Properties and Role of I in the Pacing of Subthreshold ... and Role of Ih in the Pacing of...

18
Properties and Role of I h in the Pacing of Subthreshold Oscillations in Entorhinal Cortex Layer II Neurons CLAYTON T. DICKSON, 1 JACOPO MAGISTRETTI, 2 MARK H. SHALINSKY, 1 ERIK FRANSE ´ N, 3 MICHAEL E. HASSELMO, 4 AND ANGEL ALONSO 1 1 Department of Neurology and Neurosurgery, Montreal Neurological Institute and McGill University, Montreal, Quebec H3A 2B4, Canada; 2 Department of Experimental Neurophysiology, Istituto Nazionale Neurologico C. Besta, Milan 510, 20133 Italy; 3 Department of Numerical Analysis and Computing Science, Royal Institute of Technology, S-100 44 Stockholm, Sweden; and 4 Department of Psychology, Boston University, Boston, Massachusetts 02215 Dickson, Clayton T., Jacopo Magistretti, Mark H. Shalinsky, Erik Franse ´n, Michael E. Hasselmo, and Angel Alonso. Properties and role of I h in the pacing of subthreshold oscillations in entorhinal cortex layer II neurons. J. Neurophysiol. 83: 2562–2579, 2000. Var- ious subsets of brain neurons express a hyperpolarization-activated inward current (I h ) that has been shown to be instrumental in pacing oscillatory activity at both a single-cell and a network level. A characteristic feature of the stellate cells (SCs) of entorhinal cortex (EC) layer II, those neurons giving rise to the main component of the perforant path input to the hippocampal formation, is their ability to generate persistent, Na 1 -dependent rhythmic subthreshold membrane potential oscillations, which are thought to be instrumental in imple- menting theta rhythmicity in the entorhinal-hippocampal network. The SCs also display a robust time-dependent inward rectification in the hyperpolarizing direction that may contribute to the generation of these oscillations. We performed whole cell recordings of SCs in in vitro slices to investigate the specific biophysical and pharmaco- logical properties of the current underlying this inward rectification and to clarify its potential role in the genesis of the subthreshold oscil- lations. In voltage-clamp conditions, hyperpolarizing voltage steps evoked a slow, noninactivating inward current, which also deactivated slowly on depolarization. This current was identified as I h because it was resistant to extracellular Ba 21 , sensitive to Cs 1 , completely and selec- tively abolished by ZD7288, and carried by both Na 1 and K 1 ions. I h in the SCs had an activation threshold and reversal potential at approxi- mately 245 and 220 mV, respectively. Its half-activation voltage was 277 mV. Importantly, bath perfusion with ZD7288, but not Ba 21 , gradually and completely abolished the subthreshold oscillations, thus directly implicating I h in their generation. Using experimentally derived biophysical parameters for I h and the low-threshold persistent Na 1 cur- rent (I NaP ) present in the SCs, a simplified model of these neurons was constructed and their subthreshold electroresponsiveness simulated. This indicated that the interplay between I NaP and I h can sustain persistent subthreshold oscillations in SCs. I NaP and I h operate in a “push-pull” fashion where the delay in the activation/deactivation of I h gives rise to the oscillatory process. INTRODUCTION The hyperpolarization-activated inward current (I h ; usually referred to as I f in heart) has been implicated in the pacemaking of both single-cell and network rhythmicity (for recent re- views, see DiFrancesco 1993; Lu ¨thi and McCormick 1998; Pape 1996). Typically, this current acts to promote depolariza- tion after a hyperpolarizing event. This, in combination with Ca 21 currents, functions to induce low-threshold rhythmic discharge in a number of neurons and thus contributes to brain rhythm generation (Llina ´s and Jahnsen 1982; Lu ¨ thi et al. 1998; McCormick and Pape 1990; Steriade and Llina ´s 1988; Steriade et al. 1993). In contrast to Ca 21 -dependent oscillations, numer- ous studies have shown that some, mainly cortical, neuronal populations can generate Na 1 -dependent rhythmic subthresh- old membrane potential oscillations that are thought also to be implicated in the genesis of cortical rhythms (Alonso and Llinas 1989; reviewed by Connors and Amitai 1997). The role of near-threshold conductances, including I h , in the generation of these Na 1 -dependent subthreshold oscillatory events is less clear. A prominent case of Na 1 -dependent subthreshold oscilla- tory activity is observed in the principal neurons from entorhi- nal cortex (EC) layer II (Alonso and Klink 1993; Alonso and Llina ´s 1989). These glutamatergic neurons, named by Cajal as the stellate cells (SCs) (Ramon y Cajal 1902), funnel most of the neocortical input to the hippocampal formation via the perforant pathway (for review, see Dolorfo and Amaral 1998) and appear to be generators of limbic theta rhythm (Alonso and Garcı ´a-Austt 1987a,b; Buzsa ´ki 1996; Dickson et al. 1995). In vitro current-clamp studies have shown that the current-voltage relationship of EC layer II SCs is extremely nonlinear, display- ing robust inward rectification in both the depolarizing and hyperpolarizing direction. Inward rectification in the depolar- izing direction is generated by a persistent subthreshold Na 1 current (I NaP ) (Magistretti and Alonso 1999; Magistretti et al. 1999) (for a recent review on I NaP , see also Crill 1996) that has been shown to be necessary for the development of the robust theta frequency subthreshold oscillations that the SCs display (Alonso and Klink 1993; Alonso and Llina ´s 1989). On the other hand, the time-dependent inward rectification in the hyperpolarizing direction is affected by extracellular Cs 1 but not Ba 21 and thus is likely to be generated by the nonspecific cationic current I h (Klink and Alonso 1993). Given the prop- erties and role of I h in pacemaking in other excitable cells, it was proposed that this current also could contribute to the genesis of subthreshold oscillations in SCs (Alonso and Llina ´s 1989; Klink and Alonso 1993: White et al. 1995) although the exact nature of this role was not specified. Using the whole cell patch-clamp technique in the EC slice The costs of publication of this article were defrayed in part by the payment of page charges. The article must therefore be hereby marked “advertisementin accordance with 18 U.S.C. Section 1734 solely to indicate this fact. 2562 0022-3077/00 $5.00 Copyright © 2000 The American Physiological Society

Transcript of Properties and Role of I in the Pacing of Subthreshold ... and Role of Ih in the Pacing of...

Properties and Role ofIh in the Pacing of Subthreshold Oscillationsin Entorhinal Cortex Layer II Neurons

CLAYTON T. DICKSON,1 JACOPO MAGISTRETTI,2 MARK H. SHALINSKY,1 ERIK FRANSEN,3

MICHAEL E. HASSELMO,4 AND ANGEL ALONSO1

1Department of Neurology and Neurosurgery, Montreal Neurological Institute and McGill University, Montreal, QuebecH3A 2B4, Canada;2Department of Experimental Neurophysiology, Istituto Nazionale Neurologico C. Besta, Milan 510,20133 Italy;3Department of Numerical Analysis and Computing Science, Royal Institute of Technology, S-100 44 Stockholm,Sweden; and4Department of Psychology, Boston University, Boston, Massachusetts 02215

Dickson, Clayton T., Jacopo Magistretti, Mark H. Shalinsky, ErikFransen, Michael E. Hasselmo, and Angel Alonso.Properties androle of Ih in the pacing of subthreshold oscillations in entorhinalcortex layer II neurons.J. Neurophysiol.83: 2562–2579, 2000. Var-ious subsets of brain neurons express a hyperpolarization-activatedinward current (Ih) that has been shown to be instrumental in pacingoscillatory activity at both a single-cell and a network level. Acharacteristic feature of the stellate cells (SCs) of entorhinal cortex(EC) layer II, those neurons giving rise to the main component of theperforant path input to the hippocampal formation, is their ability togenerate persistent, Na1-dependent rhythmic subthreshold membranepotential oscillations, which are thought to be instrumental in imple-menting theta rhythmicity in the entorhinal-hippocampal network.The SCs also display a robust time-dependent inward rectification inthe hyperpolarizing direction that may contribute to the generation ofthese oscillations. We performed whole cell recordings of SCs inin vitro slices to investigate the specific biophysical and pharmaco-logical properties of the current underlying this inward rectificationand to clarify its potential role in the genesis of the subthreshold oscil-lations. In voltage-clamp conditions, hyperpolarizing voltage stepsevoked a slow, noninactivating inward current, which also deactivatedslowly on depolarization. This current was identified asIh because it wasresistant to extracellular Ba21, sensitive to Cs1, completely and selec-tively abolished by ZD7288, and carried by both Na1 and K1 ions.Ih inthe SCs had an activation threshold and reversal potential at approxi-mately245 and220 mV, respectively. Its half-activation voltage was277 mV. Importantly, bath perfusion with ZD7288, but not Ba21,gradually and completely abolished the subthreshold oscillations, thusdirectly implicatingIh in their generation. Using experimentally derivedbiophysical parameters forIh and the low-threshold persistent Na1 cur-rent (INaP) present in the SCs, a simplified model of these neurons wasconstructed and their subthreshold electroresponsiveness simulated. Thisindicated that the interplay betweenINaP and Ih can sustain persistentsubthreshold oscillations in SCs.INaP and Ih operate in a “push-pull”fashion where the delay in the activation/deactivation ofIh gives rise tothe oscillatory process.

I N T R O D U C T I O N

The hyperpolarization-activated inward current (Ih; usuallyreferred to asIf in heart) has been implicated in the pacemakingof both single-cell and network rhythmicity (for recent re-views, see DiFrancesco 1993; Lu¨thi and McCormick 1998;

Pape 1996). Typically, this current acts to promote depolariza-tion after a hyperpolarizing event. This, in combination withCa21 currents, functions to induce low-threshold rhythmicdischarge in a number of neurons and thus contributes to brainrhythm generation (Llina´s and Jahnsen 1982; Lu¨thi et al. 1998;McCormick and Pape 1990; Steriade and Llina´s 1988; Steriadeet al. 1993). In contrast to Ca21-dependent oscillations, numer-ous studies have shown that some, mainly cortical, neuronalpopulations can generate Na1-dependent rhythmic subthresh-old membrane potential oscillations that are thought also to beimplicated in the genesis of cortical rhythms (Alonso andLlinas 1989; reviewed by Connors and Amitai 1997). The roleof near-threshold conductances, includingIh, in the generationof these Na1-dependent subthreshold oscillatory events is lessclear.

A prominent case of Na1-dependent subthreshold oscilla-tory activity is observed in the principal neurons from entorhi-nal cortex (EC) layer II (Alonso and Klink 1993; Alonso andLlinas 1989). These glutamatergic neurons, named by Cajal asthe stellate cells (SCs) (Ramon y Cajal 1902), funnel most ofthe neocortical input to the hippocampal formation via theperforant pathway (for review, see Dolorfo and Amaral 1998)and appear to be generators of limbic theta rhythm (Alonso andGarcıa-Austt 1987a,b; Buzsa´ki 1996; Dickson et al. 1995). Invitro current-clamp studies have shown that the current-voltagerelationship of EC layer II SCs is extremely nonlinear, display-ing robust inward rectification in both the depolarizing andhyperpolarizing direction. Inward rectification in the depolar-izing direction is generated by a persistent subthreshold Na1

current (INaP) (Magistretti and Alonso 1999; Magistretti et al.1999) (for a recent review onINaP, see also Crill 1996) that hasbeen shown to be necessary for the development of the robusttheta frequency subthreshold oscillations that the SCs display(Alonso and Klink 1993; Alonso and Llina´s 1989). On theother hand, the time-dependent inward rectification in thehyperpolarizing direction is affected by extracellular Cs1 butnot Ba21 and thus is likely to be generated by the nonspecificcationic currentIh (Klink and Alonso 1993). Given the prop-erties and role ofIh in pacemaking in other excitable cells, itwas proposed that this current also could contribute to thegenesis of subthreshold oscillations in SCs (Alonso and Llina´s1989; Klink and Alonso 1993: White et al. 1995) although theexact nature of this role was not specified.

Using the whole cell patch-clamp technique in the EC slice

The costs of publication of this article were defrayed in part by the paymentof page charges. The article must therefore be hereby marked “advertisement”in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.

