sensory re-

sponses were obtained by averaging median-filtered Vm traces aligned at the

onset of the stimuli or touches (Crochet and Petersen, 2006; Crochet et al.,

2011). Baseline Vm was defined as the mean Vm at 0–5 ms before the onset.

.

Page 13: Membrane Potential Dynamics of Neocortical Projection ... · Membrane Potential Dynamics of Neocortical Projection Neurons Driving Target-Specific Signals Takayuki Yamashita,1,*

Neuron

Target-Specific Membrane Potential Dynamics

The amplitude of PSPs was defined as the difference between the baseline Vm

and the peak Vm of averaged traces. The slope of the PSP was calculated from

a linear fit to the 20%–80% rise-time period. The PSP onset latency was

computed by detecting the time point after the stimulation period or the touch

onset at which the averaged Vm trace deflects from the baseline regression line

(Crochet and Petersen, 2006). Onset latency was defined as the time differ-

ence between this point and the time of the onset of stimuli or touches. For

normalizing touch-evoked PSPs, the peak amplitude and baseline potential

of averaged Vm traces with ICIs >0.2 s were used as a standard.

All values are mean ± SEM. Statistical testing using Wilcoxon rank-sum test

(unpaired data) or Wilcoxon signed rank test (paired data) was performed in

IgorPro unless otherwise noted.

Single-Cell Electroporation

In vivo electroporation of a single CTB-labeled cell per C57BL6J mouse was

carried out under isoflurane anesthesia with slight modifications from a previ-

ously described protocol (Judkewitz et al., 2009). In brief, glass pipettes having

resistances of 10–17 MU were filled with the pipette solution (see above) to

which 100 mMAlexa 488 and 100 ng/ml of pCAG-EGFP plasmid DNA (Matsuda

and Cepko, 2004) (Addgene plasmid 11150, kindly provided by Connie Cepko)

were added. Using shadow imaging under two-photon microscopy, the pi-

pettes were inserted through the intact dura and brought into close contact

with the cell body of the CTB-labeled neuron and 50 pulses of negative voltage

step (0.5 ms, –12 V) were delivered at 50 Hz using a pulse generator (Axopo-

rator 800A, Molecular Devices). The craniotomy was then covered with a sili-

cone elastomer (Kwik-Cast, WPI) and animals were returned to their home

cages for 4–7 days before perfusion.

Histology

After transcardial perfusion and postfixation for 2–4 hr using paraformalde-

hyde (4%, in 0.1 M phosphate buffer [PB], pH 7.3–7.4), we cut the fixed brains

in slices on a vibratome (section thickness: 60 or 100 mm). For epifluorescence

and confocal imaging, slices were mounted on slides using DABCO or Mowiol

after staining cellular nuclei by incubation for 10–20min with DAPI (2 mM in PB).

For cell reconstruction, slices were washed in PBS for 5 min, and endogenous

peroxidases were then quenched by 15 min incubation with 0.3% H2O2. The

slices were subsequently washed three times with 2% normal goat serum

(NGS) and 0.5% Triton-X and then incubated with primary anti-GFP antibody

(rabbit polyclonal, 1:500) together with 2% NGS and 0.5% Triton-X for 4 days

at 4�C. The slices were then washed with PBS containing 2% NGS and 0.5%

Triton-X and further incubated with biotinylated goat antibody against rabbit

IgG (1:500) together with 2% NGS and 0.5% Triton-X for 1.5 hr. The slices

were then rinsed in TBS (pH 8, 0.1 M, 0.9% NaCl) three times and were con-

jugated with avidin-biotinylated peroxidase following the manufacturer’s

instructions (Vectastain, Vector Labs). Slices were then washed three times

with TBS, and subsequently GFP-expressing neurons were visualized under

a reaction with 0.4 mg/ml DAB and 0.03% H2O2 for 10 min. The reaction

was stopped by rinsing the sections in TBS. Finally, the slices were mounted

on slides using Mowiol. Axonal and dendritic processes were subsequently

reconstructed from the serial sections using Neurolucida software (MBF

Bioscience).

