Comprehensive Physiology Wiley Online Library

Electrophysiology of the Cerebellar Networks

Full Article on Wiley Online Library



Abstract

The sections in this article are:

1 Comparative Aspects of Cerebellar Circuits
2 Field Potentials
2.1 Local Stimulation of Cerebellar Cortex
2.2 Field Potentials Generated by White Matter Stimulation of Cerebellar Cortex
2.3 Summary
3 Unitary Cell Recording from Cerebellar Cortex
3.1 Electrophysiology of Purkinje Cells
3.2 Activation of Climbing Fiber Afferents
3.3 Mossy Fiber Afferents
3.4 Inhibitory Systems in Cerebellar Cortex
3.5 Summary
4 Purkinje Cell Target Neurons, Cerebellar Nuclei, and Deiters' Nucleus
4.1 Summary
5 Inferior Olive
5.1 Intracellular Recording From Inferior Olivary Cells In Vivo and In Vitro
5.2 Ionic Mechanisms in Spike Generation
5.3 Inferior Olivary Cell as Single‐Cell Oscillator
5.4 Long‐Term Changes in Inferior Olive Neuronal Excitability
5.5 Electrotonic Coupling
5.6 Summary
6 Overall Function of Cerebellum
6.1 Braitenberg's Timing Model
6.2 Marr's Learning Model
6.3 Pellionisz‐Llinás Tensor Model
6.4 Boylls' Synergic Parameterization Theory
6.5 Summary
Figure 1. Figure 1.

Upper left: diagram of basic cerebellar circuit comprising mossy and climbing fiber inputs to Purkinje cells. Activity in mossy fiber (MF) is relayed through granule cells (GC) via parallel fibers (PF) to the Purkinje cell (PC). Axons of these cells are the only output system from cerebellar cortex. Second afferent is a climbing fiber (CF), which establishes a monosynaptic input to Purkinje cell dendrites. Upper right: diagram of the 2 basic inhibitory systems of cerebellar cortex. On left, basket cell axon contacts the somata of Purkinje cells. These axons also contact dendrites of these cells. Stellate cell (not shown) establishes direct contact with Purkinje cell dendrites. Both basket (BC) and stellate cells receive input via parallel fibers (PF) and constitute the inhibitory systems of the molecular layer. Golgi cell (GC to right) receives input from parallel fibers, mossy fibers (MF), and climbing fibers (not shown) and relays inhibition onto dendrites of granule cells (GrC) in granular layer. Bottom: detail of geometric organization of neuronal elements of cerebellar cortex. Drawing demonstrates different sections through a cerebellar folium. A: transverse plane. B: saggital plane. C: tangential plane. Cellular elements are displayed in drawings A, B, and C, as though the cerebellum were transparent. Orthogonal organization of parallel fibers, with respect to isoplanar characteristics of dendrites of Purkinje cells and basket and stellate cells (SC), is self‐explanatory. Note that axons of basket and stellate cells run at right angles with respect to parallel fibers and that dendritic tree of Golgi cells is close to cylindrical rather than isoplanar.

Upper left and upper right adapted from Ramón y Cajal 215; bottom from Llinás 138
Figure 2. Figure 2.

Field potentials and corresponding current densities generated by Loc stimulation of cat cerebellar cortex. A: potentials recorded at 4 parallel tracks are arranged in columns. Records were obtained at indicated depths. Lowest trace in each column is surface response. First column is accurately on beam of excited parallel fibers (0 μm). Other columns are recorded at lateralities of 200, 400, and 600 μm from the beam, respectively. Perpendicular broken lines indicate point of measurement of field‐potential amplitudes displayed in B. B: depth and transverse measurements of amplitudes of field potentials shown in A. The 7 tracks indicate potential measurements in arbitrary units at 5.3 ms latency for series in A and also for the 3 other tracks at 100, 300, and 500 μm. Approximately isopotential contour lines are drawn for values indicated by arrows. Note large lateral and deep spread of positive potential field in track 400 μm lateral to center of beam of excited parallel fibers (0 μm). C: field potentials obtained at depth indicated to left. D: corresponding current source‐density analysis, demonstrating localization of current. Note that positivity below 200 μm in C is not accompanied by large current flow at that level in D. E: surface recording of field potential generated by parallel fiber activation. F: superimposed current‐source densities in similar location before and after application of synaptic blocking agent manganese. This enables the pre‐ and postsynaptic components to be verified. Time scales are the same, recording position is just below cerebellar surface.

A and B adapted from Eccles, Llinás, and Sasaki 54; C and D from Nicholson and Llinás 189; E and F from Nicholson et al. 186
Figure 3. Figure 3.

Cerebellar field potentials recorded at different depths and evoked by local stimulation of surface of cerebellum in different vertebrates. A: frog field potentials evoked by parallel fiber activation at surface of cerebellum and recorded at different depths from cerebellar surface as indicated in microns at left of records. Depth is same for all records in a given horizontal plane. Note that in frog the late negativity is clearly reversed at 200 μm. B: similar type of records obtained from an alligator cerebellum at indicated depths. Note large negative potential that follows the parallel fiber compound action current (first negativity) from surface up to 300‐μm depth. C: records obtained from pigeon cerebellum as in A and B. Note that as for alligator, late negativity recorded at surface is present up to 300‐μm depth. D: field potential generated in cat cerebellum by parallel fiber volley. Late negativity recorded at surface of cortex (0 μm) reversed at 100 μm, reached maximum at 200 μm, and decreased progressively with depth. Time and voltage calibrations as indicated.

From Llinás et al. 147
Figure 4. Figure 4.

Laminar field potentials evoked by antidromic activation of Purkinje cells in frog, alligator, and cat. Records illustrated were generated by JF stimulation and recorded at depth indicated in microns to left of each series of traces. In all records, early negative response diminishes in amplitude, increases in latency, and quickly reverses as microelectrode is withdrawn. Difference in time courses between frog and alligator and that of cat probably reflect temperature differences.

Data for frog from Llinás et al. 144; data for alligator from Nicholson and Llinás 188; data for cats from Eccles, Llinás, and Sasaki 52
Figure 5. Figure 5.

Field potentials and current densities evoked by mossy fiber activation at different depths in cerebellar cortex. A: field potentials evoked by transfolial stimulation. A superficial negativity, the N3 wave, is first potential to be seen in depth. It corresponds to action potential in parallel fibers via mossy fiber‐granule cell relay. At 300‐ to 500‐μm depths an earlier negativity, the N2 wave, is split by a sharp positivity (marked by arrow in trace at 500 μm) attributed to activation of ascending axons of granule cells. B: current‐density analysis of mossy fiber response. Current densities were computed from laminar field potentials. Flows toward pial surface are indicated as positive. Principal excitatory components of current densities at time t1 are indicated by vertically hatched areas and peak current by vertical bar. Calibrations: 50 g·wt for tension; 0.95 μA/mm2 for molecular layer and 0.50 μm2/mm2 for granular layer current densities.

A adapted from Eccles, Llinás, and Sasaki 55; B adapted from Kwan and Murphy 130
Figure 6. Figure 6.

Field potentials produced in cat cerebellar cortex by contralateral stimulation of inferior olive. Field potential is characterized by initial negativity followed by positivity at granular and molecular layers and by superficial negativity at levels above 200 μm. Small difference in latency of early negativity at different depths reflects conduction time of climbing fiber action potentials in molecular layer.

Adapted from Eccles, Llinás, and Sasaki 51
Figure 7. Figure 7.

Ionic mechanisms for Purkinje cell firing. AF: recordings from mammalian Purkinje cell somata in vitro. AC: repetitive firing obtained with prolonged current pulses. In A, threshold current stimulus produces repetitive activation of Purkinje cell after initial local response (arrow). In B and C., increases in current injection amplitude produce high‐frequency firing and an oscillatory behavior marked with arrows in B. D–E: tetrodotoxin (TTX) sensitivity of Purkinje cell spikes. In D, control response to square pulse depolarization. In E, similar pulse after addition of TTX to bath. Note that fast spikes are blocked, whereas slower oscillations and afterdepolarization (arrow) remain. F: addition of cobalt chloride (Co) to TTX saline removes all electroresponsiveness.

Adapted from Llinás and Sugimori 154
Figure 8. Figure 8.

Dendritic recording from mammalian Purkinje cells in vitro. A: composite picture showing relationship between somatic and dendritic action potentials following D. C: depolarization through recording electrode. A clear shift in amplitude of fast action potentials and dendritic calcium‐dependent spikes is seen when comparing the more superficial recording in B with the somatic recording in E. Note that at increasing distances from the soma, fast spikes are reduced in amplitude and are barely noticeable in more peripheral recordings. Prolonged and slow‐rising burst spikes are more prominent at dendritic level, however. F: calcium‐dependent plateau and burst spikes. These plateau potentials seen intradendritically with short depolarizations are recorded in presence of tetrodotoxin (TTX). As stimulus is increased, prolonged local responses are observed which ultimately result in full dendritic spike bursts. G: addition of cadmium chloride (Cd) to TTX solution produces complete blockage of plateau and burst response recorded intradendritically.

Adapted from Llinás and Sugimori 155
Figure 9. Figure 9.

Direct stimulation of an in vitro guinea pig Purkinje cell after blockage of calcium conductance with cobalt chloride (Co). Arrows indicate onset of slow voltage‐dependent sodium conductance, which generates repetitive firing. Note that plateau level seems independent at level of current injection (lower traces).

Adapted from Llinás and Sugimori 154
Figure 10. Figure 10.

Climbing fiber activation of mammalian Purkinje cell in vitro. A: all‐or‐none Purkinje cell activation after JF stimulation. B: reversal of climbing fiber‐evoked synaptic potential. Properties of reversal are particularly clear at 18, 22, and 28 nA, where biphasic nature of reversal is clearly observed. C: voltage‐current relationship for synaptic potential. Bottom: computer display of somatic potentials after synaptic activation of the 3 different compartments in cable model. The 1st set of records to left illustrates EPSP at normal resting potential and its reversal as membrane is depolarized to +50 mV. The EPSP is generated by superposition of synaptic inputs to compartments 1, 2, and 3. Next set of records (labeled 1) is produced by synaptic input restricted to 1st compartment, whereas 2 and 3 are restricted to 2nd and 3rd compartments, respectively. Note that when synaptic input is restricted to compartment 1, EPSP reversal (at approximately −15 mV) is a mirror image of its depolarized counterpart. Potential generated by an input to compartment 2 reverses at approximately +40 mV. Input to compartment 3 does not reverse at all for same current level. When 1, 2, and 3 are activated (with progressive delay of 0.2 ms between compartments to allow for climbing fiber propagation time), reversal is biphasic as in experimental data. Current‐injection levels increase from zero at lowest level to 40 nA at uppermost level, in 5‐nA steps.

From Llinás and Nicholson 151
Figure 11. Figure 11.