2562 0022-3077/00 $5.00 Copyright © 2000 The American Physiological Society

preparation, the aim of the present study was to characterizethe specific properties ofIh in the SCs and to examine the roleof this current in the generation of subthreshold membranepotential oscillations in these cells. In addition, a simplifiedbiophysical simulation based on the voltage- and current-clampdata was used to study the interactions betweenIh and INaP inthe generation of such oscillations. Our results indicate that thedynamic interplay between the gating and kinetic properties ofIh and INaP is essential for the generation of rhythmic sub-threshold oscillations by the SCs. Given the key position of theSCs in the temporal lobe memory system, modulation ofIh inthe SCs may have major implications for the control of popu-lation dynamics in the entorhinal network and in the memoryprocesses it carries out. Some of these results have beenpresented previously in abstract form (Dickson and Alonso1996, 1998; Franse´n et al. 1998).

M E T H O D S

General

Brain slices were prepared from male Long-Evans rats (100–250 g,i.e., 30–60 days of age) as previously described (Alonso and Klink1993). Briefly, animals were decapitated quickly, and the brain wasremoved rapidly from the cranium, blocked, and placed in a cold(4°C) Ringer solution (pH 7.4 by saturation with 95% O2-5% CO2)containing (in mM) 124 NaCl, 5 KCl, 1.25 NaH2PO4, 2 CaCl2, 2MgSO4, 26 NaHCO3, and 10 glucose. Horizontal slices of the retro-hippocampal region were cut at 350–400mm on a vibratome (Pelco

Series 1000, Redding, CA) and were transferred to an incubationchamber in which they were kept submerged for$1 h at roomtemperature (24°C). Slices were transferred, one at a time, to arecording chamber and were superfused with Ringer solution, also atroom temperature. The chamber was located on the stage of anupright, fixed-stage microscope (Axioskop, Zeiss) equipped with awater immersion objective (340–63: long-working distance), No-marski optics, and a near-infrared charge-coupled device (CCD) cam-era (Sony XC-75). With this equipment, stellate and pyramidal-likecells could be distinguished based on their shape, size, and positionwithin layer II of the medial entorhinal cortex (Fig. 1A) (Klink andAlonso 1997). Stellate cells (SCs) were selected for whole cell re-cording.

Recording

Patch pipettes (4–7 MV) were filled with (in mM) 140–130 gluconicacid (potassium salt: K-gluconate), 5 NaCl, 2 MgCl2, 10 N-2-hydroxy-ethylpiperazine-N-2-ethanesulfonic acid (HEPES), 0.5 ethylene glycol-bis(b-aminoethyl ether)-N,N,N9,N9-tetraacetic acid (EGTA), 2 ATP(ATP Tris salt), and 0.4 GTP (GTP Tris salt), pH 7.25 with KOH. Inadditional experiments performed to assess the contribution of chlorideions toIh, a modified intracellular solution was made containing (in mM)120 K-gluconate, 10 KCl, 5 NaCl, 2 MgCl2, 10 HEPES, 0.5 EGTA, 2ATP-Tris, and 0.4 GTP-Tris, pH 7.25 with KOH. The liquid junctionpotential was estimated following the technique of Neher (1992). In brief,the offset was zeroed while recording the potential across the patchpipette and a commercial salt-bridge ground electrode (MERE 2, WPI,Sarasota, FL) when the chamber was filled with the same intracellularsolution as used in the pipette. After zeroing, the chamber solution was

FIG. 1. Basic electrophysiological profile of entorhinal cortex (EC) layer II stellate cells (SCs) under whole cell current-clamprecording conditions.A: digitized photomicrograph demonstrating the visualization of a patched SC. - - -, approximate borderbetween layers I and II.B: V-I relationship of the SC inA demonstrating robust time-dependent inward rectification in thehyperpolarizing direction.C: action potential from (D2) (*) at an expanded time and voltage scale. Note the fast-afterhyperpo-larization/depolarizing afterpotential/medium afterhyperpolarization (fast-AHP/DAP/medium-AHP) sequence characteristic ofthese cells.D: subthreshold membrane potential oscillations (1 and2) and spike clustering (3) develop at increasingly depolarizedmembrane potential levels positive to about255 mV. Autocorrelation function (inset in 1) demonstrates the rhythmicity of thesubthreshold oscillations.

2563Ih IN EC LAYER II NEURONS

replaced with the extracellular recording solution, and the potential re-corded was used as an estimate of the liquid junction potential. Using thismethod, we recorded a value between 2 and 3 mV. Membrane potentialvalues reported herein do not contain this correction.

Tight seals (.1 GV) were formed on cell bodies of selected EClayer II SCs, and whole cell recordings were made by rupturing thecell membrane with negative pressure. Both current- and voltage-clamp recordings were made with an Axopatch 1D patch-clampamplifier (Axon Instruments, Foster City, CA). For current-clamprecordings, the low-pass filter (23dB) was set at 10 kHz, whereas forvoltage clamp, it was set at 2 kHz. All current- and some voltage-clamp experiments were stored by PCL coding on VHS tape (Neu-rocorder, Neurodata, New York), and all voltage-clamp experimentswere stored on computer by digital sampling at 4 kHz, using pClampsoftware (V6.0, Axon Instruments). Data stored on VHS tape wasdigitized and plotted off-line by sampling at 20 kHz using Axoscopesoftware (V1.1, Axon Instruments).

The identification of neurons as SCs was confirmed by current-clamp recordings demonstrating the presence of robust inward recti-fication with hyperpolarizing current pulses in addition to the presenceof subthreshold membrane potential oscillations at depolarized levels(Fig. 1,B andC) (Alonso and Klink 1993). SCs fulfilling the follow-ing criteria were considered acceptable for further analysis: stablemembrane potential less than250 mV, input resistance.75 MV,overshooting spike, and a balanced series resistance,20 MV com-pensated between 60 and 80%.

Solutions

Various salts and drugs were added directly to the perfusate fromconcentrated stock solutions during experiments. Divalent cations such asBa21, Co21, or Cd21 were added to a modified Ringer solution withoutphosphates or sulfates. To isolateIh to compute activation curves usingthe tail current method (seeRESULTS), the following solution was used (inmM): 80 NaCl, 40 tetraethylammonium chloride (TEA-Cl), 5 KCl, 44-aminopyridine (4-AP), 2 MgCl2, 2 BaCl2, 2 CoCl2, 1 CaCl2, 0.2 CdCl2,26 NaH2CO3, and 10 glucose and 1mM tetrodotoxin (TTX). Loweringthe extracellular concentration of sodium ions (NaCl) was achieved usingequimolarN-methyl-D-glucamine substitution in a no phosphate/sulfateRinger solution. Alterations in the concentration of potassium ions (KCl)was achieved in the same way in a no phosphate/sulfate Ringer solutionwith a consistent concentration (119 mM) of NaCl. To prevent theinfluence of synaptic transmission on the subthreshold membrane behav-ior in current-clamp recordings, 6-cyano-7-nitoquinoxaline-2,3-dione(CNQX: 10 mM), DL22-amino-5-phosphonopentanoic acid (AP-5: 50mM), bicuculline methiodide (BMI: 10mM), and 2-hydroxysaclofen(2-OH saclofen:100mM) were added to the Ringer solution. To blockIh,CsCl (1–6 mM) or ZD7288 (100mM) was added directly to the Ringersolution. All salts were purchased from BDH (Toronto, CA), whereasTEA-Cl, 4-AP, TTX, and BMI were purchased from Sigma (St. Louis,MO). CNQX, AP-5, 2-OH-saclofen, and ZD7288 were purchased fromTocris Cookson (UK).

Analysis

Traces were plotted and measurements made with the use ofpClamp (Clampfit) and Origin (Microcal, Northampton, MA) soft-ware packages. Curve fitting of subtracted traces was conducted withpClamp software. The fittings were made from a time point 15 msafter the application of the hyperpolarizing voltage step so as tominimize any capacitive or membrane charging transients. All curve-fitting procedures were optimized using the least sum of squaresmethod. The standard deviation between the fit and the data were usedto estimate the goodness of the fit. Autocorrelational analysis wasconducted with Matlab (Mathworks, Natick, MA). Spectral (Fourier)analysis was conducted using both Origin and Matlab.

Biophysical simulation

A simple Hodgkin-Huxley model was assumed for describingIhactivation and deactivation. We applied the basic relationship

I h~V,t! 5 Gh~V,t! z ~V 2 Vh! (1)

where

Gh~V,t! 5 GhMax z m~V,t! (2)

andm is the probability of the activating particle to be in the permis-sive position. BecauseIh activation and deactivation could be properlyfitted with double-exponential functions,Eq. 2was applied separatelyto each exponential component, whereasEq. 1was extended to

I h~V,t! 5 @Gh1~V,t! 1 Gh2~V,t!# z ~V 2 Vh!

Conductance values were derived from amplitude coefficient (Ai)values by applying the relationship:Ghi(V,t) 5 Ai(V,t)/(V 2 Vh),where the index i is either 1 or 2. The functions describingm`i(V)were derived directly fromGhi(V) activation curves. The transitions ofthe activating particle,mi, were schematized as the following first-order kinetic reaction:

1 2 mi ºbi

ai

mi

from which it follows

mi~V,t! 5 m`i~V! 2 @m`i~V! 2 m0i# z exp~2t/t i! (3)

where

t i~V! 5 1/@a i~V! 1 b i~V!# (4)

m`i~V! 5 a i~V!/@a i~V! 1 b i~V!# (5)

Numerical values for the rate constants,ai andbi, were derived fromthe experimental values of time constants of activation and deactiva-tion (ti) and from them`i curves by applyingEqs. 4and 5. Rate-constant plots were best fitted with the empiric function

a i~or b i! 5 ~ai z Vm 1 bi!/$1 2 exp@~Vm 1 bi /aI!/ki#%

After obtaining the analytic functions appropriately describing thevoltage dependence of rate constants, the changes inIh during varioussimulated current-clamp protocols were numerically reconstructed onthe basis of theEqs. 1and2 andEq. 3 in its differential form.

The basic equations used for describingINaPwere the same as usedfor Ih (see preceding text,Eqs. 1 and 2). GNaP(V) was modeledaccording to the voltage-dependence data reported by Magistretti andAlonso (1999).INaP activation was assumed to be instantaneous.

Kinetic and voltage-dependence parameters concerningIh andINaP

were used in a simplified model of an EC SC aimed at reproducing thesubthreshold oscillatory behavior of membrane potential in thesesame neurons. In this model, the neuron was considered as mono-compartmental, and its membrane conductance consisted ofGh, GNaP,and a linear leakage conductance (Gl) whose current reversed at theequilibrium potential for K1. The parameters of the equations de-scribing conductance kinetics and voltage dependence were given thesame numerical values as returned by the analysis of the relevantexperimental data (seeRESULTS). Na1 and K1 reversal potentials hadthe theoretical (Nernst) values calculated for the ionic conditionsemployed in our current-clamp experiments (VNa 5 187 mV, VK 5283 mV). The reversal potential forIh (Vh 5 220 mV) and theamplitude ratio between the fast and slow kinetic components ofGh

(Gh1Max /Gh2Max 5 1.85) also matched exactly the experimentallyobserved average values. Only the absolute values of maximal con-ductances (GMax) were adjusted until a good concordance betweensimulations and experimental observations was achieved. In the sim-ulations here illustrated,GhMax, GNaPMax, and GlMax equaled 98.0,

2564 DICKSON, MAGISTRETTI, SHALINSKY, FRANSE´ N, HASSELMO, AND ALONSO

17.4, and 78.0 pS/pF, respectively. These values compared reason-ably, namely within a factor of 2, with the experimentally measuredvalues.

Numeric solution of the differential equations was achieved by the useof a one-step Euler integration method. The integration step size was 0.25ms. Preliminary tests on the adequacy of this integration method werecarried out by reducing the step size by#25 times, which revealed anoptimal convergence. The simulation programs were compiled usingQuickBASIC 4.5 (Microsoft). Data were analyzed using Origin.

R E S U L T S

The results presented in this study were based on a databaseof 131 EC layer II SCs intracellularly recorded under wholecell patch conditions and met the criteria specified inMETHODS.The studied neurons were identified as SCs by their grossmorphological characteristics (Klink and Alonso 1993) as af-forded by direct visualization of their somata and proximaldendrites (Fig. 1A) but mainly by their characteristic electro-physiological properties (Alonso and Klink 1993; Alonso andLlinas 1989) (Fig. 1,B–D). Indeed, as illustrated in Fig. 1,patched SCs demonstrated qualitatively the same electrore-sponsive properties that distinguish SCs recorded with sharpelectrodes. First, the patched SCs demonstrated robust time-dependent inward rectification in the hyperpolarizing direction.As shown in Fig. 1B, the membrane voltage responses tohyperpolarizing current pulses did not monotonically reach asteady value but displayed, after a certain delay, large ampli-tude “sags” back to more depolarized values. Second, theaction potential of the patched SCs also demonstrated the

characteristic fast after hyperpolarization (arrowhead Fig. 1C)followed by a depolarizing afterpotential and a medium afterhyperpolarization. Finally, and most importantly, patched SCsalso developed rhythmic subthreshold membrane potential os-cillations and demonstrated cluster discharge when depolarizedwith DC current in the membrane potential range between255and250 mV (Fig. 1D, 1–3). At an average membrane poten-tial of 252 6 1 mV, the peak frequency of these membranepotential oscillations as determined by Fourier analysis aver-aged 3.16 0.7 Hz (n 5 12). The SCs had an average restingmembrane potential of2556 3 mV and an input resistance of113 6 40 MV.