For imaging tissue-cleared brains (Figures S1A and S1B), the fixed brains

were incubated for more than 7 days with ScaleA2 solution (Hama et al.,

2011), containing: 4 M urea, 10% glycerol, and 0.1% Triton-X. The brains

were then tangentially cut in half and the left hemisphere was imaged using

a two-photon microscope.

SUPPLEMENTAL INFORMATION

Supplemental Information includes four figures and can be found with this

article online at http://dx.doi.org/10.1016/j.neuron.2013.10.059.

ACKNOWLEDGMENTS

We thank Luc Gentet, Sylvain Crochet, Celine Mateo, Szabolcs Olah, Emma-

nuel Eggermann, and Varun Sreenivasan for discussion and Rudolf Kraftsik for

Ne

help with anatomical reconstruction. This work was supported by grants from

the Swiss National Science Foundation (to C.C.H.P. and E.W.), the Human

Frontier Science Program (to C.C.H.P.) and the European Research Council

(to C.C.H.P.). T.Y. is a JSPS Postdoctoral Fellow for Research Abroad.

Accepted: October 23, 2013

Published: December 18, 2013

REFERENCES

Aronoff, R., Matyas, F., Mateo, C., Ciron, C., Schneider, B.L., and Petersen,

C.C.H. (2010). Long-range connectivity of mouse primary somatosensory bar-

rel cortex. Eur. J. Neurosci. 31, 2221–2233.

Azouz, R., and Gray, C.M. (2000). Dynamic spike threshold reveals a mecha-

nism for synaptic coincidence detection in cortical neurons in vivo. Proc.

Natl. Acad. Sci. USA 97, 8110–8115.

Branco, T., and Hausser, M. (2010). The single dendritic branch as a funda-

mental functional unit in the nervous system. Curr. Opin. Neurobiol. 20,

494–502.

Brown, S.P., and Hestrin, S. (2009). Intracortical circuits of pyramidal neurons

reflect their long-range axonal targets. Nature 457, 1133–1136.

Carvell, G.E., and Simons, D.J. (1987). Thalamic and corticocortical connec-

tions of the second somatic sensory area of the mouse. J. Comp. Neurol.

265, 409–427.

Chen, J.L., Carta, S., Soldado-Magraner, J., Schneider, B.L., and Helmchen,

F. (2013). Behaviour-dependent recruitment of long-range projection neurons

in somatosensory cortex. Nature 499, 336–340.

Conte, W.L., Kamishina, H., and Reep, R.L. (2009). Multiple neuroanatomical

tract-tracing using fluorescent Alexa Fluor conjugates of cholera toxin subunit

B in rats. Nat. Protoc. 4, 1157–1166.

Crochet, S., and Petersen, C.C.H. (2006). Correlating whisker behavior with

membrane potential in barrel cortex of awakemice. Nat. Neurosci. 9, 608–610.

Crochet, S., Poulet, J.F.A., Kremer, Y., and Petersen, C.C.H. (2011). Synaptic

mechanisms underlying sparse coding of active touch. Neuron 69, 1160–1175.

Curtis, J.C., and Kleinfeld, D. (2009). Phase-to-rate transformations encode

touch in cortical neurons of a scanning sensorimotor system. Nat. Neurosci.

12, 492–501.

Diamond, M.E., von Heimendahl, M., Knutsen, P.M., Kleinfeld, D., and Ahissar,

E. (2008). ‘Where’ and ‘what’ in the whisker sensorimotor system. Nat. Rev.

Neurosci. 9, 601–612.

Feldmeyer, D., Brecht, M., Helmchen, F., Petersen, C.C.H., Poulet, J.F.A.,

Staiger, J.F., Luhmann, H.J., and Schwarz, C. (2013). Barrel cortex function.

Prog. Neurobiol. 103, 3–27.