Climbing fiber activation of Purkinje cells recorded extra‐ and intracellularly. AD: climbing fiber responses in frog Purkinje cells. In A, extracellular recording from Purkinje cell identified by its all‐or‐none antidromic activation from underlying white matter. As stimulus is increased, an all‐or‐none burst of 6 spikes is recorded (B), which has a very regular amplitude and time course (3 superimposed traces). In C and D, intracellular records from another frog Purkinje cell showing climbing fiber EPSP after stimulus to white matter. In C, all‐or‐none nature of EPSP is shown by electrical stimulation at threshold level. In D, 2 EPSPs are superimposed to show their regularity in latency and time course. Time and voltage calibration as indicated. EF: extracellular recordings from alligator Purkinje cell. As in frog, alligator Purkinje cells are activated antidromically from white matter (E). All‐or‐none climbing fiber burst of spikes is shown in E, generating a large action potential followed by 3 small spikes. The regularity of this response is illustrated in F, where 2 sweeps have been superimposed. GH: intracellular records from another Purkinje cell show all‐or‐none climbing fiber EPSP (G) and 2 superimposed with stimuli at suprathreshold level. Time and voltage calibrations as indicated. IJ: extracellular potentials from pigeon Purkinje cell after JF stimulation. In I, antidromic invasion of Purkinje cell. In J, all‐or‐none climbing fiber bursts of spikes. In K, intracellular record from same cell. After antidromic action potential, the climbing fiber activation generates a long‐lasting burst of spikes. Time and voltage calibration as indicated. LO: extracellular potentials from cat Purkinje cells. L shows all‐or‐none antidromic Purkinje cell spike. In O, all‐or‐none climbing fiber spike burst. Intracellular record from another Purkinje cell after activation of underlying white matter (N) shows antidromic action potential followed by large climbing fiber depolarization. In O, all‐or‐none climbing fiber EPSP generated by stimulation of contralateral olive. Time and voltage calibration as indicated.

From Llinás and Hillman 146
Figure 12. Figure 12.

Intracellular recordings from Purkinje cells illustrate synaptic potentials evoked by climbing fiber activation. In 1st column, climbing fiber EPSP is reversed by currect injection in an in vitro frog preparation. In 2nd column (alligator), climbing fiber EPSP is evoked by white matter stimulation. All‐or‐none nature of EPSP is illustrated in upper trace. Reversal of EPSP is observed with depolarizing (Dep) current pulses. Arrows on left indicate onset of pulse. Arrows on right indicate stimulus artifact. In 3rd column are superimposed records of climbing fiber EPSPs evoked in a cat Purkinje cell by JF stimulation with depolarizing and hyperpolarizing (Hyp) current pulses. Note that reflexly activated repetitive climbing fiber response of this cell (marked by arrows) was altered by applied current in same way as directly evoked EPSP.

Data for frog from Hackett 83; data for alligator from Llinás and Nicholson 149; data for cat from Eccles, Llinás, and Sasaki 51
Figure 13. Figure 13.

Patches of tactile projections to left paramedian lobe (PML) portrayed by both figurine (A) and patch mosiac (B) methods. A: figurine map collated from data gathered in 2 rats. Black patch on each figurine shows receptive field (RF) location, which, when stimulated, activates multiple units in granule cell layer at site shown by black dots in C. Note that projections to PML derive from entire body, but primarily from upper and lower lips. However, RFs from these mouth parts are smaller than those from forelimb, hindlimb, and trunk. Contralateral RFs are shown by black patches on right side of body. B: schematic outlines shown here illustrate patchlike character of projections from specific body structures or regions. This mosaic of patch projections was prepared by drawing a line around all adjacent figures shown in A with similar RF projections. For example, most lateral hand patch (h) is similar in shape and size with region in A occupied by the 11 hand figurines. Boundaries between adjacent patch projections seem to be discrete rather than continuous, revealing a fractured (or disjunctive) somatotopic organization. Specific body regions may project to more than 1 patch, showing multiple representation. For example, there are several lower lip patches. Most contralateral patch projections (encircled by dots in B) are found medially. Bilateral fields (encircled by solid lines) are more lateral. A tiny contralateral projection from upper lip was found just lateral to all ipsilateral projections. fl, Forelimb; h, hand; hl, hindlimb (includes trunk and hindlimb); li, lower incisor, ll, lower lip; ui, upper incisor; ul, upper lip; V, mystacial vibrissae. C: drawing of cerebellum indicating location of electrode punctures (black dots).

From Shambes et al. 224
Figure 14. Figure 14.

Proposed radial organization of ascending axon of granule cells. Granule cells (in circle) are assumed to contact Purkinje cells (hatched) not only via parallel fibers but also via ascending portion of their axons. Granule cell to left indicates possibility of a number of contacts 4 between 1 ascending granule cell axon and 1 Purkinje cell.

From Llinás 141
Figure 15. Figure 15.

Inhibitory postsynaptic potentials in cat Purkinje cells. A: intracellular recording obtained at depth of 360 μm with graded Loc stimulation at indicated strengths; a corresponding just‐extracellular recording is below each trace. Time and voltage calibrations as indicated. B: inhibitory synaptic potential inverted by intracellular chloride injection. IPSPs recorded in Purkinje cell at 300‐μm depth was inverted by chloride diffusion out of electrode. First trace shows inhibitory synaptic noise in a series of 3 superimposed traces. In 6 subsequent traces, stimulation was progressively increased, as shown by strengths on arbitrary scale. Note difference in time course between normal and reversed IPSPs.

Adapted from Eccles, Llinás, and Sasaki 56
Figure 16. Figure 16.

Extracellular recording of responses of presumed inhibitory interneurons. A: volley of impulses recorded at 350‐μm depth was generated in parallel fibers by a stimulus of progressively increasing strength (given in arbitrary units to left) applied through a surface stimulating electrode. B: 2 parallel fiber volleys at 180‐μm depth. Conditioning and testing local stimuli were kept constant and stimulus interval was varied. Control testing response (CON) had 4 spikes. LOC, parallel fiber.

From Eccles, Llinás, and Sasaki 53
Figure 17. Figure 17.

Golgi cell inhibition. AB: diagrams illustrating position of glomerulus in neuronal network of cerebellar cortex and an idealized version of its ultrastructure. A gives a quasi‐stereoscopic view of a small part of cerebellar folium into which mossy afferents (Mo) enter; their synaptic expansions are rosettes. Rosettes are connected mainly by short, claw‐shaped dendrites of granule cells (Gr) and by the descending dendrites of large Golgi neurons (black). Ascending axons of granule cells give rise to parallel fibers (Pf), which while running in longitudinal axis of folium, pierce flattened dendritic trees of Purkinje neurons (Pu, represented here as spade‐shaped boxes). Axon branches of Golgi neurons (Go ax) enter glomeruli and give rise to a plexus of small beaded terminals. B shows synaptic relations of these elements within glomerulus as seen with electron microscope. Mo, mossy afferent; GrD, granule dendrites entering through glial capsule (Gl) of glomerulus and terminating in characteristic bulbous terminals or digits; Dd, desmosomoid dendrodendritic contacts; GoD, descending dendrite of Golgi cell neuron with characteristic small spines, which makes broad contact with mossy rosette. Golgi axon terminals situated in periphery of glomerulus (hatched) establish synaptic contacts exclusively with granule cell dendrites. C illustrates electrode placement. DG: inhibition of repetitive firing of impulses by granule cells in response to Loc stimulation. In D, single granule cell was fired repetitively by single stimulus to superficial radial nerve (SR). In E, this response was inhibited by a reconditioning Loc stimulus that preceded SR stimulus at increasing intervals. In F, spontaneous activity of several granule cells recorded at 600‐μm depth, inhibited by local stimulation in G. H–K: effects of a preceding Loc stimulation on EPSPs in Purkinje cell evoked by transfolial (TF) and local stimulation. The EPSP evoked by TF stimulation (H) was depressed by a preceding Loc stimulation of increasing strengths (I). The EPSP evoked in the same cell by an Loc stimulus (J) was not depressed by a conditioning Loc stimulus but enhanced (K).

A and B from Szentágothai 238; DK from Eccles, Llinás, and Sasaki 55
Figure 18. Figure 18.

Stellate and basket cell inhibition of Purkinje cells in different vertebrates. In elasmobranch, intracellularly recorded EPSP‐IPSP sequence following local stimuli of increasing strength. Note ripples, indicated by dots, which appear to be unitary IPSPs. In frog, intracellular recordings from Purkinje cells after surface stimulation. Graded EPSP‐IPSP sequence is illustrated for increasing amplitudes of Loc stimulation. In alligator, 1st trace represents a threshold activation of parallel fibers. Stimulus strength is then increased from this level to 2.5 times threshold in the 5th trace. Last trace shows field potential recorded extracellularly in immediate vicinity of Purkinje cell. Arrow in upper trace indicates presence of spontaneous IPSP. In cat, similar set of records obtained by local stimulation in cerebellum. Recordings obtained immediately below beam of activated parallel fibers.

Data for elasmobranch from Nicholson, Llinás, and Precht 190; data for frog from Freeman and Lubozynski 70; data for alligator from Llinás and Nicholson 149; data for cat from Eccles, Llinás, and Sasaki 56
Figure 19. Figure 19.

Diagrammatic illustration of monosynaptic connections between cerebellar cortex and Deiters' neurons. In A, Deiters' nucleus (Deit), which generates vestibulospinal tract (VST), receives excitatory inputs from both mossy and climbing fiber collaterals. In addition it receives excitatory input from fastigial nucleus (F). Purkinje cells produce direct inhibition on Deiters' neurons. BE: intracellular recordings obtained from Deiters' nucleus after stimulation of cerebellar cortex (S) and recorded with a microelectrode (M). The IPSP in Deiters' neurons is produced by stimulation of ipsilateral anterior lobe of cerebellum. Stimulation was increased from 1.9 to 30 V to vermal cortex at lobule IV. Dotted lines in E indicate time course of potential changes if similar to that shown in D. F: IPSP shown at slower sweep speed after activation of lobule III. GH: suppression and rebound facilitation of spontaneous discharge induced by stimulation of lobule III. In H, there is absence of spontaneous firing.

Adapted from Ito and Yoshida 108
Figure 20. Figure 20.

Inhibitory action on Deiters' neurons via activation of Purkinje cell after spinal cord stimulation. A: extracellular and intracellular recording from Deiters' neurons to indicate (upper trace) antidromic invasion followed by 2 periods of excitation (arrows) and 2 silent periods in between. Lower trace, intracellular correlation of these excitability changes, consisting of an early and a late EPSP‐IPSP sequence. B: synaptic potential pattern evoked in Deiters' neurons by spinal cord stimulation at 2nd cervical vertebra. The 1st response is antidromic spike, which is followed by early excitation and inhibition. This response is followed (dot) by an excitatory potential and a fast inhibition produced by mossy fiber collateral activation of the nucleus and the inhibition via mossy fiber activation of Purkinje cells. The 2nd excitatory potential with a latency of 12 ms and the large inhibition and rebound excitation that ensues is due to activation of olivocerebellar pathway. C: early IPSP generated via mossy fiber Purkinje cell activation is quite stable, and late fast IPSP generated via climbing fiber activation of Purkinje cells shows discrete components. Its activation is very dependent on frequency. In D, discrete nature of this inhibition shows a clear latency shift for both early excitatory (dot) and fast inhibitory potential. EF: disappearance of late IPSP after lesion of inferior olive. E: control inhibition, showing early and late (dot) IPSP at 2 different sweep speeds. In F, same situation as E, but after transection of olivocerebellar tract by midline section. Note lack of postinhibitory rebound in slow sweep speed record in F. Calibrations: voltage 5 mV except for top portion of A, which is 0.2 mV; time 10 ms except for right portion of E, which is 50 ms.