Although not further treated, in some instances, neuronsother than SCs were recorded from. Pyramidal-like cells of EClayer II (n 5 10), could be distinguished from SCs based ontheir pyramidal shape and their limited expression of time-dependent inward rectification (Klink and Alonso 1993). LayerIII pyramidal cells (n 5 3) were distinguishable based on theirqualitatively smaller size, their high-input resistance (2176 67MV) and the absence of time-dependent inward rectification(Dickson et al. 1997).

Hyperpolarization-activated, time-dependent inwardrectification in SCs corresponded to a slow, noninactivatinginward current

As illustrated for a typical SC in Fig. 2, the depolarizing sagsthat developed on membrane hyperpolarization in current-clamp conditions (A, 4) were paralleled by the development

FIG. 2. Characterization of inward rec-tification in the SCs.A: in current-clampconditions, hyperpolarizing current pulsesevoke voltage responses with a robusttime-dependent depolarizing sag (4). B:in the same neuron, under voltage-clampconditions, hyperpolarizing voltage stepsevoke a slow inward current (Ih) thatgrows in amplitude and rate of activationwith increasing hyperpolarization.C:peak and steady-stateV-I plot (E and ●,respectively) derived from the data inA.Time-dependent inward rectification isgraphically represented as the depolariz-ing shift between peak and steady-statepotential readings.D: instantaneous andsteady-stateI-V plot (E and ●, respec-tively) derived from the data inB. Time-dependent inward rectification is apparentas the negative difference in the holdingcurrent between the instantaneous andsteady-state measurements. All tracesshown were obtained in the standardRinger solution.

2565Ih IN EC LAYER II NEURONS

of a slow inward current on step hyperpolarization undervoltage-clamp conditions (B,3). Note that the time course andamplitude of this inward relaxation was overtly voltage depen-dent (see following text). In all cases, analysis of the subthresh-old input-output relations under current-clamp revealed that thesteady-state voltage-current (V-I) curve (Fig. 2C; F) showed amarked upward bending over the entire voltage range (260 to2120). Similarly, analysis of input-output relations under volt-age-clamp revealed that the steady-state current-voltage curve(ssI-V; Fig. 2D, F) showed a robust inward shift, as comparedwith the instantaneous current-voltage curve (Fig. 2D, E), thatgrew steadily with membrane hyperpolarization. The slowinward current relaxations were associated with a membrane-conductance increase because the instantaneous current flow-ing at the break of the hyperpolarizing commands was largerthan that recorded on first jumping to the command potential(see Fig. 4A). Thus SCs do possess a robust time-dependenthyperpolarization activated conductance (Gh).

Pharmacological block of inward rectification

In addition to a time-dependent inward rectifier such asIh,many neurons also possess a fast inward rectifier K1 current(IKir) (reviewed by Hille 1992). It has been shown that in manycells bath application of Ba21 and Cs1 can be used to phar-macologically dissectIh from IKir because Ba21 blocksIKir andnot Ih, whereas Cs1 blocks bothIKir and Ih (Hagiwara et al.1976, 1978). In agreement with this, in all SCs tested (n 5 8),bath application of Ba21 (0.5–2 mM) had no effect onIh (Fig.3, A–C), although it did block the small inward bending of the

instantaneousI-V relationship that was always observed atpotentials negative to about280 mV in control conditions(Fig. 3C, squares). This Ba21 effect suggests the presence of aminor IKir in the SCs. In contrast to Ba21, in all SCs tested (n 510), bath application of Cs1 (1–6 mM) always produced asubstantial decrease inIh (though never a completeIh block).This decrease was assessed by expressing the percentage de-crease in the difference between the instantaneous and steady-state current at potentials between260 and280 mV beforeand after application of Cs1 (cf. Ishii et al. 1999). It was dosedependent and ranged from 60 to 75% for a concentration of 2mM Cs1 (n 5 5) that produced close to maximal effects.

Given that Cs1 produced only a partial block ofIh in theSCs, we assessed the effects of the novel bradycardic agentZD7288, which has been reported to be a potent blocker ofIh

in other cells (BoSmith et al. 1993; Harris and Constanti 1995;Maccaferri and McBain 1996; Williams et al. 1997). As illus-trated in Fig. 3 (D and E), in all cases tested (n 5 9), bathapplication of ZD7288 (.10 min; 100mM) always resulted ina complete and irreversible block ofIh. When the cells wereheld at260 mV (about resting level), application of ZD7288always resulted in an outward shift of the holding current(mean5138 6 52 pA, n 5 5), indicating thatGh is active atthe resting membrane potential (see following text). Signifi-cantly, ZD7288 did not abolish the small inward shift of theinstantaneousI-V relation below280 mV (Fig. 3G) and thuswhereas ZD7288 fully blockedIh, it did not affect IKir. Al-though small, the remaining fast inward rectification, however,could be blocked fully by the further addition of Ba21 that also

FIG. 3. Effects of extracellular Ba21 andZD7288 on Ih. A and B: current responsesevoked by a series of voltage-clamp steps incontrol conditions (A) and during superfusionwith 2 mM Ba21 (B). C: I-V plot of both theinstantaneous (n and▫) and steady-state (E and●) current responses fromA andB. Note thatBa21 did not block time-dependent inwardrectification (generated byIh) though it didappear to affect fast inward rectification (ap-parent as the negative bending of the instanta-neous current plot at potentials negative to280 mV).D–F: current responses evoked by aseries of voltage-clamp steps in control condi-tions (D), in the presence of 100mM ZD7288(E) and during further addition of 2 mM Ba21

(F). G: I-V plot of both the instantaneous (n

and ▫) and steady-state (E and ●) current re-sponses from (D, control) and (E, 1ZD7288).H: I-V plot of both the instantaneous (▫ andv)and steady-state (E and R) current responsesfrom (E, 1ZD7288) and (F, 1Ba21). Notethat ZD7288 completely abolished time-de-pendent inward rectification and that Ba21

abolished the remaining, minor, fast inwardrectification. All recordings were carried out inthe presence of 1mM TTX and 2 mM Co21.

2566 DICKSON, MAGISTRETTI, SHALINSKY, FRANSE´ N, HASSELMO, AND ALONSO

caused a decrease in slope conductance due to its blockingaction on leak currents (n 5 3; Fig. 3,F andH).

Activation of Ih

We estimated the activation curve of the membrane con-ductance underlyingIh (Gh) by applying two different pro-tocols (Fig. 4). In the first protocol, a modified Ringersolution (as specified inMETHODS) was used. The activationcurve of Gh was estimated from the peak amplitude of thetail currents recorded at about240 mV (n 5 8) or at about260 mV (n 5 5) after a series of hyperpolarizing voltage-clamp steps from a holding potential in the range of245 to230 mV (Fig. 4,A–C). When stepping back to260 mV, thezero current level was the tail current amplitude after themost depolarized voltage step (at least240 mV). Tailcurrent amplitudes were normalized to the maximal value(Imax) and plotted as a function of the membrane potentialduring the hyperpolarizing prepulse. In all cases (n 5 13),the data were well fitted with a Boltzmann equation of theform

I /Imax 5 ~1 1 e~~Vm2V1/2!/k!!21

whereVm is the membrane potential of the prepulse,V1/2 themembrane potential at whichGh is half activated,k a slopefactor, andI is the amplitude of the tail current recorded afterthe prepulse. The tail current analysis yielded an activationrange ofGh between245 and2115 mV, a mean value forV1/2

of 277 6 5 mV and a slope factor (k) of 11.26 1.8 (Fig. 4C).For comparison, in four cells we applied slow (,10

mV/s) 100 mV hyperpolarizing voltage ramps from a hold-ing potential of230 mV in control and after block ofIh withZD7288 (Fig. 4D). In these experiments, 1mM TTX, 2 mMCo21, and 2 mM Ba21 were added to the control Ringer.Subtraction of the ZD7288I-V curve from the control curveyielded the steady-stateIh I-V relationship from which weestimatedGh according to the formulaGh5 Ih/(Vm 2 Vh)whereVh is the reversal potential forIh estimated to be221mV (see following text, Fig. 6). The resulting values werenormalized to the maximal conductance (Gmax; 10.4 6 2.3nS) and plotted againstVm. In all cases the curves werewell fitted with a Boltzmann equation as in the precedingtext. This ramp analysis yielded aV1/2 of 276 6 4 mV anda slope factor of 12.16 2, which were not significantlydifferent to those obtained by tail current analysis by two-tailed t-tests [t(15) 5 0.36, P . 0.05; t(15) 5 20.85,P . 0.05].

FIG. 4. Activation curve for the conductanceGh. A: incremental hyperpolarizing voltage-clamp steps increase the amount ofIh

that is activated as well as the amplitude of the tail currents that follow (inset). B: tail currents fromA shown at an expanded timeand current scale. Tail current amplitude measurements were taken at the time indicated by the empty circle and broken line.C:plot of the activation curve ofGh. Filled circles, averaged activation curve for 13 SCs. Individual experiments first were fitted withBoltzmann functions and the interpolated values for the steps at each 10-mV increment were used to compute the average value.Small open circles, data derived from cell inA. Line, best Boltzmann fit to the average (seeRESULTS) where the half activationvoltage and slope factor were found to be277 6 5 mV and 11.26 1.8, respectively.D: Gh activation curves also were estimatedin four cells fromIh-V relations whereIh was isolated by subtracting ZD7288I-V curves from controlI-V curves obtained throughslow ramp protocols (inset,seeRESULTSfor details). As can be seen, theGh activation curve obtained in this manner (noisy line)and its Boltzmann fit (dotted line) largely overlap the activation curve obtained from tail current analysis (continuous line; sameas in C). Tail current protocols were conducted using the specialized solution described inMETHODS, and ramp protocols wereconducted in solution containing 1mM TTX, 2 mM Co21, and 2 mM Ba21.

2567Ih IN EC LAYER II NEURONS

Time course of activation and deactivation

As stated previously, the rate of activation ofIh increasedsharply with hyperpolarization (e.g., Fig. 2B). This qualitativeobservation was further explored in a more quantitative man-ner. To maximize the accuracy of our kinetic analysis, weisolatedIh by subtracting from control current traces evoked byhyperpolarizing voltage-clamp steps, the current traces evokedto the same potentials in the presence of the selectiveIh blockerZD7288 (n 5 5; Fig. 5A). Over the whole voltage range tested,the Ih current relaxations were best fitted with a double expo-nential function of the form

I h~t! 5 A1e~2t/t1! 1 A2e

~2t/t2! 1 C

where Ih(t) is the amplitude of the current at timet, C is aconstant, andA1, andA2 reflect the amplitude coefficients ofthe fast (t1) and slow (t2) time constants, respectively. At-tempts to fit the subtracted traces with a single exponentialfunction were judged to be unsuccessful based on visual in-spection and by comparison of the standard deviations of thefits using single or double exponential functions (not shown).Both the first and second time constants were found to bevoltage dependent, as shown in Fig. 5,C and D, becomingfaster with increased hyperpolarization. The first time constantranged between 786 12 and 396 6 ms for voltage steps to270 and2110 mV, respectively. The second time constantranged between 3726 39 and 1646 45 ms for the samevoltage steps. The ratio of the amplitude coefficients for thefirst (A1) and second (A2) time constants increased from 1 at270 mV to just over 2 at2110 mV.

An equivalent method as the one described in the precedingtext was conducted to study the rate of deactivation ofIhisolated with the use of ZD7288 (Fig. 5B). IsolatedIh currenttraces evoked by depolarizing voltage steps from a holdingpotential of260 mV were also well fitted by a double-expo-nential function. As for the time constants of activation, both

the first and second time constants of deactivation were foundto be voltage dependent, becoming faster with increasing de-polarization (Fig. 5,C and D). The first time constant ofdeactivation ranged from 236 9 to 58 6 13 ms for voltagesteps to240 and 250 mV, respectively. The second timeconstant of deactivation ranged between 2416 38 and 326658 ms for the same voltage steps. The amplitude of the fasttime constant was roughly 1.25 that of the slower, and this ratioremained constant over the voltage range tested.