Felleman, D.J., and Van Essen, D.C. (1991). Distributed hierarchical process-

ing in the primate cerebral cortex. Cereb. Cortex 1, 1–47.

Ferezou, I., Haiss, F., Gentet, L.J., Aronoff, R., Weber, B., and Petersen, C.C.H.

(2007). Spatiotemporal dynamics of cortical sensorimotor integration in

behaving mice. Neuron 56, 907–923.

Ferraina, S., Pare, M., and Wurtz, R.H. (2002). Comparison of cortico-cortical

and cortico-collicular signals for the generation of saccadic eye movements.

J. Neurophysiol. 87, 845–858.

Gentet, L.J., Avermann, M., Matyas, F., Staiger, J.F., and Petersen, C.C.H.

(2010). Membrane potential dynamics of GABAergic neurons in the barrel cor-

tex of behaving mice. Neuron 65, 422–435.

Gentet, L.J., Kremer, Y., Taniguchi, H., Huang, Z.J., Staiger, J.F., and

Petersen, C.C.H. (2012). Unique functional properties of somatostatin-ex-

pressing GABAergic neurons in mouse barrel cortex. Nat. Neurosci. 15,

607–612.

Glickfeld, L.L., Andermann, M.L., Bonin, V., and Reid, R.C. (2013). Cortico-

cortical projections in mouse visual cortex are functionally target specific.

Nat. Neurosci. 16, 219–226.

Goodale, M.A., and Milner, A.D. (1992). Separate visual pathways for percep-

tion and action. Trends Neurosci. 15, 20–25.

uron 80, 1477–1490, December 18, 2013 ª2013 Elsevier Inc. 1489

Page 14: Membrane Potential Dynamics of Neocortical Projection ... · Membrane Potential Dynamics of Neocortical Projection Neurons Driving Target-Specific Signals Takayuki Yamashita,1,*

Neuron

Target-Specific Membrane Potential Dynamics

Hama, H., Kurokawa, H., Kawano, H., Ando, R., Shimogori, T., Noda, H.,

Fukami, K., Sakaue-Sawano, A., and Miyawaki, A. (2011). Scale: a chemical

approach for fluorescence imaging and reconstruction of transparent mouse

brain. Nat. Neurosci. 14, 1481–1488.

Hill, D.N., Curtis, J.C., Moore, J.D., and Kleinfeld, D. (2011). Primary motor cor-

tex reports efferent control of vibrissa motion on multiple timescales. Neuron

72, 344–356.

Hoffmann, K.P., Bremmer, F., Thiele, A., and Distler, C. (2002). Directional

asymmetry of neurons in cortical areas MT and MST projecting to the NOT-

DTN in macaques. J. Neurophysiol. 87, 2113–2123.

Jarosiewicz, B., Schummers, J., Malik, W.Q., Brown, E.N., and Sur, M. (2012).

Functional biases in visual cortex neurons with identified projections to higher

cortical targets. Curr. Biol. 22, 269–277.

Judkewitz, B., Rizzi, M., Kitamura, K., and Hausser, M. (2009). Targeted single-

cell electroporation of mammalian neurons in vivo. Nat. Protoc. 4, 862–869.

Kiritani, T., Wickersham, I.R., Seung, H.S., and Shepherd, G.M. (2012).

Hierarchical connectivity and connection-specific dynamics in the corticospi-

nal-corticostriatal microcircuit in mouse motor cortex. J. Neurosci. 32, 4992–

5001.

Kitamura, K., Judkewitz, B., Kano, M., Denk, W., and Hausser, M. (2008).

Targeted patch-clamp recordings and single-cell electroporation of unlabeled

neurons in vivo. Nat. Methods 5, 61–67.

Langer, D., van ’t Hoff, M., Keller, A.J., Nagaraja, C., Pfaffli, O.A., Goldi, M.,

Kasper, H., and Helmchen, F. (2013). HelioScan: a software framework for

controlling in vivo microscopy setups with high hardware flexibility, functional

diversity and extendibility. J. Neurosci. Methods 215, 38–52.