A and B from Bruggencate et al. 31; CF from Bruggencate et al. 32
Figure 21. Figure 21.

Electrical excitability of mammalian inferior olivary cells tested in vitro. AC: normal electroresponsiveness. DF: electrophysiology after calcium blockage by extracellular cobalt (Co). GI: electrophysiological properties after blockage of sodium conductance by extracellular tetrodotoxin (TTX), indicating 2 different sets of calcium‐dependent action potentials. In B, subthreshold current pulse is given at resting membrane potential. In A, subthreshold stimulus is superimposed on a small DC depolarization (dotted line, with respect to solid line), generating a fast action potential followed by an afterdepolarization and a prolonged afterhyperpolarization. In C, hyperpolarization about 8 mV from rest (dotted line) also produces an increase in excitability as seen by action potential generated by otherwise subthreshold stimulus. Records A–C have been separated for clarity. In DF, there is a similar sequence as in AC but after blockage of calcium conductance by extracellular cobalt. Notice that in D, action potential lacks afterdepolarization and prolonged afterhyperpolarization seen in A. In F, subthreshold stimulus riding on hyperpolarization is now incapable of generating an action potential. In GI, sodium spike has been blocked by TTX. Again in H, subthreshold stimulus can generate an action potential by either depolarizing (G) or hyperpolarizing (I) membrane potential change. Although fast action potentials in A, C, and D are generated by sodium‐dependent conductances, the afterdepolarizations in A and C and in G and I are generated by calcium conductances. Those of A and G are generated by high‐threshold calcium‐dependent action potentials from dendrites. Those in C and I are generated by inactivating calcium conductances at somatic level.

From Llinás and Yarom 160
Figure 22. Figure 22.

Rhythmic firing of inferior olivary neurons. A: rebound calcium spike in presence of tetrodotoxin. Direct stimulation of inferior olivary neuron produces a dendritic calcium spike, followed by an afterhyperpolarization and a rebound spike (arrow). Small changes in DC hyperpolarizations (note current record) facilitate rebound spike, which becomes larger and moves to left. B: diagram illustrating sequence of events that generates inferior olivary cell rebound leading to oscillation. Antidromic or direct stimulation generates a somatic sodium spike having a fast rise and about 1‐ms duration (broken line). At appropriate membrane potential, this action potential generates a dendritic calcium spike that produces a plateau afterdepolarization followed by a sizable potassium conductance that generates prolonged afterhyperpolarization. This membrane hyperpolarization removes inactivation from somatic calcium conductance, which can then produce a rebound depolarization and can start the sequence once again.

Adapted from Llinás and Yarom 159
Figure 23. Figure 23.

Electrotonic coupling between inferior olivary (IO) neurons. A: 2 simultaneously recorded IO cells fired by antidromic stimulus. B: In a 2nd pair, hyperpolarization of lower cells through recording electrode hyperpolarizes 2nd cell (upper trace). To right: diagram of inferior olivary glomerulus. Top: general organization of IO glomerulus. In central core, dendritic branches are seen coupled by means of gap junctions (arrowheads). Central core is surrounded by synaptic terminals (ST), which establish contact with core elements. Bottom: on left, path of coupling current between 2 IO neurons and, on right, hypothetical function for synaptic junction at glomerulus. When synapses are activated, conductance change produced by synaptic transmitter action on postsynaptic membrane produces a shunt at glomerular level that reduces coupling coefficient between cells, since current tends to be lost across shunt.

A and B from Llinás and Yarom 159; diagrams on right from Llinás 139
Figure 24. Figure 24.

Limb movement as tensorial entity. A: on left, an upward displacement vector is a physical entity that can be expressed in different reference frames: e.g., by x,y coordinate system, or by α,β,γ‐ordered set of 3 quantities. On right, 2 reference frames shown are of fundamentally different kinds: applies to CNS‐independent external space; α,β,γ applies to space inherently connected to CNS. Limb‐displacement vector occurs in both spaces. Different expressions of the vector are related by limb‐displacement tensor, . BD: covariant analysis and contravariant synthesis via a metric tensor. B: given a 2‐dimensional intended vector and 3 α,β,γ‐axes of an overcomplete reference frame, the decomposition could be performed by a 2‐step operation. First, covariant components of can be established (C), using geometry of 2‐space, to any number of directions independently. (Perpendicular projections, i.e., the inner products, provide “features” of desired vector in any coordinate direction.) Physical sum of covariant components, however, is not equal to displacement. Second, provided that metric tensor is available (in contravariant expression) for the α,β,γ‐space, corresponding set of contravariant components can be established (D). Physical sum of contravariant components physically generates displacement vector .

Adapted from Pellionisz and Llinás 209
Figure 25. Figure 25.

Circuit exemplifying certain mechanisms characteristic of Boylls' synergic parameterization theory. This circuit represents only 1 of a number of alternative realizations consistent with existing physiological knowledge. PF, parallel fibers; MF, mossy fibers; CF, climbing fibers; g, granule cells.



Figure 1.

Upper left: diagram of basic cerebellar circuit comprising mossy and climbing fiber inputs to Purkinje cells. Activity in mossy fiber (MF) is relayed through granule cells (GC) via parallel fibers (PF) to the Purkinje cell (PC). Axons of these cells are the only output system from cerebellar cortex. Second afferent is a climbing fiber (CF), which establishes a monosynaptic input to Purkinje cell dendrites. Upper right: diagram of the 2 basic inhibitory systems of cerebellar cortex. On left, basket cell axon contacts the somata of Purkinje cells. These axons also contact dendrites of these cells. Stellate cell (not shown) establishes direct contact with Purkinje cell dendrites. Both basket (BC) and stellate cells receive input via parallel fibers (PF) and constitute the inhibitory systems of the molecular layer. Golgi cell (GC to right) receives input from parallel fibers, mossy fibers (MF), and climbing fibers (not shown) and relays inhibition onto dendrites of granule cells (GrC) in granular layer. Bottom: detail of geometric organization of neuronal elements of cerebellar cortex. Drawing demonstrates different sections through a cerebellar folium. A: transverse plane. B: saggital plane. C: tangential plane. Cellular elements are displayed in drawings A, B, and C, as though the cerebellum were transparent. Orthogonal organization of parallel fibers, with respect to isoplanar characteristics of dendrites of Purkinje cells and basket and stellate cells (SC), is self‐explanatory. Note that axons of basket and stellate cells run at right angles with respect to parallel fibers and that dendritic tree of Golgi cells is close to cylindrical rather than isoplanar.

Upper left and upper right adapted from Ramón y Cajal 215; bottom from Llinás 138


Figure 2.

Field potentials and corresponding current densities generated by Loc stimulation of cat cerebellar cortex. A: potentials recorded at 4 parallel tracks are arranged in columns. Records were obtained at indicated depths. Lowest trace in each column is surface response. First column is accurately on beam of excited parallel fibers (0 μm). Other columns are recorded at lateralities of 200, 400, and 600 μm from the beam, respectively. Perpendicular broken lines indicate point of measurement of field‐potential amplitudes displayed in B. B: depth and transverse measurements of amplitudes of field potentials shown in A. The 7 tracks indicate potential measurements in arbitrary units at 5.3 ms latency for series in A and also for the 3 other tracks at 100, 300, and 500 μm. Approximately isopotential contour lines are drawn for values indicated by arrows. Note large lateral and deep spread of positive potential field in track 400 μm lateral to center of beam of excited parallel fibers (0 μm). C: field potentials obtained at depth indicated to left. D: corresponding current source‐density analysis, demonstrating localization of current. Note that positivity below 200 μm in C is not accompanied by large current flow at that level in D. E: surface recording of field potential generated by parallel fiber activation. F: superimposed current‐source densities in similar location before and after application of synaptic blocking agent manganese. This enables the pre‐ and postsynaptic components to be verified. Time scales are the same, recording position is just below cerebellar surface.

A and B adapted from Eccles, Llinás, and Sasaki 54; C and D from Nicholson and Llinás 189; E and F from Nicholson et al. 186


Figure 3.

Cerebellar field potentials recorded at different depths and evoked by local stimulation of surface of cerebellum in different vertebrates. A: frog field potentials evoked by parallel fiber activation at surface of cerebellum and recorded at different depths from cerebellar surface as indicated in microns at left of records. Depth is same for all records in a given horizontal plane. Note that in frog the late negativity is clearly reversed at 200 μm. B: similar type of records obtained from an alligator cerebellum at indicated depths. Note large negative potential that follows the parallel fiber compound action current (first negativity) from surface up to 300‐μm depth. C: records obtained from pigeon cerebellum as in A and B. Note that as for alligator, late negativity recorded at surface is present up to 300‐μm depth. D: field potential generated in cat cerebellum by parallel fiber volley. Late negativity recorded at surface of cortex (0 μm) reversed at 100 μm, reached maximum at 200 μm, and decreased progressively with depth. Time and voltage calibrations as indicated.

From Llinás et al. 147


Figure 4.

Laminar field potentials evoked by antidromic activation of Purkinje cells in frog, alligator, and cat. Records illustrated were generated by JF stimulation and recorded at depth indicated in microns to left of each series of traces. In all records, early negative response diminishes in amplitude, increases in latency, and quickly reverses as microelectrode is withdrawn. Difference in time courses between frog and alligator and that of cat probably reflect temperature differences.

Data for frog from Llinás et al. 144; data for alligator from Nicholson and Llinás 188; data for cats from Eccles, Llinás, and Sasaki 52


Figure 5.

Field potentials and current densities evoked by mossy fiber activation at different depths in cerebellar cortex. A: field potentials evoked by transfolial stimulation. A superficial negativity, the N3 wave, is first potential to be seen in depth. It corresponds to action potential in parallel fibers via mossy fiber‐granule cell relay. At 300‐ to 500‐μm depths an earlier negativity, the N2 wave, is split by a sharp positivity (marked by arrow in trace at 500 μm) attributed to activation of ascending axons of granule cells. B: current‐density analysis of mossy fiber response. Current densities were computed from laminar field potentials. Flows toward pial surface are indicated as positive. Principal excitatory components of current densities at time t1 are indicated by vertically hatched areas and peak current by vertical bar. Calibrations: 50 g·wt for tension; 0.95 μA/mm2 for molecular layer and 0.50 μm2/mm2 for granular layer current densities.

A adapted from Eccles, Llinás, and Sasaki 55; B adapted from Kwan and Murphy 130


Figure 6.

Field potentials produced in cat cerebellar cortex by contralateral stimulation of inferior olive. Field potential is characterized by initial negativity followed by positivity at granular and molecular layers and by superficial negativity at levels above 200 μm. Small difference in latency of early negativity at different depths reflects conduction time of climbing fiber action potentials in molecular layer.