Reversal of Ih

Estimation of the reversal potential ofIh was achieved bytwo different methods, which took advantage of the fact that at280 mV, Gh was strongly activated and did not show time-dependent inactivation (Fig. 6,A andB). In all experiments, thesuperfusing Ringer solution contained 1mM TTX, 2 mMCoCl2, and 2 mM BaCl2. Thus in the first method, we esti-mated the reversal potential ofIh (Vh) from the intersection ofthe instantaneous (chord) current-voltage relationships re-corded at holding potentials of280 and240 mV (i.e., in thepresence and absence ofIh; Fig. 6C) (Mayer and Westbrook1983). In 17 neurons examined, this method provided an av-erage value forVh of 221 6 5mV.

To support the preceding estimation, we used a secondmethod in which we took advantage of the fact thatIh isselectively and fully blocked by ZD7288 (see preceding text).Chord conductance measurements were made from voltagesteps from a holding potential of280 mV before and afterblock of Ih using ZD7288 and the instantaneous I/V relation-ships in both conditions were constructed (Fig. 6,D–F). Ineight neurons examined, the average voltage at which thelinear fits for both plots intersected, i.e., the reversal potentialfor Ih, was 222 6 6 mV, a value that was not significantlydifferent from that found with the tail current analysis methodabove [t(23) 5 0.44,P . 0.05].

FIG. 5. Activation and deactivation kineticsof Ih. Ih was isolated by subtraction of ZD7288current traces from control current traces re-corded in the presence of 1mM TTX, 2 mMCo21, and 2 mM Ba21. A: averaged (n 5 5) Ih

traces (z z z ) evoked by hyperpolarizing stepsfrom 260 mV. Ih activation time course waswell fitted by a double-exponential decay func-tion (—, seeRESULTS for details).B: averaged(n 5 5) Ih traces (z z z ) evoked by depolarizingsteps from260 mV (same collection of cells asin A). Ih deactivation time course was also wellfitted by an incremental asymptotic double-ex-ponential function (—, seeRESULTS for details).C andD: plots of the 1st and 2nd time constantsof activation and deactivation as a function ofmembrane potential. Kinetics of activation ofIh

(n) can be seen to be voltage dependent with boththe 1st (C) and 2nd (D) time constants decreas-ing with hyperpolarization. Kinetics of deactiva-tion of Ih (▫) also can be seen to be voltagedependent with both the 1st (C) and 2nd (D) timeconstants decreasing with depolarization.

2568 DICKSON, MAGISTRETTI, SHALINSKY, FRANSE´ N, HASSELMO, AND ALONSO

Ionic basis of Ih

The fact that in the SCsIh reverses at about220 mVsuggests that, as in other neurons (Crepel and Penit-Soria 1986;Halliwell and Adams 1982; Mayer and Westbrook 1983; Mc-Cormick and Pape 1990; Spain et al. 1987; Takahashi 1990),this hyperpolarization-activated inward current might be car-ried by a mixture of both Na1 and K1 ions. Indeed, increasingthe extracellular concentration of K1 ([K1]o) (Fig. 7) producedan increase inIh (with no change in theGh activation curve; notshown) as well as an increase in instantaneous conductance. Asexpected for K1 being an important carrier forIh, an increasein [K1]o from 1 to 10 mM produced an average positive shiftin Vh of 10 6 4mV (n 5 4, Fig. 7D).

On the other hand, reductions in the concentration of extracel-lular Na1 from control levels (151 mM) to 26 mM reversiblyreduced the amplitude ofIh (Fig. 8) without changing the activa-tion properties of the conductance underlying this current (notshown). Concomitant with this reduction,Vh shifted in the hyper-polarizing direction by an average of 216 5mV (n 5 5). Theseresults indicate that Na1 ions also largely contribute toIh.

Finally, a number of neurons (5) were recorded using amodified intracellular solution containing an additional 10 mMCl2 in the pipette solution (seeMETHODS). Although in thesecases, the chloride reversal potential was theoretically shiftedby ;20 mV in a positive direction, no significant differencewas observed in either the average reversal potential [223 66 mV; t(20)5 21.48,P . 0.05] or the activation properties ofIh (not shown). Thus using the Goldman-Hodgkin-Katz equa-tion and an estimatedVh of 221.5 mV, we calculated apermeability (conductance) ratio for Na1 and K1 (pNa1/pK1)of ;0.4 for Ih in the SCs.

Role of Ih in membrane potential oscillations

Given the overlap between the activation range ofGh(threshold at about245 mV), and the voltage range at whichsubthreshold membrane potential oscillations occur in SCs(260 to 250 mV), we sought to define the involvement ofIhin these oscillations by exploring the effects on them of Cs1,ZD7288 and Ba21. Because these agents, particularly Cs1 andBa21, greatly enhance spontaneous synaptic events, we carriedout this analysis during synaptic transmission block withCNQX (10 mM), AP5 (50 mM), bicuculline (10 mM), and2-OH-saclofen (100mM). In line with a role of Ih in thegeneration of the rhythmic subthreshold oscillations, we ob-served that the addition of Cs1 (1–2 mM; n 5 4) to thesuperfusate resulted in a progressive disruption (and slow-down) of the oscillations. However, as previously reported(Klink and Alonso 1993), some trains of subthreshold oscilla-tory activity could consistently be observed in the presence ofCs1. This result might be interpreted as suggestive that, inaddition toIh, another conductance operating in the subthresh-old range, such as the M current, may play a major role in thegeneration of the rhythmic subthreshold oscillations by theSCs. Alternatively, it also might be that the expression ofsubthreshold oscillatory activity by the SCs is rather insensitiveto the level ofIh expression and that a major decrease inIh isnecessary to abolish the oscillations. To explore these possi-bilities, we first tested the effects of the more potentIh blockerZD7288. In all cells, application of 50–100mM ZD7288always resulted in membrane hyperpolarization (96 4 mV;n 5 8) concomitant with the block of the typical depolarizingvoltage sag evoked by hyperpolarizing current pulses. Simi-larly to Cs1, ZD7288 always produced a progressive disrup-tion of the oscillations, though, in contrast to what was ob-

FIG. 6. Reversal potential ofIh. Extrapolated reversal potential ofIh (Vh) was estimated by 2 different approaches.A–C: SCswere held at240 and280 mV (A andB, respectively) and the instantaneousI-V relations constructed.Vh was estimated from theintersection of the extrapolatedI-V relations derived from both holding potentials (C). D andE: SCs were held at280 mV, andthe instantaneousI-V relation was constructed in control conditions (D) and duringIh block with ZD7288 (E). Vh was estimatedfrom the intersection of the extrapolatedI-V relations derived in both conditions (F). All recordings were conducted in a solutioncontaining 1mM TTX, 2 mM Co21, and 2 mM Ba21.

2569Ih IN EC LAYER II NEURONS

served with Cs1, this disruption always proceeded to acomplete block (Fig. 9,A–C). Although these data suggest that,indeed, a major block ofIh is necessary to completely abolishthe oscillations, it could be argued that the blocking effect ofZD7288 might have been due to a nonselective action of thedrug on another conductance operating in the oscillatory range.To exclude this possibility, we performed a series of voltage-clamp experiments in which we examined the effects ofZD7288 on the outward current relaxations evoked by a seriesof depolarizing voltage-clamp steps from260 mV (aboutresting level) to the voltage range where subthreshold oscilla-tions develop (255 to 250 mV; Fig. 10A) and up to theGh

activation threshold (245 mV). These experiments were con-ducted in the presence of 1mM TTX and 2 mM Co21. Asshown in Fig. 10,B–D, ZD7288 (100mM) always caused arobust outward shift in the holding current and a complete andselective block of both the outward current relaxations inresponse to membrane depolarization as well as the associatedtail currents on return to the holding potential (n 5 4). Incontrast, there was a nearly perfect overlap between the tracesat 245 mV, the threshold for activation ofIh, before and afterZD7288. This indicates that, in the voltage range from260 to245 (which includes the voltage range at which the membranepotential oscillations occur) the action of ZD7288 was specific

for Ih. Thus the block of the oscillations by ZD7288 cannot beattributed to a nonspecific effect of the drug.

Finally, it also might be argued that the disappearance ofsustained subthreshold oscillations with ZD7288 resulted fromthe membrane conductance decrease due to theIh block and notby theIh block per se. This possibility was tested with the useof Ba21 (1–2 mM;n 5 7), which, in contrast to ZD7288, doesnot affectIh (cf. Fig. 3) but which, similarly to ZD7288, alsoproduces a major decrease in membrane conductance. Impor-tantly, and in sharp contrast to the ZD7288 results, bath su-perfusion with Ba21 resulted in both a significant increase inthe amplitude [1.26 0.5 mV; t(4) 5 2.7; P , 0.05] and asignificant decrease in the frequency [21.3 6 0.1 Hz; t(6) 58.8;P , 0.01] of the subthreshold oscillations (Fig. 9,D–F). Inconsequence, the above indicates thatIh plays an essential rolein the generation of rhythmic subthreshold oscillations by theSCs and that leak conductances can modulate their amplitudeand frequency through their effects on passive membrane prop-erties.

Role of Ih and INaP in the generation of subthresholdoscillations

Although the preceding experimental data indicate thatIh isnecessary for the genesis of subthreshold oscillations by the

FIG. 7. Ih is increased andVh shifted in a depolarized direction in raised [K1]o. A: Ih was evoked by hyperpolarizing voltagesteps from a holding potential of240 mV in a control Ringer solution containing 1 mM [K1]o. B: increasing the concentration of[K1]o markedly increased the amplitude ofIh evoked at the same potential levels.C: washing with 1 mM [K1]o reversed thisincrease.D: cells were held at240 and280 mV (top and bottom left,respectively) and the instantaneousI-V relations wereconstructed in both 1 and 10 mM [K1]o. Right: Vh in both 1 mM (●, n, and —) and 10 mM [K1]o (E, ▫, and - - -) was estimatedfrom the intersection of the extrapolated I/V relations constructed from240 (▫ andn) and280 mV (● andE). Note that in thiscell changing the [K1]o from 1 to 10 mM shifted the extrapolatedVh from 224 to 217 mV. All recordings were conducted in asolution containing 1mM TTX, 2 mM Co21, and 2 mM Ba21.

2570 DICKSON, MAGISTRETTI, SHALINSKY, FRANSE´ N, HASSELMO, AND ALONSO

SCs, previous studies have shown that these oscillations arealso dependent on the activation of a subthreshold persistentNa1 current (INaP) (Alonso and Llinas 1989). To generate anoscillatory phenomenon, a process is needed the action ofwhich feeds-back to slow down the rate of the process itselfand, most critically, a delay in the execution of the feedback. InSCs, the slow kinetics of activation and deactivation ofIh

potentially can implement such a feedback process. To furtherclarify the role ofIh in the generation of subthreshold oscilla-tions by the SCs and to complement the preceding experimen-tal data, we next implemented a simplified biophysical simu-lation of the subthreshold membrane voltage behavior of theseneurons. Using classical Hodgkin-Huxley formalism, a theo-retical reconstruction of the biophysical properties ofIh firstwas carried out. To be consistent with our experimental dataindicating both a fast and a slow kinetic component ofIh (seepreceding text; Fig. 5), we constructed activation plots of thecorresponding fast and slow conductance components (Gh1 andGh2, respectively; Fig. 11A, 1 and2) from which them`i curveswere derived directly by using a standard Boltzmann fitting(seeMETHODS and legend of Fig. 11). The voltage dependenceof the fast and slow rate constants (ai andbi; Fig. 11C, 1 and2) was estimated from the corresponding time constants of

activation and deactivation (ti; Fig. 11B, 1 and2) and them`i

curves, as explained in detail inMETHODS.The derived parameters for the kinetics and voltage depen-

dence ofIh and those forINaP as described previously (Mag-istretti and Alonso 1999) then were incorporated in a singlecompartment model of the SC (seeMETHODS). The model thenwas explored to test whether it could reproduce characteristiccurrent-clamp phenomena such as the sag in membrane poten-tial during hyperpolarizing current steps and the generation ofsubthreshold membrane potential oscillations.