Larkum, M.E., Nevian, T., Sandler, M., Polsky, A., and Schiller, J. (2009).

Synaptic integration in tuft dendrites of layer 5 pyramidal neurons: a new uni-

fying principle. Science 325, 756–760.

Le Be, J.V., Silberberg, G., Wang, Y., and Markram, H. (2007). Morphological,

electrophysiological, and synaptic properties of corticocallosal pyramidal cells

in the neonatal rat neocortex. Cereb. Cortex 17, 2204–2213.

Lee, C.C., and Sherman, S.M. (2008). Synaptic properties of thalamic and in-

tracortical inputs to layer 4 of the first- and higher-order cortical areas in the

auditory and somatosensory systems. J. Neurophysiol. 100, 317–326.

Lee, S.H., and Dan, Y. (2012). Neuromodulation of brain states. Neuron 76,

209–222.

Madisen, L., Zwingman, T.A., Sunkin, S.M., Oh, S.W., Zariwala, H.A., Gu, H.,

Ng, L.L., Palmiter, R.D., Hawrylycz, M.J., Jones, A.R., et al. (2010). A robust

and high-throughput Cre reporting and characterization system for the whole

mouse brain. Nat. Neurosci. 13, 133–140.

Mao, T., Kusefoglu, D., Hooks, B.M., Huber, D., Petreanu, L., and Svoboda, K.

(2011). Long-range neuronal circuits underlying the interaction between sen-

sory and motor cortex. Neuron 72, 111–123.

Margrie, T.W., Meyer, A.H., Caputi, A., Monyer, H., Hasan, M.T., Schaefer,

A.T., Denk, W., and Brecht, M. (2003). Targeted whole-cell recordings in the

mammalian brain in vivo. Neuron 39, 911–918.

Mateo, C., Avermann, M., Gentet, L.J., Zhang, F., Deisseroth, K., and

Petersen, C.C.H. (2011). In vivo optogenetic stimulation of neocortical excit-

atory neurons drives brain-state-dependent inhibition. Curr. Biol. 21, 1593–

1602.

Matsuda, T., and Cepko, C.L. (2004). Electroporation and RNA interference in

the rodent retina in vivo and in vitro. Proc. Natl. Acad. Sci. USA 101, 16–22.

Matyas, F., Sreenivasan, V., Marbach, F., Wacongne, C., Barsy, B., Mateo, C.,

Aronoff, R., and Petersen, C.C.H. (2010). Motor control by sensory cortex.

Science 330, 1240–1243.

Mesulam, M.M. (1998). From sensation to cognition. Brain 121, 1013–1052.

Morishima, M., and Kawaguchi, Y. (2006). Recurrent connection patterns of

corticostriatal pyramidal cells in frontal cortex. J. Neurosci. 26, 4394–4405.

Morishima, M., Morita, K., Kubota, Y., and Kawaguchi, Y. (2011). Highly differ-

entiated projection-specific cortical subnetworks. J. Neurosci. 31, 10380–

10391.

1490 Neuron 80, 1477–1490, December 18, 2013 ª2013 Elsevier Inc

Movshon, J.A., and Newsome, W.T. (1996). Visual response properties of stri-

ate cortical neurons projecting to area MT in macaque monkeys. J. Neurosci.

16, 7733–7741.

Nassi, J.J., and Callaway, E.M. (2009). Parallel processing strategies of the pri-

mate visual system. Nat. Rev. Neurosci. 10, 360–372.

Okun, M., Naim, A., and Lampl, I. (2010). The subthreshold relation between

cortical local field potential and neuronal firing unveiled by intracellular record-

ings in awake rats. J. Neurosci. 30, 4440–4448.

Otsuka, T., and Kawaguchi, Y. (2011). Cell diversity and connection specificity

between callosal projection neurons in the frontal cortex. J. Neurosci. 31,

3862–3870.