Adapted from Eccles, Llinás, and Sasaki 51


Figure 7.

Ionic mechanisms for Purkinje cell firing. AF: recordings from mammalian Purkinje cell somata in vitro. AC: repetitive firing obtained with prolonged current pulses. In A, threshold current stimulus produces repetitive activation of Purkinje cell after initial local response (arrow). In B and C., increases in current injection amplitude produce high‐frequency firing and an oscillatory behavior marked with arrows in B. D–E: tetrodotoxin (TTX) sensitivity of Purkinje cell spikes. In D, control response to square pulse depolarization. In E, similar pulse after addition of TTX to bath. Note that fast spikes are blocked, whereas slower oscillations and afterdepolarization (arrow) remain. F: addition of cobalt chloride (Co) to TTX saline removes all electroresponsiveness.

Adapted from Llinás and Sugimori 154


Figure 8.

Dendritic recording from mammalian Purkinje cells in vitro. A: composite picture showing relationship between somatic and dendritic action potentials following D. C: depolarization through recording electrode. A clear shift in amplitude of fast action potentials and dendritic calcium‐dependent spikes is seen when comparing the more superficial recording in B with the somatic recording in E. Note that at increasing distances from the soma, fast spikes are reduced in amplitude and are barely noticeable in more peripheral recordings. Prolonged and slow‐rising burst spikes are more prominent at dendritic level, however. F: calcium‐dependent plateau and burst spikes. These plateau potentials seen intradendritically with short depolarizations are recorded in presence of tetrodotoxin (TTX). As stimulus is increased, prolonged local responses are observed which ultimately result in full dendritic spike bursts. G: addition of cadmium chloride (Cd) to TTX solution produces complete blockage of plateau and burst response recorded intradendritically.

Adapted from Llinás and Sugimori 155


Figure 9.

Direct stimulation of an in vitro guinea pig Purkinje cell after blockage of calcium conductance with cobalt chloride (Co). Arrows indicate onset of slow voltage‐dependent sodium conductance, which generates repetitive firing. Note that plateau level seems independent at level of current injection (lower traces).

Adapted from Llinás and Sugimori 154


Figure 10.

Climbing fiber activation of mammalian Purkinje cell in vitro. A: all‐or‐none Purkinje cell activation after JF stimulation. B: reversal of climbing fiber‐evoked synaptic potential. Properties of reversal are particularly clear at 18, 22, and 28 nA, where biphasic nature of reversal is clearly observed. C: voltage‐current relationship for synaptic potential. Bottom: computer display of somatic potentials after synaptic activation of the 3 different compartments in cable model. The 1st set of records to left illustrates EPSP at normal resting potential and its reversal as membrane is depolarized to +50 mV. The EPSP is generated by superposition of synaptic inputs to compartments 1, 2, and 3. Next set of records (labeled 1) is produced by synaptic input restricted to 1st compartment, whereas 2 and 3 are restricted to 2nd and 3rd compartments, respectively. Note that when synaptic input is restricted to compartment 1, EPSP reversal (at approximately −15 mV) is a mirror image of its depolarized counterpart. Potential generated by an input to compartment 2 reverses at approximately +40 mV. Input to compartment 3 does not reverse at all for same current level. When 1, 2, and 3 are activated (with progressive delay of 0.2 ms between compartments to allow for climbing fiber propagation time), reversal is biphasic as in experimental data. Current‐injection levels increase from zero at lowest level to 40 nA at uppermost level, in 5‐nA steps.

From Llinás and Nicholson 151


Figure 11.

Climbing fiber activation of Purkinje cells recorded extra‐ and intracellularly. AD: climbing fiber responses in frog Purkinje cells. In A, extracellular recording from Purkinje cell identified by its all‐or‐none antidromic activation from underlying white matter. As stimulus is increased, an all‐or‐none burst of 6 spikes is recorded (B), which has a very regular amplitude and time course (3 superimposed traces). In C and D, intracellular records from another frog Purkinje cell showing climbing fiber EPSP after stimulus to white matter. In C, all‐or‐none nature of EPSP is shown by electrical stimulation at threshold level. In D, 2 EPSPs are superimposed to show their regularity in latency and time course. Time and voltage calibration as indicated. EF: extracellular recordings from alligator Purkinje cell. As in frog, alligator Purkinje cells are activated antidromically from white matter (E). All‐or‐none climbing fiber burst of spikes is shown in E, generating a large action potential followed by 3 small spikes. The regularity of this response is illustrated in F, where 2 sweeps have been superimposed. GH: intracellular records from another Purkinje cell show all‐or‐none climbing fiber EPSP (G) and 2 superimposed with stimuli at suprathreshold level. Time and voltage calibrations as indicated. IJ: extracellular potentials from pigeon Purkinje cell after JF stimulation. In I, antidromic invasion of Purkinje cell. In J, all‐or‐none climbing fiber bursts of spikes. In K, intracellular record from same cell. After antidromic action potential, the climbing fiber activation generates a long‐lasting burst of spikes. Time and voltage calibration as indicated. LO: extracellular potentials from cat Purkinje cells. L shows all‐or‐none antidromic Purkinje cell spike. In O, all‐or‐none climbing fiber spike burst. Intracellular record from another Purkinje cell after activation of underlying white matter (N) shows antidromic action potential followed by large climbing fiber depolarization. In O, all‐or‐none climbing fiber EPSP generated by stimulation of contralateral olive. Time and voltage calibration as indicated.

From Llinás and Hillman 146


Figure 12.

Intracellular recordings from Purkinje cells illustrate synaptic potentials evoked by climbing fiber activation. In 1st column, climbing fiber EPSP is reversed by currect injection in an in vitro frog preparation. In 2nd column (alligator), climbing fiber EPSP is evoked by white matter stimulation. All‐or‐none nature of EPSP is illustrated in upper trace. Reversal of EPSP is observed with depolarizing (Dep) current pulses. Arrows on left indicate onset of pulse. Arrows on right indicate stimulus artifact. In 3rd column are superimposed records of climbing fiber EPSPs evoked in a cat Purkinje cell by JF stimulation with depolarizing and hyperpolarizing (Hyp) current pulses. Note that reflexly activated repetitive climbing fiber response of this cell (marked by arrows) was altered by applied current in same way as directly evoked EPSP.

Data for frog from Hackett 83; data for alligator from Llinás and Nicholson 149; data for cat from Eccles, Llinás, and Sasaki 51


Figure 13.

Patches of tactile projections to left paramedian lobe (PML) portrayed by both figurine (A) and patch mosiac (B) methods. A: figurine map collated from data gathered in 2 rats. Black patch on each figurine shows receptive field (RF) location, which, when stimulated, activates multiple units in granule cell layer at site shown by black dots in C. Note that projections to PML derive from entire body, but primarily from upper and lower lips. However, RFs from these mouth parts are smaller than those from forelimb, hindlimb, and trunk. Contralateral RFs are shown by black patches on right side of body. B: schematic outlines shown here illustrate patchlike character of projections from specific body structures or regions. This mosaic of patch projections was prepared by drawing a line around all adjacent figures shown in A with similar RF projections. For example, most lateral hand patch (h) is similar in shape and size with region in A occupied by the 11 hand figurines. Boundaries between adjacent patch projections seem to be discrete rather than continuous, revealing a fractured (or disjunctive) somatotopic organization. Specific body regions may project to more than 1 patch, showing multiple representation. For example, there are several lower lip patches. Most contralateral patch projections (encircled by dots in B) are found medially. Bilateral fields (encircled by solid lines) are more lateral. A tiny contralateral projection from upper lip was found just lateral to all ipsilateral projections. fl, Forelimb; h, hand; hl, hindlimb (includes trunk and hindlimb); li, lower incisor, ll, lower lip; ui, upper incisor; ul, upper lip; V, mystacial vibrissae. C: drawing of cerebellum indicating location of electrode punctures (black dots).

From Shambes et al. 224


Figure 14.

Proposed radial organization of ascending axon of granule cells. Granule cells (in circle) are assumed to contact Purkinje cells (hatched) not only via parallel fibers but also via ascending portion of their axons. Granule cell to left indicates possibility of a number of contacts 4 between 1 ascending granule cell axon and 1 Purkinje cell.

From Llinás 141


Figure 15.

Inhibitory postsynaptic potentials in cat Purkinje cells. A: intracellular recording obtained at depth of 360 μm with graded Loc stimulation at indicated strengths; a corresponding just‐extracellular recording is below each trace. Time and voltage calibrations as indicated. B: inhibitory synaptic potential inverted by intracellular chloride injection. IPSPs recorded in Purkinje cell at 300‐μm depth was inverted by chloride diffusion out of electrode. First trace shows inhibitory synaptic noise in a series of 3 superimposed traces. In 6 subsequent traces, stimulation was progressively increased, as shown by strengths on arbitrary scale. Note difference in time course between normal and reversed IPSPs.

Adapted from Eccles, Llinás, and Sasaki 56


Figure 16.

Extracellular recording of responses of presumed inhibitory interneurons. A: volley of impulses recorded at 350‐μm depth was generated in parallel fibers by a stimulus of progressively increasing strength (given in arbitrary units to left) applied through a surface stimulating electrode. B: 2 parallel fiber volleys at 180‐μm depth. Conditioning and testing local stimuli were kept constant and stimulus interval was varied. Control testing response (CON) had 4 spikes. LOC, parallel fiber.

From Eccles, Llinás, and Sasaki 53


Figure 17.

Golgi cell inhibition. AB: diagrams illustrating position of glomerulus in neuronal network of cerebellar cortex and an idealized version of its ultrastructure. A gives a quasi‐stereoscopic view of a small part of cerebellar folium into which mossy afferents (Mo) enter; their synaptic expansions are rosettes. Rosettes are connected mainly by short, claw‐shaped dendrites of granule cells (Gr) and by the descending dendrites of large Golgi neurons (black). Ascending axons of granule cells give rise to parallel fibers (Pf), which while running in longitudinal axis of folium, pierce flattened dendritic trees of Purkinje neurons (Pu, represented here as spade‐shaped boxes). Axon branches of Golgi neurons (Go ax) enter glomeruli and give rise to a plexus of small beaded terminals. B shows synaptic relations of these elements within glomerulus as seen with electron microscope. Mo, mossy afferent; GrD, granule dendrites entering through glial capsule (Gl) of glomerulus and terminating in characteristic bulbous terminals or digits; Dd, desmosomoid dendrodendritic contacts; GoD, descending dendrite of Golgi cell neuron with characteristic small spines, which makes broad contact with mossy rosette. Golgi axon terminals situated in periphery of glomerulus (hatched) establish synaptic contacts exclusively with granule cell dendrites. C illustrates electrode placement. DG: inhibition of repetitive firing of impulses by granule cells in response to Loc stimulation. In D, single granule cell was fired repetitively by single stimulus to superficial radial nerve (SR). In E, this response was inhibited by a reconditioning Loc stimulus that preceded SR stimulus at increasing intervals. In F, spontaneous activity of several granule cells recorded at 600‐μm depth, inhibited by local stimulation in G. H–K: effects of a preceding Loc stimulation on EPSPs in Purkinje cell evoked by transfolial (TF) and local stimulation. The EPSP evoked by TF stimulation (H) was depressed by a preceding Loc stimulation of increasing strengths (I). The EPSP evoked in the same cell by an Loc stimulus (J) was not depressed by a conditioning Loc stimulus but enhanced (K).