Voltage responses to hyperpolarizing current steps in themodel SC are illustrated in Fig. 12A. Note that the model celldid display the typical delayed large-amplitude depolarizingsags in response to membrane hyperpolarization as well asrobust rebound potentials at the break of the hyperpolarizingcurrent pulses. More importantly, as shown in Fig. 12B, themodel SC also developed sustained rhythmic membrane po-tential oscillations in response to DC membrane depolarizationfrom its resting level to about253 mV. The combined exper-imental and model work thus demonstrates that in the SCs theinterplay betweenIh andINaP is essential for the generation ofsustained rhythmic subthreshold membrane potential oscilla-tions.

As illustrated in Fig. 12C, the use of the model SC also

FIG. 8. Ih is decreased and its reversal potential shifted in a hyperpolarized direction in lowered [Na1]o. A: Ih was evoked byhyperpolarizing voltage steps from a holding potential of240 mV in a control Ringer solution (151 mM [Na1]o). B: lowering[Na1]o to 26 mM (substitution withN-methyl-D-glucamine) diminished the amplitude ofIh evoked at the same potential levels.C:washing with control Ringer reversed this reduction.D: cells were held at240 and280 mV (top andbottom left,respectively)and the instantaneousI-V relations constructed in both 151 mM [Na1]o (control) and 26 mM [Na1]o (low Na1). Right: Vh in both151 mM [Na1]o (n, ●, and —) and 26 mM [Na1]o (▫, E, and - - -) was estimated from the intersection of the extrapolatedI-Vrelations constructed from240 (▫ andn) and280 mV (E and●). Note that in this cell changing the [Na1]o from 151 to 26 mMshifted the extrapolatedVh from 219.8 to245.9 mV. All recordings were conducted in a solution containing 1mM TTX, 2 mMCo21, and 2 mM Ba21.

2571Ih IN EC LAYER II NEURONS

allowed us to understand the dynamics of the interplay betweenINaP and Ih (as well asGNaP and Gh) during the oscillatorycycle. Note that at the trough of an oscillation (1st verticaldashed line)INaP andGNaP are at a minimum, whereasIh andGh are approaching, but not yet at, their respective maxima.This occurs after a certain delay (Gh maximum lagsGNaP andVm minimum by 21 ms at 3 Hz), and thus the attainment of amaximum level byIh coincides with the initiation of a depo-larizing phase. As depolarization proceeds,INaP rapidly in-creases becauseGNaPbecomes progressively activated. In turn,the depolarization boosted byINaPleads toGh deactivation andthus a decrease inIh, which slows down and eventually stopsthe depolarization. Note that at the peak of the oscillation (2nddashed line)Ih andGh (similarly to what reciprocally occurs atthe trough) are approaching, but they are still not at, a mini-mum. This occurs after the peak. Thus the deactivation ofGhis now responsible for initiating the repolarizing phase of theoscillation. Membrane hyperpolarization leads to a rapid de-crease inINaPthat boosts further hyperpolarization. Eventuallymembrane hyperpolarization leads to the new activation ofIhand a new oscillatory cycle is initiated.

It can be observed from the traces in Fig. 12C that INaPessentially changes instantaneously with changes inVm,whereas changes inIh follow with a certain delay (caused by itstime-dependent properties). This differential behavior may bebetter understood by plottingINaPandIh as a function ofVm asillustrated in Fig. 13. Note that although theIh curve demon-strates substantial hysteresis, theINaP curve demonstrates very

little to none. The increase and decrease inINaP during thedepolarizing and hyperpolarizing phase of an oscillation main-tains an almost perfect linear relationship with changes inVm.In contrast, the trajectory followed byIh during the depolariz-ing phase of the oscillation is substantially different from thatfollowed during the hyperpolarizing phase.Ih decreases slowlyduring the initial part of the depolarizing phase but this rate ofdecrease accelerates as the membrane potential approaches itspeak value. Similarly, during the hyperpolarizing phase of theoscillation Ih increases slowly during the initial portion, andthis rate of increase accelerates as the membrane potentialapproaches its minimal value. Basically, the hysteresis intro-duced by the kinetic properties ofIh implement a delayedfeedback mechanism to the voltage changes led byINaP thatallows sustained oscillatory activity to occur.

D I S C U S S I O N

The present results demonstrate that the robust hyperpolar-ization-activated time-dependent inward rectification displayedby the stellate cells from EC layer II is due to an inward currentthat we identified asIh on the basis of its pharmacological andbiophysical profile as described in other excitable cells (seerecent revisions by Clapham 1998; Pape 1996). Importantly,our study also points out that some of the specific biophysicalproperties ofIh in the stellate cells determine that this currentplays a major “pacemaker” role in the generation of the theta-like subthreshold oscillations typical of these neurons (Alonso

FIG. 9. Effects of ZD7288 and Ba21 on subthreshold oscillations.A and D: in control Ringer solution containing synapticblockers (seeMETHODS), DC depolarization elicited rhythmic subthreshold membrane potential oscillations in 2 different cells.B:in the cell shown inA, the subthreshold oscillatory activity was abolished after application of 50mM ZD7288. Inset in B showsthe membrane responses to a 0.1-nA hyperpolarizing pulse before and after ZD7288 application, demonstrating a full block oftime-dependent inward rectification.C: autocorrelation analysis of membrane potential in the just-subthreshold range before andafter application of ZD7288 (— and - - -, respectively). Note lack of rhythmical oscillatory activity after drug application.E: in thecell shown inD, the oscillations were still present after application of 1 mM Ba21; however, they were increased markedly inamplitude (from 3.6 to 4.5 mV: peak to peak), and reduced in frequency (from 2.1 to 1.1 Hz).Inset in D shows the membraneresponse to a hyperpolarizing pulse of 0.1-nA demonstrating a large increase in input resistance without block of time-dependentinward rectification.F: autocorrelation analysis of membrane potential in the just-subthreshold range before and after applicationof Ba21 (— and - - -, respectively). Note the longer period of rhythmical oscillatory activity after Ba21 application.

2572 DICKSON, MAGISTRETTI, SHALINSKY, FRANSE´ N, HASSELMO, AND ALONSO

and Llinas 1989). Indeed, the combined experimental andmodeling analysis we carried out indicate that the kineticproperties of both the activation and deactivation ofIh imple-ment a delayed feed-back mechanism to the voltage changesled by a subthreshold Na1 current that allows persistent sub-threshold oscillatory activity to occur. Althoughh currentshave been shown to contribute to rebound activity in manyneuronal types (Crepel and Penit-Soria 1986; Mayer and West-brook 1983; Spain et al. 1987) and to interact with low-threshold Ca21 currents (at rather negative levels; about270mV) to generate oscillatory activity (Bal and McCormick1997; Brown and DiFrancesco 1980; McCormick and Pape1990), to our knowledge, the SCs are the first case in which anh current has been shown to generate persistent oscillatoryactivity by interacting with a sustained Na1 current in thesubthreshold voltage range (about255 mV). Intrinsic oscilla-tory activity in the subthreshold voltage range may be offundamental importance in defining the integrative propertiesof the participating neurons (cf. Hopfield 1995; Lampl andYarom 1993; Llinas 1988).

A characteristic pharmacological feature ofIh in all othercell types studied is its blockade by Cs1 but not Ba21. Con-sistent with this,Ih in the SCs was largely reduced, though notcompletely blocked, by Cs1 (1–6 mM) and not significantlyaffected by Ba21 (1–2 mM). The incomplete block ofIh byCs1, which also has been observed in other cells (Champignyand Lenfant 1986; Ishii et al. 1999), prompted us to use therecently described bradycardic agent ZD7288 (BoSmith et al.1993). As found in other neurons (Gasparini and DiFrancesco1997; Harris and Constanti 1995; Lu¨thi et al. 1998; Maccaferri

and McBain 1996), ZD7288 fully and selectively (in the volt-age range explored:245 to2120 mV) blockedIh in the SCs.

In general terms, the biophysical properties ofIh as recordedin the SCs were similar to those reported for other brainneurons though some differences in activation threshold andreversal potential were apparent, and these appear to be func-tionally significant.Ih in the SCs appeared as a noninactivatinginward current that turned on (activated) relatively slowly withhyperpolarization and also turned off (deactivated) slowly withdepolarization. As with other voltage-dependent currents, theconductance that gives rise toIh showed a sigmoidal activationcurve with a threshold at around245 mV, a half-activationpoint at277 mV, and a slope factor of 12.1. A half-activationpoint in the range of270 to 280 mV is typical of otherneurons; however, theIh activation threshold that we observedin the SCs was;10–15 mV more positive than that reportedfor other subcortical or cortical neurons (Bayliss et al. 1994;Crepel and Penit-Soria 1986; Halliwell and Adams 1982; Ka-mondi and Reiner 1991; Mayer and Westbrook 1983; McCor-mick and Pape 1990; Mercuri et al. 1995; Spain et al. 1987).An activation threshold at about245 mV is, however, com-parable with that found in cardiac sinoatrial cells and Purkinjefibers (DiFrancesco 1981; DiFrancesco et al. 1986; Yanagiharaand Irisawa 1980) and to recent findings for hippocampal andbrain stem neurons (Maccaferri and McBain 1996; Maccaferriet al. 1993; Travagli and Gillis 1994). In addition, the esti-mated reversal potential forIh in the SCs (about220 mV) was10–30 mV more positive than that reported for other subcor-tical or cortical neurons (Bayliss et al. 1994; Crepel and Penit-Soria 1986; Halliwell and Adams 1982; Kamondi and Reiner

FIG. 10. Blocking effect of ZD7288 on the subthreshold membrane potential oscillations can be attributed to the blocking actionof this drug onIh. A: subthrehold oscillations in the SCs develop on DC depolarization to the voltage range between about255to 250 mV. B: voltage-clamp experiment in the same cell showing outward current relaxations (corresponding toIh deactivation)during depolarizing voltage steps through the oscillatory voltage level (255, 250, and245 mV), as well as inward currentrelaxations (corresponding toIh activation) on return to the holding potential (260 mV).C: addition of 100mM ZD7288 eliminatedIh and thus caused a robust outward shift in the holding current and the concomitant disappearance of the outward currentrelaxations on depolarization as well as the inward current relaxations on return to the holding potential.D: superimposition oftraces inB andC. Note that control and ZD7288 current traces to245 mV overlap perfectly at steady-state, thus indicating thatin the voltage range examined the actions of ZD7288 were selective forIh. Voltage-clamp recordings were performed in thepresence of 1mM TTX and 2 mM Co21.

2573Ih IN EC LAYER II NEURONS

1991; Mayer and Westbrook 1983; McCormick and Pape1990; Mercuri et al. 1995; Spain et al. 1987), though, again,very similar to the value reported for cardiac sinoatrial cellsand Purkinje fibers (around225 mV) (DiFrancesco 1981;DiFrancesco et al. 1986; Yanagihara and Irisawa 1980). Areversal potential around220 mV clearly suggests thatIh inthe SCs, as in all other excitable cells studied, is a mixedcationic current carried by both Na1 and K1 ions. Consistentwith this interpretation, we found that raising [K1]o shiftedVh

in the depolarizing direction (without any significant effect onthe activation curve) (Halliwell and Adams 1982; Mayer andWestbrook 1983; Spain et al. 1987; Takahashi 1990), whereaslowering [Na1]o shifted Vh in the hyperpolarizing direction.Alterations in [Cl2]i changed neither the reversal potential northe activation curve corresponding to this current, suggestingthat it is cation-specific.

With regard to the time-dependent properties ofIh in theSCs, an interesting feature was their biexponential kineticnature. We found that the time course of activation ofIh wasdescribed best by a biexponential function having fast (t1) andslow (t2) time constants that were voltage dependent—de-creasing with increasing hyperpolarization. A dual exponentialnature of theIh deactivation time course also was confirmed,with fast and slow time constants that also showed voltagedependency—becoming faster with increasing depolarization.