Palmer, L.M., Schulz, J.M., Murphy, S.C., Ledergerber, D., Murayama,M., and

Larkum,M.E. (2012). The cellular basis of GABA(B)-mediated interhemispheric

inhibition. Science 335, 989–993.

Petersen, C.C.H. (2007). The functional organization of the barrel cortex.

Neuron 56, 339–355.

Petersen, C.C.H., and Crochet, S. (2013). Synaptic computation and sensory

processing in neocortical layer 2/3. Neuron 78, 28–48.

Petersen, C.C.H., Hahn, T.T.G., Mehta, M., Grinvald, A., and Sakmann, B.

(2003). Interaction of sensory responses with spontaneous depolarization in

layer 2/3 barrel cortex. Proc. Natl. Acad. Sci. USA 100, 13638–13643.

Petreanu, L., Gutnisky, D.A., Huber, D., Xu, N.L., O’Connor, D.H., Tian, L.,

Looger, L., and Svoboda, K. (2012). Activity in motor-sensory projections re-

veals distributed coding in somatosensation. Nature 489, 299–303.

Poulet, J.F.A., and Petersen, C.C.H. (2008). Internal brain state regulates

membrane potential synchrony in barrel cortex of behaving mice. Nature

454, 881–885.

Poulet, J.F.A., Fernandez, L.M., Crochet, S., and Petersen, C.C.H. (2012).

Thalamic control of cortical states. Nat. Neurosci. 15, 370–372.

Rauschecker, J.P., and Tian, B. (2000). Mechanisms and streams for process-

ing of ‘‘what’’ and ‘‘where’’ in auditory cortex. Proc. Natl. Acad. Sci. USA 97,

11800–11806.

Sachidhanandam, S., Sreenivasan, V., Kyriakatos, A., Kremer, Y., and

Petersen, C.C.H. (2013). Membrane potential correlates of sensory perception

in mouse barrel cortex. Nat. Neurosci. 16, 1671–1677.

Sato, T.R., and Svoboda, K. (2010). The functional properties of barrel cortex

neurons projecting to the primary motor cortex. J. Neurosci. 30, 4256–4260.

Spreafico, R., Barbaresi, P., Weinberg, R.J., and Rustioni, A. (1987). SII-pro-

jecting neurons in the rat thalamus: a single- and double-retrograde-tracing

study. Somatosens. Res. 4, 359–375.

Tamamaki, N., and Tomioka, R. (2010). Long-range GABAergic connections

distributed throughout the neocortex and their possible function. Front.

Neurosci. 4, 202.

Tamamaki, N., Yanagawa, Y., Tomioka, R., Miyazaki, J., Obata, K., and

Kaneko, T. (2003). Green fluorescent protein expression and colocalization

with calretinin, parvalbumin, and somatostatin in the GAD67-GFP knock-in

mouse. J. Comp. Neurol. 467, 60–79.

Taniguchi, H., He, M., Wu, P., Kim, S., Paik, R., Sugino, K., Kvitsiani, D., Fu, Y.,

Lu, J., Lin, Y., et al. (2011). A resource of Cre driver lines for genetic targeting of

GABAergic neurons in cerebral cortex. Neuron 71, 995–1013.

Turner, R.S., and DeLong, M.R. (2000). Corticostriatal activity in primary motor

cortex of the macaque. J. Neurosci. 20, 7096–7108.

Welker, E., Hoogland, P.V., and Van der Loos, H. (1988). Organization of feed-

back and feedforward projections of the barrel cortex: a PHA-L study in the

mouse. Exp. Brain Res. 73, 411–435.

Xu, N.L., Harnett, M.T., Williams, S.R., Huber, D., O’Connor, D.H., Svoboda,

K., and Magee, J.C. (2012). Nonlinear dendritic integration of sensory and

motor input during an active sensing task. Nature 492, 247–251.

Zagha, E., Casale, A.E., Sachdev, R.N., McGinley, M.J., and McCormick, D.A.

(2013). Motor cortex feedback influences sensory processing by modulating

network state. Neuron 79, 567–578.

.