A and B from Szentágothai 238; DK from Eccles, Llinás, and Sasaki 55


Figure 18.

Stellate and basket cell inhibition of Purkinje cells in different vertebrates. In elasmobranch, intracellularly recorded EPSP‐IPSP sequence following local stimuli of increasing strength. Note ripples, indicated by dots, which appear to be unitary IPSPs. In frog, intracellular recordings from Purkinje cells after surface stimulation. Graded EPSP‐IPSP sequence is illustrated for increasing amplitudes of Loc stimulation. In alligator, 1st trace represents a threshold activation of parallel fibers. Stimulus strength is then increased from this level to 2.5 times threshold in the 5th trace. Last trace shows field potential recorded extracellularly in immediate vicinity of Purkinje cell. Arrow in upper trace indicates presence of spontaneous IPSP. In cat, similar set of records obtained by local stimulation in cerebellum. Recordings obtained immediately below beam of activated parallel fibers.

Data for elasmobranch from Nicholson, Llinás, and Precht 190; data for frog from Freeman and Lubozynski 70; data for alligator from Llinás and Nicholson 149; data for cat from Eccles, Llinás, and Sasaki 56


Figure 19.

Diagrammatic illustration of monosynaptic connections between cerebellar cortex and Deiters' neurons. In A, Deiters' nucleus (Deit), which generates vestibulospinal tract (VST), receives excitatory inputs from both mossy and climbing fiber collaterals. In addition it receives excitatory input from fastigial nucleus (F). Purkinje cells produce direct inhibition on Deiters' neurons. BE: intracellular recordings obtained from Deiters' nucleus after stimulation of cerebellar cortex (S) and recorded with a microelectrode (M). The IPSP in Deiters' neurons is produced by stimulation of ipsilateral anterior lobe of cerebellum. Stimulation was increased from 1.9 to 30 V to vermal cortex at lobule IV. Dotted lines in E indicate time course of potential changes if similar to that shown in D. F: IPSP shown at slower sweep speed after activation of lobule III. GH: suppression and rebound facilitation of spontaneous discharge induced by stimulation of lobule III. In H, there is absence of spontaneous firing.

Adapted from Ito and Yoshida 108


Figure 20.

Inhibitory action on Deiters' neurons via activation of Purkinje cell after spinal cord stimulation. A: extracellular and intracellular recording from Deiters' neurons to indicate (upper trace) antidromic invasion followed by 2 periods of excitation (arrows) and 2 silent periods in between. Lower trace, intracellular correlation of these excitability changes, consisting of an early and a late EPSP‐IPSP sequence. B: synaptic potential pattern evoked in Deiters' neurons by spinal cord stimulation at 2nd cervical vertebra. The 1st response is antidromic spike, which is followed by early excitation and inhibition. This response is followed (dot) by an excitatory potential and a fast inhibition produced by mossy fiber collateral activation of the nucleus and the inhibition via mossy fiber activation of Purkinje cells. The 2nd excitatory potential with a latency of 12 ms and the large inhibition and rebound excitation that ensues is due to activation of olivocerebellar pathway. C: early IPSP generated via mossy fiber Purkinje cell activation is quite stable, and late fast IPSP generated via climbing fiber activation of Purkinje cells shows discrete components. Its activation is very dependent on frequency. In D, discrete nature of this inhibition shows a clear latency shift for both early excitatory (dot) and fast inhibitory potential. EF: disappearance of late IPSP after lesion of inferior olive. E: control inhibition, showing early and late (dot) IPSP at 2 different sweep speeds. In F, same situation as E, but after transection of olivocerebellar tract by midline section. Note lack of postinhibitory rebound in slow sweep speed record in F. Calibrations: voltage 5 mV except for top portion of A, which is 0.2 mV; time 10 ms except for right portion of E, which is 50 ms.

A and B from Bruggencate et al. 31; CF from Bruggencate et al. 32


Figure 21.

Electrical excitability of mammalian inferior olivary cells tested in vitro. AC: normal electroresponsiveness. DF: electrophysiology after calcium blockage by extracellular cobalt (Co). GI: electrophysiological properties after blockage of sodium conductance by extracellular tetrodotoxin (TTX), indicating 2 different sets of calcium‐dependent action potentials. In B, subthreshold current pulse is given at resting membrane potential. In A, subthreshold stimulus is superimposed on a small DC depolarization (dotted line, with respect to solid line), generating a fast action potential followed by an afterdepolarization and a prolonged afterhyperpolarization. In C, hyperpolarization about 8 mV from rest (dotted line) also produces an increase in excitability as seen by action potential generated by otherwise subthreshold stimulus. Records A–C have been separated for clarity. In DF, there is a similar sequence as in AC but after blockage of calcium conductance by extracellular cobalt. Notice that in D, action potential lacks afterdepolarization and prolonged afterhyperpolarization seen in A. In F, subthreshold stimulus riding on hyperpolarization is now incapable of generating an action potential. In GI, sodium spike has been blocked by TTX. Again in H, subthreshold stimulus can generate an action potential by either depolarizing (G) or hyperpolarizing (I) membrane potential change. Although fast action potentials in A, C, and D are generated by sodium‐dependent conductances, the afterdepolarizations in A and C and in G and I are generated by calcium conductances. Those of A and G are generated by high‐threshold calcium‐dependent action potentials from dendrites. Those in C and I are generated by inactivating calcium conductances at somatic level.

From Llinás and Yarom 160


Figure 22.

Rhythmic firing of inferior olivary neurons. A: rebound calcium spike in presence of tetrodotoxin. Direct stimulation of inferior olivary neuron produces a dendritic calcium spike, followed by an afterhyperpolarization and a rebound spike (arrow). Small changes in DC hyperpolarizations (note current record) facilitate rebound spike, which becomes larger and moves to left. B: diagram illustrating sequence of events that generates inferior olivary cell rebound leading to oscillation. Antidromic or direct stimulation generates a somatic sodium spike having a fast rise and about 1‐ms duration (broken line). At appropriate membrane potential, this action potential generates a dendritic calcium spike that produces a plateau afterdepolarization followed by a sizable potassium conductance that generates prolonged afterhyperpolarization. This membrane hyperpolarization removes inactivation from somatic calcium conductance, which can then produce a rebound depolarization and can start the sequence once again.

Adapted from Llinás and Yarom 159


Figure 23.

Electrotonic coupling between inferior olivary (IO) neurons. A: 2 simultaneously recorded IO cells fired by antidromic stimulus. B: In a 2nd pair, hyperpolarization of lower cells through recording electrode hyperpolarizes 2nd cell (upper trace). To right: diagram of inferior olivary glomerulus. Top: general organization of IO glomerulus. In central core, dendritic branches are seen coupled by means of gap junctions (arrowheads). Central core is surrounded by synaptic terminals (ST), which establish contact with core elements. Bottom: on left, path of coupling current between 2 IO neurons and, on right, hypothetical function for synaptic junction at glomerulus. When synapses are activated, conductance change produced by synaptic transmitter action on postsynaptic membrane produces a shunt at glomerular level that reduces coupling coefficient between cells, since current tends to be lost across shunt.

A and B from Llinás and Yarom 159; diagrams on right from Llinás 139


Figure 24.

Limb movement as tensorial entity. A: on left, an upward displacement vector is a physical entity that can be expressed in different reference frames: e.g., by x,y coordinate system, or by α,β,γ‐ordered set of 3 quantities. On right, 2 reference frames shown are of fundamentally different kinds: applies to CNS‐independent external space; α,β,γ applies to space inherently connected to CNS. Limb‐displacement vector occurs in both spaces. Different expressions of the vector are related by limb‐displacement tensor, . BD: covariant analysis and contravariant synthesis via a metric tensor. B: given a 2‐dimensional intended vector and 3 α,β,γ‐axes of an overcomplete reference frame, the decomposition could be performed by a 2‐step operation. First, covariant components of can be established (C), using geometry of 2‐space, to any number of directions independently. (Perpendicular projections, i.e., the inner products, provide “features” of desired vector in any coordinate direction.) Physical sum of covariant components, however, is not equal to displacement. Second, provided that metric tensor is available (in contravariant expression) for the α,β,γ‐space, corresponding set of contravariant components can be established (D). Physical sum of contravariant components physically generates displacement vector .

Adapted from Pellionisz and Llinás 209


Figure 25.

Circuit exemplifying certain mechanisms characteristic of Boylls' synergic parameterization theory. This circuit represents only 1 of a number of alternative realizations consistent with existing physiological knowledge. PF, parallel fibers; MF, mossy fibers; CF, climbing fibers; g, granule cells.