Although many of the studies onIh report a time course ofactivation that appears to be well fitted by a mono-exponentialfunction (Crepel and Penit-Soria 1986; DiFrancesco et al.1986; Halliwell and Adams 1982; McCormick and Pape 1990),some also have shown that a biexponential fitting best repre-sented the time course ofIh in the corresponding cells (Mayerand Westbrook 1983; Solomon and Nerbonne 1993; Spain etal. 1987). The fact that the activation time constants reported indifferent neurons (measured at similar temperatures) rangefrom tens of milliseconds (Crepel and Penit-Soria 1986),through hundreds of milliseconds (Kamondi and Reiner 1991)and even seconds (McCormick and Pape 1990; Soltesz et al.1991), might be interpreted as indicative of multipleIh channelsubtypes with different gating properties. Thus cells expressingmore than oneIh channel subtype would show multiple kineticcomponents during both the activation and deactivation pro-cesses. Strong evidence for this comes from a zebrafish mutant(slow mo), which exhibits a slower than normal heart rate. Avoltage-clamp analysis in isolated cardiac myocytes from thewild-type and the mutant zebrafish demonstrated thatIh wasselectively decreased in the mutant. Importantly, this decreaseappeared to result from the selective diminution of the fastkinetic component of the current (Baker et al. 1997). Finally,very recently, three different groups have cloned various genesresponsible forIh clearly demonstrating the existence of a

FIG. 11. Modeling of the biophysical proper-ties of Ih. A: voltage dependence of the normal-ized conductances derived from amplitude coef-ficients returned by double-exponential fittings ofaverageIh traces. Amplitude coefficients relatingto the fast and slow time constants (A, 1 and 2,respectively) were used to calculate the values ofthe underlying conductances (GA1 and GA2, re-spectively) as explained inMETHODS. Boltzmannfittings to the experimental data points also areshown. Fitting parameters are:V1⁄21 5 267.4 mV,k1 5 12.66 mV (A1); V1⁄22 5 257.92 mV,k2 59.26 mV (A2). B: voltage dependence of fast (1)and slow (2) time constants (t1 and t2, respec-tively) returned by double-exponential fittings ofaverageIh traces. —, theoretical functions de-scribing the behavior of the experimental plots, asobtained from the theoretical functions derivedfrom fittings of rate-constant plots (see followingtext; see alsoMETHODS for details). C: voltagedependence of fast and slow rate constants (a1,b1, anda2, b2, respectively) calculated from nor-malized conductance values and time-constantvalues as explained inMETHODS. ●, “on” reactionrate constants (a); E, “off”-reaction rate constants(b). —, fittings obtained using the theoreticalfunction described inMETHODS. Fitting parame-ters are:a 5 22.89 z 1023 mV21 z ms21, b 520.445 ms21, k 5 24.02 mV (a1); a 5 2.71 z1022 mV21 z ms21, b 5 21.024 ms21, k 5217.4 mV (b1); a 5 23.18z 1023 mV21 z ms21,b 5 20.695 ms21, k 5 26.72 mV (a2); a 52.16 z 1022 mV21 z ms21, b 5 21.065 ms21,k 5 214.25 mV (b2).

2574 DICKSON, MAGISTRETTI, SHALINSKY, FRANSE´ N, HASSELMO, AND ALONSO

family of hyperpolarization-activated cation channels with dif-ferent activation kinetics (Gauss et al. 1998; Ludwig et al.1998, 1999; Santoro et al. 1998).

Function of Ih in the stellate cells

In the stellate cells, the rather positive values that we foundfor Ih activation (245 mV) and reversal potential (220 mV)indicate that in these neuronsIh must be a major contributor tothe resting membrane potential (around260 mV) (Alonso and

Klink 1993). Indeed, full block ofIh with ZD7288 induced arobust hyperpolarization of the resting membrane potential incurrent-clamp conditions and a large outward shift in theholding current at260 mV in voltage-clamp conditions. Inaddition, in the stellate cells,Ih also efficiently modulatesmembrane hyperpolarization, as observed by the pronounceddepolarizing sags in response to membrane hyperpolarizationand the robust rebound potentials that follow. In this respect,the stellate cells are densely innervated (Finch et al. 1988;Jones 1994; Jones and Buhl 1993; Kohler 1988; Kohler et al.

FIG. 12. Modeling of the current-clamp behavior of SCs.A: current-clampV-I simulation in a simplified model of a SC. In thissimulation, 600-ms hyperpolarizing current pulses of20.49,20.98, and21.46 pA/pF were injected into the cell from the restingmembrane potential level (256.7 mV). Simulation parameters were:GhMax 5 98.0 pS/pF,Gh1Max /Gh2Max5 1.85,GNaPMax5 17.4pS/pF,GlMax 5 78.0 pS/pF,Vh 5 220 mV,VNa 5 187 mV,Vl 5 283 mV.B: simulations ofIh- andINaP-dependent subthresholdmembrane-voltage (Vm) oscillations. Oscillations were elicited by injection of steady depolarizing current of variable amplitude(from bottomto top: 0.00,10.19, and10.21 pA/pF). Note the absence of oscillations at rest (256.7 mV) and the gradual increasein the amplitude of the oscillations with increasing current amplitude. Simulation parameters were the same as described forA. C:expanded view of the simulated membrane potential oscillations evoked by the injection of10.21 pA/pF and of the behavior ofthe underlying currents (Ih, INaP) and conductances (Gh, GNaP). Vertical, dashed lines help in the appraisal of the phase relationshipsbetween the membrane potential, the currents, and their conductances. Note the almost perfect phase correspondence betweenVm

and INaP, although there appears to be a phase lag betweenVm and Ih.

FIG. 13. Phase plane plot of the simulated mem-brane potential vs. the amplitude ofINaP and Ih. Forthe simulation shown in Fig. 12C, the amplitude ofboth INaP and Ih were plotted with respect to themembrane potential. Plot corresponding toINaPtrav-els a straight line with negative slope showing essen-tially no hysteresis. Therefore during both the up-swing and downswing of the oscillation there is arelatively “instantaneous” increase and decrease, re-spectively, in the amplitude ofINaP. On the otherhand, the plot corresponding toIh demonstrates alarge degree of hysteresis. Therefore oppositely toINaP, during the upswing and downswing of the os-cillation there is a delayed decrease and increase,respectively, in the amplitude ofIh. Because ofthisdelay, the maxima and minima of the amplitude ofIh

tend to occur just after the membrane potential hasachieved its most hyperpolarized and depolarizedvalues, respectively (i.e., just after the commence-ment of the upswing and downswing of the oscilla-tion, respectively).

2575Ih IN EC LAYER II NEURONS

1985; Wouterlood et al. 1995) by GABAergic inputs, and wehave observed thatIh efficiently limits the inhibitory postsyn-aptic potentials that these inputs trigger. IndeedIh activation inresponse to aninhibitory postsynaptic potentialimplements a“resetting” mechanism for the intrinsic oscillations (Alonsoand Dickson, unpublished observations).

As directly demonstrated in this study, a major role ofIh inthe stellate cells is that of a “pacemaker” current. This role isnot, however, one of providing a “background” depolarizingcurrent that sustains rhythmic discharge but of contributing tothe generation of persistent subthreshold oscillatory activity.We confirmed that the bradycardic agent ZD7288 selectivelyabolishedIh in the range of potentials corresponding to theoscillatory range and found that it progressively and com-pletely abolished the rhythmic subthreshold oscillations. Thisevidence directly implicatesIh in their generation.

In the present study, we found no evidence of a slowoutward K1 conductance, such as the M current, being neces-sary for the generation of the subthreshold oscillations by thestellate cells. Indeed Ba21, which blocks the M-current as wellas leak K1 conductances andIKir, increased the amplitude ofthe oscillations, which also became more regular and lower infrequency. This result is in contrast to our previous study usingsharp electrodes (Alonso and Klink 1993) in which we reportedan apparent block of the oscillations with Ba21. However, inthat study Ba21 was applied in the absence of glutamatergicand GABAergic neurotransmission and thus the apparent blockmay have been due to the increased synaptic backgroundactivity that is caused by Ba21. In addition, because Ba21 alsoprolonged the action potential, allowing for an increased Ca21

influx and an enhancement of the slow afterhyperpolarization(see Fig. 10 in Klink and Alonso 1993), it could have been thatthe apparent block of the oscillations also was related to amembrane shunting effect by a Ca21-dependent conductance.To test these possibilities, we performed additional sharp elec-trodes recordings from stellate cells in a similar manner aspreviously described (Alonso and Klink 1993) in which Ba21

application (1 mM) was conducted in the presence of gluta-matergic and GABAergic antagonists and/or the Ca21 channelblocker Co21 (2 mM) (n 5 6; not shown). In these experi-ments, in agreement with the above interpretations, the appli-cation of Ba21 during synaptic and Ca21 conductance blockwith Co21 did not abolish the oscillations, which, however, didbecome slower in frequency (as in our present whole cell patchstudy). In addition, a similar result was observed when werepeated the preceding experiment using perforated patch re-cordings (Horn and Marty 1988) (n 5 2; not shown). We did,however, observe that, using either sharp electrodes or theperforated patch technique, block of Ca21 conductances withCo21 was required to best reproduce our present results usingconventional whole cell patch recording. This suggests that arun-down of Ca21 conductances occurred during the conven-tional whole cell technique that allowed for a better expressionof the oscillations in the presence of Ba21 (as seen in Fig. 9).

The modeling analysis that was carried out on the basis ofthe biophysical properties ofIh and INaP, as experimentallyderived from the stellate cells, indicated that the interplaybetween these two currents appears sufficient for the genera-tion of persistent subthreshold oscillations in these neurons.First, in the SCs, theGh activation curve has a rather positivethreshold for activation (at about245 mV) and overlaps sub-

stantially with theGNaP activation curve (threshold at about265 mV) (Magistretti and Alonso 1999), thus setting the stagefor their potential interplay because these conductances acti-vate with voltage changes in opposite directions. Second, andmost importantly, the slow activation and deactivation kineticsof Ih implement a delayed feed-back mechanism to the voltagechanges led by the “instantaneous” changes inINaP, thus caus-ing the emergence of sustained oscillatory activity (given thepassive membrane properties of the cells). Alonso and Llina´s(1989) initially had postulated an interaction betweenIh andINaPas the basis for the generation of subthreshold oscillationsby the SCs. Using a bifurcation analysis, White et al. (1995)also had implicatedIh (in addition to an as-yet undeterminedslow outward rectifier) in the generation of these oscillations.More recently, an interaction betweenIh andINaPalso has beenproposed as the basis for subthreshold membrane potentialresonance in response to oscillatory intracellular current injec-tion in sensorimotor cortex neurons (Hutcheon et al. 1996a,b).During the last 10 years, Na1-dependent subthreshold oscilla-tions similar to those expressed by the SCs also have beenobserved in other brain neurons (Alonso et al. 1996; Amitai1994; Gutfreund et al. 1995; Leung and Yim 1991; Llina´s et al.1991; Pape et al. 1998), and they have been considered typi-cally to emerge from the interplay betweenINaP and an out-ward K1 current, such as an M-type current (Gutfreund et al.1995; Llinas et al. 1991; Pape et al. 1998; White et al. 1995,1998) or an A-type current (Wang 1993). As in other situa-tions, distinct ionic mechanisms may be responsible for asimilar electrophysiological event in different neurons accord-ing to their specific integrative needs.

Interestingly, we found that complete abolition of the rhyth-mic subthreshold oscillations required a very substantial reduc-tion of Ih. Cs1, which we found to reduceIh by ;65%, did notcompletely abolish the oscillations, thus suggesting that theoscillatory process per se is very robust and rather insensitiveto variations in the level ofIh expression. Indeed our simplifiedmodel maintained sustained oscillatory activity even whenIhwas decreased by 50% (not shown). Further reductions causeda very rapidly damping oscillation. However, in an elaboratedSC model that we have constructed (Franse´n et al. 1998) whenthe impact of channel noise (White et al. 1998) is considered,the SC model behaved much more closely to the “real” cell inthat Ih reductions of up to;70% were required to completelyabolish subthreshold rhythmicity. The critical importance ofstochastic noise in sculpturing the membrane voltage behaviorof the SCs has been treated in detail by White et al. (1998).

Functional role of the subthreshold oscillations

The potential functional role of subthreshold oscillatoryactivity in neocortical neurons was reviewed recently by Con-nors (Connors and Amitai 1997) and also has been treated indetail by others (Engel et al. 1992; Gray 1994; Laurent 1996;Singer 1993). We will discuss briefly here the potential func-tional implications that the presence of rhythmic subthresholdoscillatory activity in the stellate cells may have with regard totemporal lobe function. A well-established role of the entorhi-nal network is that of memory function. Indeed the stellatecells from EC layer II occupy a privileged position in theneocortico-hippocampo-neocortical circuit. They are the tar-gets of convergent information from polysensory associational

2576 DICKSON, MAGISTRETTI, SHALINSKY, FRANSE´ N, HASSELMO, AND ALONSO

areas that they funnel to the dentate gyrus of the hippocampalformation via the perforant path. Thus a potential role of theintrinsic oscillatory activity of the stellate cells could be that ofimplementing a synchronizing mechanism by which conver-gent sensory information is coordinated temporally for itstransfer to the hippocampal processing machinery and theformation of a memory event (Buzsa´ki 1989, 1996; Lopes daSilva et al. 1985; Rudell et al. 1980; ). Although still far fromdirectly testing such a hypothesis given the complex nature ofthe information processed by the EC, we do know that EClayer II neurons are powerful generators of theta rhythmicity“in vivo” (Alonso and Garcı´a-Austt 1987a,b; Dickson et al.1995; Mitchell and Ranck 1980), and the theta rhythm has beenimplicated in memory processes (Greenstein et al. 1988; Hol-scher et al. 1997; Huerta and Lisman 1993; Larson et al. 1986;Lisman and Idiart 1995; Pavlides et al. 1988; Winson 1978).Importantly, analysis of inhibitory synaptic input on EC layerII neurons show that inhibitory postsynaptic potentials, whichEC stellates are likely to receive during theta (cf. Fox 1989;Leung and Yim 1986; Soltesz and Descheˆnes 1993; Ylinen etal. 1995) can very efficiently reset the intrinsic oscillations(Alonso and Dickson, unpublished observations; see also Cobbet al. 1995). In addition, theoretical, experimental and model-ing studies also indicate that the intrinsic subthreshold oscilla-tions of individual cells can engender synchronized populationactivity (Engel et al. 1992; Franse´n et al. 1999; Jefferys et al.1996; Konig et al. 1995; Lampl and Yarom 1993; Singer 1993;Steriade et al. 1990).