References
 1. Adrian, E. D. Discharge frequencies in the cerebral and cerebellar cortex. J. Physiol. London 83: 32–33, 1935.
 2. Albus, J. S. A theory of cerebellar function. Math. Biosci. 10: 25–61, 1971.
 3. Altman, J., W. J. Anderson, and K. A. Wright. Selective destruction of precursors of microneurons of the cerebellar cortex with fractionated low‐dose x‐rays. Exp. Neurol. 17: 481–497, 1967.
 4. Andersen, P., J. C. Eccles, and P. E. Voorhoeve. Postsynaptic inhibition of cerebellar Purkinje cells. J. Neurophysiol. 27: 1138–1153, 1964.
 5. Armstrong, D. M., R. J. Harvey, and R. F. Schild. Branching of inferior olivary axons to terminate in different folia, lobules or lobes of the cerebellum. Brain Res. 54: 365–371, 1973.
 6. Baker, P. F., A. L. Hodgkin, and E. B. Ridgway. Depolarization and calcium entry in squid giant axons. J. Physiol. London 218: 709–755, 1971.
 7. Bantli, H. Analysis of difference between potentials evoked by climbing fibers in cerebellum of cat and turtle. J. Neurophysiol. 37: 573–593, 1974.
 8. Belbenoit, P. Conditionnement instrumental de l'électroperception des objets chez Gnathonemus petersii (Mormyridae, Teleostei, Pisces). Z. Vgl. Physiol. 67: 192–204, 1970.
 9. Belcari, P., A. Francesconi, C. Maioli, and P. Strata. Spontaneous activity of the Purkinje cells in the pigeon cerebellum. Pfluegers Arch. 371: 147–154, 1977.
 10. Bell, C. C., and R. J. Grimm. Discharge properties of Purkinje cells recorded on single and double microelectrodes. J. Neurophysiol. 32: 1044–1055, 1969.
 11. Bennett, M. V. L. Physiology of electrotonic junctions. Ann. NY Acad. Sci. 137: 509–539, 1966.
 12. Bernstein, N. The Coordination and Regulation of Movements. New York: Pergamon, 1967.
 13. Bisti, S., G. Iosip, G. F. Marchesi, and P. Strata. Pharmacological properties of inhibition in the cerebellar cortex. Exp. Brain Res. 14: 24–37, 1971.
 14. Bloedel, J. R., R. S. Gregory, and S. H. Martin. Action of interneurons and axon collaterals in cerebellar cortex of a primate. J. Neurophysiol. 35: 847–863, 1972.
 15. Bloedel, J. R., and W. J. Roberts. Action of climbing fibers in cerebellar cortex of the cat. J. Neurophysiol. 34: 17–31, 1971.
 16. Blomfield, S., and D. Marr. How the cerebellum may be used. Nature London 227: 1224–1228, 1970.
 17. Bloom, F. E., Chemical integrative processes in the central nervous system. In: The Neurosciences: Fourth Study Program, edited by F. O. Schmitt and F. G. Worden. Cambridge: MIT Press, 1979, p. 51–58.
 18. Bower, J. M., D. C. Woolston, and J. M. Gibson. Congruence of spatial patterns of receptive field projections to Purkinje cell and granule cell layers in the cerebellar cortex of the rat. Soc. Neurosci. Abstr. 6: 511, 1980.
 19. Boylls, C. C. A Theory of Cerebellar Function With Applications to Locomotion. I. The Physiological Role of Climbing Fiber Inputs in Anterior Lobe Operation. Amherst: Univ. of Massachusetts, 1975. (COINS Tech. Rep. 75C‐6.).
 20. Boylls, C. C. A Theory of Cerebellar Function With Applications to Locomotion. II. The Relation of Anterior Lobe Climbing Fiber Function to Locomotor Behavior in the Cat. Amherst: Univ. of Massachusetts, 1975. (COINS Tech. Rep. 76–1.).
 21. Boylls, C. C., Cerebellar strategies for movement coordination. In: Tutorials in Motor Behavior, edited by G. E. Stelmach and J. Reguin. New York: Elsevier/North Holland, 1980, p. 83–94.
 22. Boylls, C. C., Contributions to locomotor coordination of an olivo‐cerebellar projection to the vermis in the cat: experimental results and theoretical proposals. In: The Inferior Olivary Nucleus: Anatomy and Physiology, edited by J. Courville, C. de Montigny, and Y. Lamarre. New York: Raven, 1980, p. 321–348.
 23. Braitenberg, V. Functional interpretation of cerebellar histology. Nature London. 190: 539–540, 1961.
 24. Braitenberg, V., Is the cerebellar cortex a biological clock in the millisecond range? In: Progress in Brain Research. The Cerebellum, edited by C. A. Fox and R. S. Snider. Amsterdam: Elsevier, 1967, vol. 25, p. 334–346.
 25. Braitenberg, V., and R. P. Atwood. Morphological observations in the cerebellar cortex. J. Comp. Neurol. 109: 1–34, 1958.
 26. Braitenberg, V., and N. Onesto. The cerebellar cortex as a timing organ. Discussion of an hypothesis. Proc. 1st Congr. Int. Med. Cibern. Naples: Giannini, 1961, p. 1–19.
 27. Brand, S., A.‐L. Dahl, and E. Mugnaini. The length of parallel fibers in the cat cerebellar cortex. An experimental light and electron microscopic study. Exp. Brain Res. 26: 39–58, 1976.
 28. Brindley, G. S. The use made by the cerebellum of the information that it receives from sense organs. Int. Brain Res. Org. Bull. 3: 80, 1964.
 29. Brodal, A., and B. Høivik. Site and termination of primary vestibulocerebellar fibers in the cat: an experimental study with silver impregnation methods. Arch. Ital. Biol. 102: 1–21, 1964.
 30. Brookhart, J. M., G. Moruzzi, and R. S. Snider. Spike discharges of single units in the cerebellar cortex. J. Neurophysiol. 13: 465–486, 1950.
 31. Bruggencate, G. ten, R. Teichmann, and E. Weller. Neuronal activity in the lateral vestibular nucleus of the cat. I. Patterns of postsynaptic potentials and discharges in Deiters' neurones evoked by stimulation of the spinal cord. Pfluegers Arch. 337: 119–134, 1972.
 32. ten Bruggencate, G., R. Teichmann, and E. Weller. Neuronal activity in the lateral vestibular nucleus of the cat. III. Inhibitory actions of cerebellar Purkinje cells evoked via mossy and climbing fibre afferents. Pfluegers Arch. 337: 147–162, 1972.
 33. Buisseret, P., and C. Batini. Réponses complexes des cellules de Purkinje: conduction avec leur caractère répétitif. C. R. Acad. Sci. Ser. D 273: 2306–2308, 1971.
 34. Burke, R. E., and P. Rudomín. Spinal neurons and synapses. In: Handbook of Physiology. The Nervous System, edited by J. M. Brookhart and V. B. Mountcastle. Bethesda, MD: Am. Physiol. Soc., 1977, sect. 1, vol. I, pt. 2, chapt. 24, p. 877–944.
 35. Buser, P., and A. Rougeul. Etude des réponses du cervelet du pigeon à la stimulation du nerf optique. Boll. Soc. Ital. Biol. Sper. 30: 758–760, 1954.
 36. Clarke, P. G. H. The organization of visual processing in the pigeon cerebellum. J. Physiol. London 243: 267–284, 1974.
 37. Colin, F., J. Manil, and J. C. Desclin. The olivocerebellar system. I. Delayed and slow inhibitory effects: an overlooked salient feature of cerebellar climbing fibers. Brain Res. 187: 3–27, 1980.
 38. Coombs, J. S., J. C. Eccles, and P. Fatt. The electrical properties of the motoneurone membrane. J. Physiol. London 130: 291–325, 1955.
 39. Courville, J., J. R. Augustine, and P. Martel. Projections from the inferior olive to the cerebellar nuclei in the cat demonstrated by retrograde transport of horseradish peroxidase. Brain Res. 130: 405–419, 1977.
 40. Crepel, F., and J. Mariani. Multiple innervation of Purkinje cells by climbing fibers in the cerebellum of the Weaver mutant mouse. J. Neurobiol. 1: 579–582, 1976.
 41. Crill, W. E. Unitary multiple‐spiked responses in cat inferior olive nucleus. J. Neurophysiol. 33: 199–209, 1970.
 42. Desclin, J. C., and F. Colin. The olivocerebellar system. II. Some ultrastructural correlates of inferior olive destruction in the rat. Brain Res. 187: 29–46, 1980.
 43. Dow, R. S. The electrical activity of the cerebellum and its functional significance. J. Physiol London 94: 67–86, 1938.
 44. Dow, R. S. Cerebellar action potentials in response to stimulation of various afferent connections. J. Neurophysiol. 2: 543–555, 1939.
 45. Dow, R. S. Action potentials of cerebellar cortex in response to local electrical stimulation. J. Neurophysiol. 12: 245–256, 1949.
 46. Eccles, J. C. Circuits in the cerebellar control of movement. Proc. Natl. Acad. Sci. USA 58: 336–343, 1967.
 47. Eccles, J. C., The dynamic loop hypothesis of movement control. In: Information Processing in the Central Nervous System, edited by K. N. Leibovic. Berlin: Springer‐Verlag, 1969, p. 245–269.
 48. Eccles, J. C. The Understanding of the Brain. New York: McGraw‐Hill, 1973.
 49. Eccles, J. C. The cerebellum as a computer: pattern in space and time. J. Physiol. London 229: 1–32, 1973.
 50. Eccles, J. C., M. Ito, and J. Szentágothai. The Cerebellum as a Neuronal Machine. Berlin: Springer‐Verlag, 1967.
 51. Eccles, J. C., R. Llinás, and K. Sasaki. The excitatory synaptic action of climbing fibres on the Purkinje cells of the cerebellum. J. Physiol. London 182: 268–296, 1966.
 52. Eccles, J. C., R. Llinás, and K. Sasaki. The action of antidromic impulses on the cerebellar Purkinje cells. J. Physiol. London 182: 316–345, 1966.
 53. Eccles, J. C., R. Llinás, and K. Sasaki. The inhibitory interneurons within the cerebellar cortex. Exp. Brain Res. 1: 1–16, 1966.
 54. Eccles, J. C., R. Llinás, and K. Sasaki. Parallel fibre stimulation and the responses induced thereby in the Purkinje cells of the cerebellum Exp. Brain Res. 1: 17–39, 1966.
 55. Eccles, J. C., R. Llinás, and K. Sasaki. The mossy fibre‐granule cell relay of the cerebellum and its inhibitory control by Golgi cells. Exp. Brain Res. 1: 82–101, 1966.
 56. Eccles, J. C., R. Llinás, and K. Sasaki. Intracellularly recorded responses of the cerebellar Purkinje cells. Exp. Brain Res. 1: 161–183, 1966.
 57. Eccles, J. C., R. Llinás, K. Sasaki, and P. E. Voorhoeve. Interaction experiments on the responses evoked in Purkinje cells by climbing fibres. J. Physiol. London 182: 297–315, 1966.
 58. Eccles, J. C., K. Sasaki, and P. Strata. The profiles of physiological events produced by a parallel fiber volley in the cerebellar cortex. Exp. Brain Res. 2: 18–34, 1966.
 59. Eccles, J. C., K. Sasaki, and P. Strata. Interpretation of the potential fields generated in the cerebellar cortex by a mossy fiber volley. Exp. Brain Res. 3: 58–80, 1967.
 60. Eccles, J. C., K. Sasaki, and P. Strata. A comparison of the inhibitory actions of Golgi cells and of basket cells. Exp. Brain Res. 3: 81–94, 1967.
 61. Eccles, J. C., H. Táboříá, and N. Tsukahara. Excitation and inhibition of Purkyne cells in the cerebellum of Mustelus canis. Biol. Bull. 135: 418, 1968.