In conclusion, the stellate cells from EC layer II display avery robustIh current, which, in conjunction withINaP, givesrise to the expression of rhythmic subthreshold oscillations ofthe membrane potential in these cells. These oscillations arelikely to implement a synchronizing mechanism that may beimportant in the memory function of the temporal lobe. Un-derstanding the biophysical/molecular and neuromodulatoryaspects of the distinct voltage-gated ion channels underlyingIhand INaP thus may be of critical importance in exploring thefunctions of the EC network.

We thank I. Manns for kind and efficient aid in Matlab programming.This study was funded by grants from the Fonds de la Recherche en Sante´

du Quebec and the Medical Research Council (MRC) to A. Alonso and a grantfrom the Human Frontier Science Program Organization (HFSP) to A. Alonsoand M. Hasselmo. C. T. Dickson was supported by fellowships from theNational Sciences and Engineering Research Council and the Savoy Founda-tion. M. H. Shalinsky was supported by a studentship from the MRC. E.Franse´n was supported by the Swedish Foundation for International Cooper-ation in Research in Higher Education (STINT) and the P. E. Lindahls fund.

Address for reprint requests: A. Alonso, Dept. of Neurology and Neurosur-gery, McGill University, 3801 University St., Montreal, Quebec H3A 2B4,Canada.

Received 27 August 1999; accepted in final form 11 January 2000.

REFERENCES

ALONSO, A. AND GARCIA-AUSTT, E. Neuronal sources of theta rhythm in theentorhinal cortex of the rat. I. Laminar distribution of theta field potentials.Exp. Brain Res.67: 493–501, 1987a.

ALONSO, A. AND GARCIA-AUSTT, E. Neuronal sources of theta rhythm in theentorhinal cortex of the rat. II. Phase relations between unit discharges andtheta field potentials.Exp. Brain Res.67: 502–509, 1987b.

ALONSO, A., KHATEB, A., FORT, P., JONES, B. E., AND MUHLETHALER, M.Differential oscillatory properties of cholinergic and non-cholinergic neu-rons in the nucleus basalis.Eur. J. Neurosci.8: 169–182, 1996.

ALONSO, A. AND KLINK , R. Differential electroresponsiveness of stellate andpyramidal-like cells of medial entorhinal cortex layer II.J. Neurophysiol.70: 128–143, 1993.

ALONSO, A. AND LLINAS, R. R. Subthreshold Na1-dependent theta-like rhyth-micity in stellate cells of entorhinal cortex layer II.Nature342: 175–177,1989

AMITAI , Y. Membrane potential oscillations underlying firing patterns in neo-cortical neurons.Neuroscience63: 151–161, 1994.

BAKER, K., WARREN, K. S., YELLEN, G., AND FISHMAN, M. C. Defective“pacemaker” current (Ih) in a zebrafish mutant with a slow heart rate.Proc.Natl. Acad. Sci. USA94: 4554–4559, 1997.

BAL, T. AND MCCORMICK, D. A. Synchronized oscillations in the inferior oliveare controlled by the hyperpolarization-activated cation current.J. Neuro-physiol.77: 3145–3156, 1997.

BAYLISS, D. A., VIANA , F., BELLINGHAM , M. C., AND BERGER, A. J. Charac-teristics and postnatal development of a hyperpolarization-activated inwardcurrent in rat hypoglossal motoneurons in vitro.J. Neurophysiol.71: 119–128, 1994.

BOSMITH, R. E., BRIGGS, I., AND STURGESS, N. C. Inhibitory actions ofZENECA ZD7288 on whole-cell hyperpolarization activated inward current(If) in guinea-pig dissociated sinoatrial node cells.Br. J. Pharmacol.110:343–349, 1993.

BROWN, H. AND DIFRANCESCO, D. Voltage-clamp investigations of membranecurrents underlying pace-maker activity in rabbit sino-atrial node.J. Physiol.(Lond.)308: 331–351, 1980.

BUZSAKI , G. Two-stage model of memory trace formation: a role for “noisy”brain states.Neuroscience31: 551–570, 1989.

BUZSAKI , G. The hippocampo-neocortical dialogue.Cereb. Cortex6: 81–92,1996.

CHAMPIGNY, G.AND LENFANT, J. Block and activation of the hyperpolarization-activated inward current by Ba and Cs in frog sinus venosus.Pflugers ArchEur. J. Physiol.407: 684–690, 1986.

CLAPHAM, D. E. Not so funny anymore: pacing channels are cloned.Neuron21: 5–7, 1998.

COBB, S. R., BUHL, E. H., HALASY, K., PAULSEN, O., AND SOMOGYI, P.Synchronization of neuronal activity in hippocampus by individualGABAergic interneurons.Nature378: 75–78, 1995.

CONNORS, B. W. AND AMITAI , Y. Making waves in the neocortex.Neuron18:347–349, 1997.

CREPEL, F. AND PENIT-SORIA, J. Inward rectification and low threshold calciumconductance in rat cerebellar purkinje cells. An in vitro study.J. Physiol.(Lond.)372: 1–23, 1986.

CRILL, W. E. Persistent sodium current in mammalian central neurons.Annu.Rev. Physiol.58: 349–362, 1996.

DICKSON, C. T. AND ALONSO, A. Hyperpolarization-activated current (Ih) inlayer II cells of the medial entorhinal cortex.Soc. Neurosci. Abstr.22: 1027,1996.

DICKSON, C. T. AND ALONSO, A. Role of Ih in the generation of subthresholdmembrane potential oscillations (MPOs) of entorhinal cortex (EC) layer IIstellate cells.Soc. Neurosci. Abstr.24: 2035, 1998.

DICKSON, C. T., KIRK, I. J., ODDIE, S. D.,AND BLAND, B. H. Classification oftheta-related cells in the entorhinal cortex: cell discharges are controlled bythe ascending brainstem synchronizing pathway in parallel with hippocam-pal theta-related cells.Hippocampus5: 306–319, 1995.

DICKSON, C. T., MENA, A. R., AND ALONSO, A. Electroresponsiveness ofmedial entorhinal cortex layer III neurons in vitro.Neuroscience81: 937–950, 1997.

DIFRANCESCO, D. A new interpretation of the pace-maker current in calfPurkinje fibers.J. Physiol. (Lond.)314: 359–376, 1981.

DIFRANCESCO, D. Pacemaker mechanisms in cardiac tissue.Annu. Rev.Physiol.55: 455–472, 1993.

DIFRANCESCO, D., FERRONI, M., MAZZANTI , M., AND TROMBA, C. Properties ofthe hyperpolarizing-activated current (If) in cells isolated from the rabbitsino-atrial node.J. Physiol. (Lond.)377: 61–88, 1986.

DOLORFO, C. L. AND AMARAL , D. G. Entorhinal cortex of the rat: topographicorganization of the cells of origin of the perforant path projection to thedentate gyrus.J. Comp. Neurol.398: 25–48, 1998.

ENGEL, A. K., KONIG, P., KREITER, A. K., SCHILLEN, T. B., AND SINGER, W.Temporal coding in the visual cortex: new vistas on integration in thenervous system.Trends Neurosci.15: 218–226, 1992.

FINCH, D. M., TAN, A. M., AND ISOKAWA-AKESSON, M. Feedforward inhibitionof the rat entorhinal cortex and subicular complex.J. Neurosci.8: 2213–2226, 1988.

2577Ih IN EC LAYER II NEURONS

FOX, S. E. Membrane potential and impedance changes in hippocampal pyra-midal cells during theta rhythm.Exp. Brain Res.77: 283–294, 1989.

FRANSEN, E., DICKSON, C. T., MAGISTRETTI, J., ALONSO, A., AND HASSELMO,M. E. Modeling the generation of subthreshold membrane potential oscil-lations of entorhinal cortex layer II stellate cells.Soc. Neurosci. Abstr.24:814.815, 1998.

FRANSEN, E., WALLENSTEIN, G. V., ALONSO, A., DICKSON, C. T., AND HAS-SELMO, M. E. A biophysical simulation of intrinsic and network propertiesof entorhinal cortex.Neurocomputing26–27: 375–380, 1999.

GASPARINI, S.AND DIFRANCESCO, D. Action of the hyperpolarization-activatedcurrent (Ih) blocker ZD 7288 in hippocampal CA1 neurons.Pflugers Arch.Eur. J. Physiol.435: 99–106, 1997.

GAUSS, R., SEIFERT, R., AND KAUPP, U. B. Molecular identification of ahyperpolarization-activated channel in sea urchin sperm.Nature393: 583–587, 1998.

GRAY, C. M. Synchronous oscillations in neuronal systems: mechanisms andfunctions.J. Comput. Neurosci.1: 11–38, 1994.

GREENSTEIN, Y. J., PAVLIDES, C.,AND WINSON, J. Long-term potentiation in thedentate gyrus is preferentially induced at theta rhythm periodicity.BrainRes.438: 331–334, 1988.

GUTFREUND, Y., YAROM, Y., AND SEGEV, I. Subthreshold oscillations andresonant frequency in guinea-pig cortical neurons: physiology and model-ling. J. Physiol. (Lond.)483: 621–640, 1995.

HAGIWARA, S., MIYAZAKI , S., MOODY, W., AND PATLAK , J. Blocking effects ofbarium and hydrogen ions on the potassium current during anomalousrectification in the starfish egg.J. Physiol. (Lond.)279: 167–185, 1978.

HAGIWARA, S., MIYAZAKI , S.,AND ROSENTHAL, N. P. Potassium current and theeffect of cesium on this current during anomalous rectification of the eggcell membrane of a starfish.J. Gen. Physiol.67: 621–638, 1976.

HALLIWELL , J. V. AND ADAMS, P. R. Voltage-clamp analysis of muscarinicexcitation in hippocampal neurons.Brain Res.250: 71–92, 1982.

HARRIS, N. C. AND CONSTANTI, A. Mechanism of block by ZD 7288 of thehyperpolarization-activated inward rectifying current in guinea pig substan-tia nigra neurons in vitro.J. Neurophysiol.74: 2366–2378, 1995.

HILLE, B. Ionic Channels of Excitable Membranes.Sunderland, MA: SinauerAssoc., 1992.

HOLSCHER, C., ANWYL, R., AND ROWAN, M. J. Stimulation on the positivephase of hippocampal theta rhythm induces long-term potentiation that canbe depotentiated by stimulation on the negative phase in area CA1 in vivo.J. Neurosci.17: 6470–6477, 1997.

HOPFIELD, J. J. Pattern recognition computation using action potential timingfor stimulus representation.Nature376: 33–36, 1995.

HORN, R. AND MARTY, A. Muscarinic activation of ionic currents measured bya new whole-cell recording method.J. Gen. Physiol.92: 145–159, 1988.

HUERTA, P. T.AND LISMAN, J. E. Heightened synaptic plasticity of hippocam-pal CA1 neurones during a cholinergically induced rhythmic state.Nature364: 723–725, 1993.

HUTCHEON, B., MIURA, R. M., AND PUIL, E. Models of subthreshold membraneresonance in neocortical neurons.J. Neurophysiol.76: 698–714, 1996a.

HUTCHEON, B., MIURA, R. M., AND PUIL, E. Subthrehold membrane resonancein neocortical neurons.J. Neurophysiol.76: 683–697; 1996b.