 62. Enger, P. S., and T. Szabo. Activity of central neurons involved in electroreception in some weakly electric fish (Gymnotidae). J. Neurophysiol. 28: 800–818, 1965.
 63. Faber, D. S., and H. Korn. Inhibition in the frog cerebellar cortex following parallel fiber activation. Brain Res. 17: 506–510, 1970.
 64. Fatt, P. Electric potentials occurring around a neurone during its antidromic activation. J. Neurophysiol. 20: 27–60, 1957.
 65. Fisher, D. S., and A. M. Jonas. Cerebellar hypoplasia resulting from cytosine arabinoside treatment in the neonatal hamster. Clin. Res. 13: 540: 1965.
 66. Flourens, P. Recherches experimentales sur les proprietes et les systèmes nerveux dans les animaux vertébrés. Paris: Crevot, 1824.
 67. Frank, K., and M. G. F. Fuortes. Potentials recorded from the spinal cord with microelectrodes. J. Physiol. London 130: 625–654, 1955.
 68. Frederickson, R. C. A., M. Neuss, S. L. Morzorati, and W. J. McBride. A comparison of the inhibitory effects of taurine and GABA on identified Purkinje cells and other neurons in the cerebellar cortex of the rat. Brain Res. 145: 117–126, 1978.
 69. Freeman, J. A., The cerebellum as a timing device: an experimental study in the frog. In: Neurobiology of Cerebellar Evolution and Development, edited by R. Llinás. Chicago: Am. Med. Assoc., 1969, p. 397–420.
 70. Freeman, J. A., and M. Lubozynski. Modulation of Purkinje cell firing by laterally‐mediated inhibition in the cerebellum of Bufo marinus. Federation Proc. 32: 365, 1973.
 71. Freeman, J. A., and C. Nicholson. Experimental optimization of current source‐density technique for anuran cerebellum. J. Neurophysiol. 38: 369–382, 1975.
 72. Freeman, J. A., and J. Stone. A technique for current density analysis of field potentials and its application to the frog cerebellum. In: Neurobiology of Cerebellar Evolution and Development, edited by R. Llinás. Chicago: Am. Med. Assoc., 1969, p. 421–430.
 73. Furukawa, T., and E. J. Furshpan. Two inhibitory mechanisms in the Mauthner neurons of goldfish. J. Neurophysiol. 26: 140–176, 1963.
 74. Gardner‐Medwin, A. R. An extreme supernormal period in cerebellar parallel fibres. J. Physiol. London 222: 357–371, 1972.
 75. Gilbert, P. F. C. A theory of memory that explains the function and structure of the cerebellum. Brain Res. 70: 1–18, 1974.
 76. Gilbert, P. F. C. How the cerebellum could memorise movements. Nature London 254: 688–689, 1975.
 77. Gilbert, P. F. C., and W. T. Thach. Purkinje cell activity during motor learning. Brain Res. 128: 309–328, 1977.
 78. Gould, B. B., and A. M. Graybiel. Afferents to the cerebellar cortex in the cat. Evidence for an intrinsic pathway leading from the deep nuclei to the cortex. Brain Res. 110: 601–611, 1976.
 79. Granit, R., and C. G. Phillips. Excitatory and inhibitory processes acting upon individual Purkinje cells of the cerebellum in cats. J. Physiol. London 133: 520–547, 1956.
 80. Granit, R., and C. G. Phillips. Effect on Purkinje cells of surface stimulation of the cerebellum. J. Physiol. London 135: 73–92, 1957.
 81. Gross, N. Sensory representation within the cerebellum of the pigeon. Brain Res. 21: 280–283, 1970.
 82. Guselnikov, V. I., and V. I. Ivanova. Electrical responses of cerebellum to the effects of various stimuli in lower vertebrates. Fiziol. Zh. SSSR im. I. M. Sechenova 44: 118–125, 1958.
 83. Hackett, J. T. Calcium dependency of excitatory chemical synaptic transmission in the frog cerebellum in vitro. Brain Res. 114: 35–46, 1976.
 84. Hagiwara, S. Calcium spikes. Adv. Biophys. 4: 71–102, 1973.
 85. Hámori, J., and J. Szentágothai. Participation of Golgi neuron processes in the cerebellar glomeruli: an electron microscope study. Exp. Brain Res. 2: 35–48, 1966.
 86. Haugedé‐carré, F., T. Szabo, and F. Kirschbaum. Development of the gigantocerebellum of the weakly electric fish pollimyrus. J. Physiol. Paris 75: 381–395, 1979.
 87. Heier, P. Fundamental principles in the structure of the brain; a study of the brain of Petromyzon fluviatilis. Acta Anat. Suppl. 8: 1–112, 1948.
 88. Herndon, R. M., G. Margolis, and L. Kilham. Virus‐induced cerebellar malformation. An electron microscopic study. J. Neuropathol. Exp. Neurol. 28: 164, 1969.
 89. Herndon, R. M., G. Margolis, and L. Kilham. Synaptic organization of the malformed cerebellum induced by perinatal infection with feline panleukopenia virus (PLV). II. The Purkinje cell and its afferents. J. Neuropathol. Exp. Neurol. 30: 557–560, 1971.
 90. Herrick, C. J. Origin and evolution of the cerebellum. Arch. Neurol. Psychiatry Chicago 11: 621–652, 1924.
 91. Hess, R., and J. I. Simpson. Visual and somatosensory messages to the rabbit's cerebellar flocculus. Neurosci. Lett. Suppl. 1: 146, 1978.
 92. Hillman, D. E. Morphological organization of frog cerebellar cortex: a light and electron microscopic study. J. Neurophysiol. 32: 818–846, 1969.
 93. Hillman, D. E., Neuronal organization of the cerebellar cortex in amphibia and reptilia. In: Neurobiology of Cerebellar Evolution and Development, edited by R. Llinás. Chicago: Am. Med. Assoc, 1969, p. 279–325.
 94. Hodgkin, A. L. The local electric changes associated with repetitive action in a nonmedulated axon. J. Physiol. London 107: 165–181, 1948.
 95. Hoffer, B. J., G. R. Siggins, and F. E. Bloom. Studies on norepinephrine‐containing afferents to Purkinje cells of rat cerebellum. II. Sensitivity of Purkinje cells to norepinephrine and related substances administered by microiontophoresis. Brain Res. 25: 523–534, 1971.
 96. Holmes, G. The cerebellum in man. Brain 63: 1–30, 1939.
 97. Hounsgaard, J., and C. Yamamoto. Dendritic spikes in Purkinje cells of the guinea pig cerebellum studied in vitro. Exp. Brain Res. 37: 387–398, 1979.
 98. Ito, M. Neurophysiological aspects of the cerebellar motor control system. Int. J. Neurol. 7: 162–176, 1970.
 99. Ito, M. Neural design of the cerebellar motor control system. Brain Res. 40: 81–84, 1972.
 100. Ito, M., The control mechanism of cerebellar motor systems. In: The Neurosciences: Third Study Program, edited by F. O. Schmitt and F. G. Worden. Cambridge: MIT Press, 1974, p. 293–303.
 101. Ito, M., Adaptive modification of the vestibulo‐ocular reflex in rabbits affected by visual inputs and its possible neuronal mechanisms. In: Progress in Brain Research. Reflex Control of Posture and Movement, edited by R. Granit and O. Pompeiano. Amsterdam: Elsevier, 1979, vol. 50, p. 757–761.
 102. Ito, M., Eye movements and the cerebellum. In: The Cerebellum: New Vistas, edited by S. L. Palay and V. Chan‐Palay. Heidelberg: Springer‐Verlag. In press.
 103. Ito, M., N. Kawai, and M. Udo. The origin of cerebellar‐induced inhibition of Deiters' neurons. III. Distribution of the inhibitory zone. Exp. Brain Res. 4: 310–320, 1968.
 104. Ito, M., N. Nisimaru, and K. Shibuki. Destruction of inferior olive induces rapid depression in synaptic action of cerebellar Purkinje cells. Nature London 227: 568–569, 1979.
 105. Ito, M., K. Obata, and R. Ochi. The origin of cerebellar‐evoked inhibition of Deiters' neurons. II. Temporal correlation between the trans‐synaptic activation of Purkinje cells and the inhibition of Deiters' neurons. Exp. Brain Res. 2: 350–364, 1966.
 106. Ito, M., I. Olov, and I. Shimoyama. Reduction of the cerebellar stimulus effect on rat Deiters' neurons after chemical destruction of the inferior olive. Exp. Brain Res. 33: 143–145, 1978.
 107. Ito, M., and J. I. Simpson. Discharges in Purkinje cell axons during climbing fiber activation. Brain Res. 31: 215–219, 1971.
 108. Ito, M., and M. Yoshida. The origin of cerebellar‐induced inhibition of Deiters' neurons. I. Monosynaptic initiation of the inhibitory postsynaptic potentials. Exp. Brain Res. 2: 330–349, 1966.
 109. Ito, M., M. Yoshida, K. Obata, N. Kawai, and M. Udo. Inhibitory control of intracerebellar nuclei by the Purkinje cell axons. Exp. Brain Res. 10: 64–80, 1970.
 110. Jansen, J. K. S. Afferent impulses to the cerebellar hemispheres from the cerebral cortex and certain subcortical nuclei. Acta Physiol. Scand. Suppl. 143: 1–99, 1957.
 111. Johnston, J. B. The brain of Petromyzon. J. Comp. Neurol. 12: 1–106, 1902.
 112. Johnston, J. B. The Nervous System of Vertebrates. Philadelphia: Blakiston, 1909.
 113. Kaiserman‐Abramof, I. R., and S. L. Palay. Fine structural studies of the cerebellar cortex in mormyrid fish. In: Neurobiology of Cerebellar Evolution and Development, edited by R. Llinás. Chicago: Am. Med. Assoc., 1969, p. 171–205.
 114. Kandel, E. R., and W. A. Spencer. Electrophysiological properties of an archicortical neuron. Ann. NY Acad. Sci. 94: 570–603, 1961.
 115. Karamian, A. I., V. V. Fanardjian, and A. A. Kosareva. The functional and morphological evolution of the cerebellum and its role in behavior. In: Neurobiology of Cerebellar Evolution and Development, edited by R. Llinás. Chicago: Am. Med. Assoc., 1969, p. 639–673.
 116. Kawamura, H., and L. Provini. Depression of cerebellar Purkinje cells by microiontophoretic application of GABA and related amino acids. Brain Res. 24: 293–304, 1979.
 117. Kennedy, D. T., T. Shimono, and S. T. Kitai. Parallel fiber and white matter activation of Purkinje cells in a reptilian cerebellum (Lacerta viridis). Brain Res. 22: 381–385, 1970.
 118. Kidokoro, Y. Direct inhibitory innervation of teleost oculomotor neurons by cerebellar Purkinje cells. Brain Res. 10: 453–456, 1968.
 119. Kidokoro, Y., Cerebellar and vestibular control of fish oculomotor neurones. In: Neurobiology of Cerebellar Evolution and Development, edited by R. Llinás. Chicago: Am. Med. Assoc., 1969, p. 257–276.
 120. King, J. S., J. A. Andrezik, W. M. Falls, and G. F. Martin. Synaptic organization of cerebello‐olivary circuit. Exp. Brain Res. 26: 159–170, 1976.
 121. Kitai, S. T., T. Shimono, and D. T. Kennedy. Inhibition in the cerebellar cortex of the lizard, Lacerta viridis. In: Neurobiology of Cerebellar Evolution and Development, edited by R. Llinás. Chicago: Am. Med. Assoc., 1969, p. 481–489.
 122. Korn, H., and H. Axelrod. Electrical inhibition of Purkinje cells in the cerebellum of the rat. Proc. Natl. Acad. Sci. USA 77: 6244–6247, 1980.