ISHII, T. M., TAKANO, M., XIE, L.-H., NOMA, A., AND OHMORI, H. Molecularcharacterization of the hyperpolarization-activated cation channel in rabbithear sinoatrial node.J. Biol. Chem.274: 12835–12839, 1999.

JEFFERYS, J., TRAUB, R.,AND WHITTINGTON, M. Neuronal networks for induced“40 Hz” rhythms.Trends Neurosci.19: 202–208, 1996.

JONES, R.S.G. Synaptic and intrinsic properties of neurons of origin of theperforant path in layer II of the rat entorhinal cortex in vitro.Hippocampus4: 335–353, 1994.

JONES, R.S.G.AND BUHL, E. H. Basket-like interneurons in layer II of theentorhinal cortex exhibit a powerful NMDA-mediated synaptic excitation.Neurosci. Lett.149: 35–39, 1993.

KAMONDI, A. AND REINER, P. B. Hyperpolarization-activated inward current inhistaminergic tuberomammillary neurons of the rat hypothalamus.J. Neu-rophysiol.66: 1902–1911, 1991.

KLINK , R. AND ALONSO, A. Ionic mechanisms for the subthreshold oscillationsand differential electroresponsiveness of medial entorhinal cortex layer IIneurons.J. Neurophysiol.70: 144–157, 1993.

KLINK , R. AND ALONSO, A. Morphological characteristics of layer II projectionneurons in the rat medial entorhinal cortex.Hippocampus7: 571–583, 1997.

KOHLER, C. Cytochemical architecture of the entorhinal area.Adv. Exp. Med.Biol. 203: 83–98, 1988.

KOHLER, C., CHAN-PALAY , V., AND WU, J.-Y. Neurons and terminals in theretrohippocampal region in the rat’s brain identified by anti-g-aminobutyric

acid and anti-glutamic acid decarboxylase immunocytochemistry.Anat.Embryol.173: 35–44, 1985.

KONIG, P., ENGEL, A. K., AND SINGER, W. Relation between oscillatory activityand long-range synchronization in cat visual cortex.Proc. Natl. Acad. Sci.USA92: 290–294, 1995.

LAMPL, I. AND YAROM, Y. Subthreshold oscillations of the membrane potential:a functional synchronizing and timing device.J. Neurophysiol.70: 2181–2186, 1993.

LARSON, J., WONG, D., AND LYNCH, G. Patterned stimulation at the thetafrequency is optimal for the induction of hippocampal long-term potentia-tion. Brain Res.368: 347–350, 1986.

LAURENT, G. Dynamical representation of odors by oscillating and evolvingneural assemblies.Trends Neurosci.19: 489–496, 1996.

LEUNG, L.-W.S. AND YIM, C. Y. Intracellular records of theta rhythm inhippocampal CA1 cells of the rat.Brain Res.367: 323–327, 1986.

LEUNG, L.-W.S.AND YIM, C.-Y.C. Intrinsic membrane potential oscillations inhippocampal neurons in vitro.Brain Res.553: 261–274, 1991.

LISMAN, J. E.AND IDIART, M.A.P. Storage of 76 2 short term memories inoscillatory subcycles.Science267: 1512–1515, 1995.

LLINAS, R. R. The intrinsic electrophysiological properties of mammalianneurons: Insights into central nervous system function.Science242: 1654–1664, 1988.

LLINAS, R. R., GRACE, A. A., AND YAROM, Y. In vitro neurons in mammaliancortical layer 4 exhibit intrinsic oscillatory activity in the 10- to 50-Hzfrequency range.Proc. Natl. Acad. Sci. USA88: 897–901, 1991.

LLINAS, R. AND JAHNSEN, H. Electrophysiology of mammalian thalamic neu-rones in vitro.Nature297: 406–408, 1982.

LOPES DA SILVA , F. H., GROENEWEGEN, H. J., HOSHEIMER, J., ROOM, P.,WITTER, M. P., VAN GROEN, T., AND WADMAN , W. J. The hippocampus asa set of partially overlapping segments with a topographically organizedsystem of inputs and outputs: the entorhinal cortex as a sensory gate, themedial septum as a gain-setting system and the ventral striatum as a motorinterface. In:Electrical Activity of the Archicortex,edited by G. Buzsaki andC. H. Vanderwolf. Budapest: Akademiai Kiado, 1985, p. 83–106.

LUDWIG. A., ZONG, X., JEGLITSCH, M., HOFMANN, F.,AND BIEL, M. A family ofhyperpolarization-activated mammalian cation channels.Nature393: 587–591, 1998.

LUDWIG, A., ZONG, X., STIEBER, J., HULLIN , R., HOFMANN, F., AND BIEL, M.Two pacemaker channels from human heart with profoundly differentactivation kinetics.EMBO J.18: 2323–2329, 1999.

LUTHI, A., BAL, T., AND MCCORMICK, D. A. Periodicity of thalamic spindlewaves is abolished by ZD7288, a blocker ofIh. J. Neurophysiol.79:3284–3289, 1998.

LUTHI, A. AND MCCORMICK, D. A. H-current: properties of a neuronal andnetwork pacemaker.Neuron21: 9–12, 1998.

MACCAFERRI, G., MANGONI, M., LAZZARI, A., AND DIFRANCESCO, D. Propertiesof the hyperpolarization-activated current in rat hippocampal CA1 pyrami-dal cells.J. Neurophysiol.69: 2129–2136, 1993.

MACCAFERRI, G. AND MCBAIN, C. J. The hyperpolarization-activated current (Ih)and its contribution to pacemaker activity in rat CA1 hippocampal stratumoriens-alveus interneurones.J. Physiol. (Lond.)497: 119–130, 1996.

MAGISTRETTI, J. AND ALONSO, A. Biophysical properties and slow voltage-dependent inactivation of a sustained sodium current in entorhinal cortexlayer II principal neurons. A whole-cell and single-channel study.J. Gen.Physiol.114: 491–509, 1999.

MAGISTRETTI, J., RAGSDALE, D. S., AND ALONSO, A. High conductance sus-tained single channel activity responsible for the low threshold persistentNa1 current in entorhinal cortex neurons.J. Neurosci.19: 7334–7341,1999.

MAYER, M. L. AND WESTBROOK, G. L. A voltage-clamp analysis of inward(anomalous) rectification in mouse spinal sensory ganglion neurones.J. Physiol. (Lond.)340: 19–45, 1983.

MCCORMICK, D. A. AND PAPE, H.-C. Properties of a hyperpolarization-acti-vated cation current and its role in rhythmic oscillation in thalamic relayneurones.J. Physiol. (Lond.)431: 291–318, 1990.

MERCURI, N. B., BONCI, A., CALABRESI, P., STEFANI, A., AND BERNARDI, G.Properties of the hyperpolarization-activated cation current Ih in rat mid-brain dopaminergic neurons.Eur. J. Neurosci.7: 462–469, 1995.

MITCHELL, S. J. AND RANCK, J.B.J. Generation of theta rhythm in medialentorhinal cortex of freely moving rats.Brain Res.189: 49–66, 1980.

NEHER, E. Correction of liquid junction potentials in patch clamp experiments.Methods Enzymol.207: 123–131, 1992.

NOMA, A. AND IRISAWA, H. A time- and voltage-dependent potassium currentin the rabbit sinoatrial node cell.Pflugers Arch.366: 251–258, 1976.

2578 DICKSON, MAGISTRETTI, SHALINSKY, FRANSE´ N, HASSELMO, AND ALONSO

PAPE, H.-C. Queer current and pacemaker: the hyperpolarization-activatedcation current in neurons.Annu. Rev. Physiol.58: 299–327, 1996.

PAPE, H.-C., PARE, D., AND DRIESANG, R. B. Two types of intrinsic oscillationsin neurons of the lateral and basolateral nuclei of the amygdala.J. Neuro-physiol.79: 205–216, 1998.

PAVLIDES, C., GREENSTEIN, Y., GRUDMAN, M., AND WINSON, J. Long-termpotentiation in the dentate gyrus is induced preferentially on the positivephase of theta-rhythm.Brain Res.439: 383–387, 1988.

RAMON Y CAJAL, S. Sobre un ganglio especial de la corteza esfeno-occipital.Trab. del Lab. de Invest. Biol. Univ. Madrid1: 189–201, 1902.

RUDELL, A. P., FOX, S. E.,AND RANCK, J. B. Hippocampal excitability phase-locked to the theta rhythm in walking rats.Exp. Neurol.68: 87–96, 1980.

SANTORO, B., LIU, D. T., YAO, H., BARTSCH, D., KANDEL, E. R., SIEGELBAUM,S. A., AND TIBBS, G. R. Identification of a gene encoding a hyperpolariza-tion-activated pacemaker channel in the brain.Cell 93: 717–729, 1998.

SINGER, W. Synchronization of cortical activity and its putative role in infor-mation processing and learning.Annu. Rev. Physiol.55: 349–374, 1993.

SOLOMON, J. S.AND NERBONNE, J. M. Two kinetically distinct components ofhyperpolarization-activated current in rat superior colliculus-projecting neu-rons.J. Physiol. (Lond.)469: 291–313, 1993.

SOLTESZ, I. AND DESCHENES, M. Low- and high-frequency membrane potentialoscillations during theta activity in CA1 and CA3 pyramidal neurons of therat hippocampus under ketamine-xylazine anesthesia.J. Neurophysiol.70:97–116, 1993.

SOLTESZ, I., LIGHTOWLER, S., LERESCHE, N., JASSIK-GERSCHENFELD, D., POL-LARD, C. E.,AND CRUNELLI, V. Two inward currents and the transformationof low-frequency oscillations of rat and cat thalamocortical cells.J. Physiol.(Lond.)441: 175–197, 1991.

SPAIN, W. J., SCHWINDT, P. C.,AND CRILL, W. E. Anomalous rectification inneurons from cat sensorimotor cortex in vitro.J. Neurophysiol.57: 1555–1576, 1987.

STERIADE, M., GLOOR, P., LLINAS, R. R., LOPES DASILVA , F. H.,AND MESULAM,M.-M. Basic mechanisms of cerebral rhythmic activities.EEG Clin. Neu-rophysiol.76: 481–508, 1990.

STERIADE, M. AND LLINAS, R. R. The functional states of the thalamus and theassociated neuronal interplay.Physiol. Rev.68: 649–742, 1988.

STERIADE, M., MCCORMICK, D. A., AND SEJNOWSKI, T. J. Thalamocorticaloscillations in the sleeping and aroused brain.Science262: 679 – 685,1993.

TAKAHASHI , T. Inward rectification in neonatal rat spinal motoneurones.J. Physiol. (Lond.)423: 57–62, 1990.

TRAVAGLI , R. A. AND GILLIS, R. A. Hyperpolarization-activated currents,Ih andIkir, in rat dorsal motor nucleus of the vagus neurons in vitro.J. Neuro-physiol.71: 1308–1317, 1994.

WANG, X.-J. Ionic basis for intrinsic 40 Hz neuronal oscillations.Neuroreport5: 221–224, 1993.

WHITE, J. A., BUDDE, T., AND KAY, A. R. A bifurcation analysis of neuronalsubthreshold oscillations.Biophys. J.69: 1203–1217, 1995.

WHITE, J. A., KLINK , R., ALONSO, A., AND KAY, A. R. Noise from voltage-gated ion channels may influence neuronal dynamics in the entorhinalcortex.J. Neurophysiol.80: 262–269, 1998.

WILLIAMS , S. R., TURNER, J. P., HUGHES, S. W., AND CRUNELLI, V. On thenature of anomalous rectification in thalamocortical neurones of the catventrobasal thalamus in vitro.J. Physiol. (Lond.)505: 727–747, 1997.

WINSON, J. Loss of hippocampal theta rhythm results in spatial memory deficitin the rat.Science201: 160–163, 1978.

WOUTERLOOD, P. G., HARTIG, W., BRUCKNER, G., AND WITTER, M. P. Parval-bumin-immunoreactive neurons in the entorhinal cortex of the rat: localiza-tion, morphology, connectivity and ultraestructure.J. Neurocytol.24: 135–153, 1995.

YANAGIHARA , K. AND IRISAWA, H. Inward current activated during hyperpo-larization in the rabbit sino atrial node cell.Pfugers Arch. Eur. J. Physiol.385: 11–19, 1980.

YLINEN, A., SOLTESZ, I., BRAGIN, A., PENTTONEN, M., SIK, A., AND BUZSAKI ,G. Intracellular correlates of hippocampal theta rhythm in identifiedpyramidal cells, granule cells, and basket cells.Hippocampus5: 78 –90,1995.

2579Ih IN EC LAYER II NEURONS