 123. Kornhuber, H. H. Motor functions of cerebellum and basal ganglia: the cerebello‐cortical saccadic (ballistic) clock, the cerebello‐nuclear hold regulator, and the basal ganglia ramp (voluntary speed smooth movement) generator. Kybernetik 8: 157–162, 1971.
 124. Kostyuk, P. G., and O. A. Krishtal. Separation of sodium and calcium currents in the somatic membrane of mollusc neurones. J. Physiol. London 270: 545–568, 1977.
 125. Kotchabhakdi, N. Functional circuitry of the goldfish cerebellum. J. Comp. Physiol. 112: 47–73, 1976.
 126. Kotchabhakdi, N. Functional organization of the goldfish cerebellum. J. Comp. Physiol. 112: 75–93, 1976.
 127. Kron, G. Tensor Analysis of Networks. New York: Wiley, 1939.
 128. Kwan, H. C., and J. T. Murphy. A basis for extracellular current density analysis in cerebellar cortex. J. Neurophysiol. 37: 170–180, 1974.
 129. Kwan, H. C., and J. T. Murphy. Extracellular current density analysis of responses in cerebellar cortex to climbing fiber activation. J. Neurophysiol. 37: 333–345, 1974.
 130. Kwan, H. C., and J. T. Murphy. Extracellular current density analysis of responses in cerebellar cortex to mossy fiber activation. J. Neurophysiol. 37: 947–953, 1974.
 131. Lamarre, Y., C. de Montigny, M. Dumont, and M. Weiss. Harmaline‐induced rhythmic activity of cerebellar and lower brain stem neurons. Brain Res. 32: 246–250, 1971.
 132. Larramendi, L. M. H., and T. Victor. Synapses on spines of the Purkinje cell of the mouse. An electron microscopic study. Brain Res. 5: 15–30, 1967.
 133. Larsell, O. The cerebellum of myxinoids and petromyzonts, including developmental stages in the lampreys. J. Comp. Neurol. 86: 395–446, 1947.
 134. Larsell, O. Comparative Anatomy and Histology of the Cerebellum from Myxinoids through Birds, edited by J. Jansen. Minneapolis: Univ. of Minnesota Press, 1967.
 135. Latham, A., and D. H. Paul. Spontaneous activity of cerebellar Purkinje cells and their responses to impulses in climbing fibres. J. Physiol. London 213: 135–156, 1971.
 136. Libouban, S., and T. Szabo. An integration centre of the mormyrid fish brain: the auricula cerebelli. An HRP study. Neurosci. Lett. 6: 115–119, 1977.
 137. Llinás, R. Mechanisms of supraspinal actions upon spinal cord activities. Differences between reticular and cerebellar inhibitory actions upon alpha extensor motoneurons. J. Neurophysiol. 27: 1117–1126, 1964.
 138. Llinás, R., Neuronal operations in cerebellar transactions. In: The Neurosciences: Second Study Program, edited by F. O. Schmitt. New York: Rockefeller Univ. Press, 1970, p. 409–426.
 139. Llinás, R. 18th Bowditch Lecture: motor aspects of cerebellar control. Physiologist 17: 19–46, 1974.
 140. Llinás, R., The role of calcium in neuronal function. In: The Neurosciences: Fourth Study Program, edited by F. O. Schmitt and F. G. Worden. Cambridge: MIT Press, 1979, p. 555–571.
 141. Llinás, R., Radial connectivity in the cerebellar cortex: a novel view regarding the functional organization of the molecular layer. In: The Cerebellum: New Vistas, edited by S. L. Palay and V. Chan‐Palay. Heidelberg: Springer‐Verlag. In press.
 142. Llinás, R., R. Baker, and C. Sotelo. Electrotonic coupling between neurons in cat inferior olive. J. Neurophysiol. 37: 560–571, 1974.
 143. Llinás, R., J. R. Bloedel, and D. E. Hillman. Functional characterization of neuronal circuitry of frog cerebellar cortex. J. Neurophysiol. 32: 847–870, 1969.
 144. Llinás, R., J. R. Bloedel, and W. Roberts. Antidromic invasion of Purkinje cells in frog cerebellum. J. Neurophysiol. 32: 881–891, 1969.
 145. Llinás, R., and R. Hess. Tetrodotoxin‐resistant dendritic spikes in avian Purkinje cells. Proc. Natl. Acad. Sci. USA 73: 2520–2523, 1976.
 146. Llinás, R., and D. E. Hillman. Physiological and morphological organization of the cerebellar circuits of various vertebrates. In: Neurobiology of Cerebellar Evolution and Development, edited by R. Llinás. Chicago: Am. Med. Assoc., 1969, p. 43–73.
 147. Llinás, R., D. E. Hillman, and W. Precht. Functional aspects of cerebellar evolution. In: The Cerebellum in Health and Disease, edited by W. S. Fields and W. D. Willis, Jr. St. Louis: Green, 1970, p. 269–291.
 148. Llinás, R., D. E. Hillman, and W. Precht. Neuronal circuit reorganization in mammalian agranular cerebellar cortex. J. Neurobiol. 4: 69–94, 1973.
 149. Llinás, R., and C. Nicholson. Electrophysiological analysis of alligator cerebellar cortex: a study on dendritic spikes. In: Neurobiology of Cerebellar Evolution and Development, edited by R. Llinás. Chicago: Am. Med. Assoc., 1969, p. 431–465.
 150. Llinás, R., and C. Nicholson. Electrophysiological properties of dendrites and somata in alligator Purkinje cells. J. Neurophysiol. 34: 532–551, 1971.
 151. Llinás, R., and C. Nicholson. Reversal properties of climbing fiber potential in cat Purkinje cells: an example of a distributed synapse. J. Neurophysiol. 39: 311–323, 1976.
 152. Llinás, R., and J. I. Simpson. Cerebellar control of movement. In: Handbook of Behavioral Neurobiology, edited by A. Towe and E. Luschei. New York: Plenum, vol. II. In press.
 153. Llinás, R., and M. Sugimori. Calcium conductances in Purkinje cell dendrites: their role in development and integration. In: Progress in Brain Research, Development and Chemical Specificity of Neurons, edited by M. Cuénod. Amsterdam: Elsevier, 1979, vol. 51, p. 323–334.
 154. Llinás, R., and M. Sugimori. Electrophysiological properties of in vitro Purkinje cell somata in mammalian cerebellar slices. J. Physiol. London 305: 171–195, 1980.
 155. Llinás, R., and M. Sugimori. Electrophysiological properties of in vitro Purkinje cell dendrites in mammalian cerebellar slices. J. Physiol. London 305: 197–213, 1980.
 156. Llinás, R., and M. Sugimori. Functional significance of the climbing fiber input to Purkinje cells: an in vitro study in mammalian slices. In: The Cerebellum: New Vistas, edited by S. L. Palay and V. Chan‐Palay. Heidelberg: Springer‐Verlag. In press.
 157. Llinás, R., and R. A. Volkind. The olivo‐cerebellar system: functional properties as revealed by harmaline‐induced tremor. Exp. Brain Res. 18: 69–87, 1973.
 158. Llinás, R., Y. Yarom, and M. Sugimori. The isolated mammalian brain in vitro: a new technique for the analysis of the electrical activity of neuronal circuit function. Federation Proc. in press.
 159. Llinás, R., and Y. Yarom. Electrophysiological properties of mammalian inferior olivary cells in vitro. In: The Inferior Olivary Nucleus: Anatomy and Physiology, edited by J. Courville, C. de Montigny, and Y. Lamarre. New York: Raven, 1980, p. 379–388.
 160. Llinás, R., and Y. Yarom. Electrophysiology of mammalian inferior olivary neurons in vitro. Different types of voltage‐dependent ionic conductances. J. Physiol. London in press.
 161. Llinás, R., and Y. Yarom. Properties and distribution of ionic conductances generating electroresponsiveness of inferior olivary neurons in vitro. J. Physiol. London in press.
 162. Lorente de Nó, R. Action potential of the motoneurones of the hypoglossus nucleus. J. Cell Comp. Physiol. 29: 207–288, 1947.
 163. Maekawa, K., and T. Takeda. Mossy fiber responses evoked in the cerebellar flocculus of rabbits by stimulation of the optic pathway. Brain Res. 98: 590–595, 1975.
 164. Magendie, F. Précis élémentaire de physiologic Paris: Meguiguon‐Marvis, 1825, vols. I and II.
 165. Maler, L., H. J. Karten, and M. V. L. Bennett. The central connections of the posterior lateral line nerve of Gnathonemus petersii. J. Comp. Neurol. 151: 57–66, 1973.
 166. Maler, L., H. J. Karten, and M. V. L. Bennett. The central connections of the anterior lateral line nerve of Gnathonemus petersii. J. Comp. Neurol. 151: 67–84, 1973.
 167. Mano, N. Simple and complex spike activities of the cerebellar Purkinje cell in relation to selective alternate movements in intact monkey. Brain Res. 70: 381–393, 1974.
 168. Marr, D. A theory of cerebellar cortex. J. Physiol. London 202: 437–470, 1969.
 169. Marshall, K. C., J. M. Wojtowicz, and W. J. Hendelman. Patterns of functional synaptic connections in organized cultures of cerebellum. Neuroscience 5: 1847–1857, 1980.
 170. Mason, S. T., and S. D. Iverson. An investigation of the role of cortical and cerebellar noradrenaline in associative motor learning in the rat. Brain Res. 134: 513–527, 1977.
 171. Matsushita, M., and Y. Hosoya. The location of spinal projection neurons in the cerebellar nuclei (cerebellospinal tract neurons) of the cat: a study with the horseradish peroxidase technique. Brain Res. 142: 237–248, 1978.
 172. Matsushita, M., and M. Ikeda. Olivary projections to the cerebellar nuclei in the cat. Exp. Brain Res. 10: 488–500, 1970.
 173. Matthews, P. B. C., C. G. Phillips, and G. Rushworth. Afferent systems converging upon cerebellar Purkinje cells in the frog. Q. J. Exp. Physiol. 43: 38–52, 1958.
 174. McBride, W. J., and R. C. A. Frederickson. Taurine as a possible inhibitory transmitter in the cerebellum. Federation Proc. 39: 2701–2705, 1980.
 175. McCarley, R. W., and J. A. Hobson. Simple spike firing patterns of cat cerebellar Purkinje cells in sleep and waking. Electroencephalogr. Clin. Neurophysiol. 33: 471–483, 1972.
 176. de Montigny, C., and Y. Lamarre. Activity in the olivocerebello‐bulbar system of the cat during ibogaline‐ and oxotremorine‐induced tremor. Brain Res. 82: 369–373, 1974.
 177. Moruzzi, G. Effects at different frequencies of cerebellar stimulation upon postural tonus and myotatic reflexes. Electroencephalogr. Clin. Neurophysiol. 2: 463–469, 1950.
 178. Mugnaini, E., Ultrastructural studies on the cerebellar histogenesis. II. Maturation of nerve cell populations and establishment of synaptic connections in the cerebellar cortex of the chick. In: Neurobiology of Cerebellar Evolution and Development, edited by R. Llinás. Chicago: Am. Med. Assoc., 1969, p. 749–782.
 179. Mugnaini, E., The histology and cytology of the cerebellar cortex. In: The Comparative Anatomy and Histology of the Cerebellum: Human Cerebellum, Cerebellar Connections, and Cerebellar Cortex, edited by O. Larsell and J. Jansen. Minneapolis: Univ. of Minnesota Press, 1972, p. 201–262.
 180. Mugnaini, E., R. L. Atluri, and J. C. Hour. Fine structure of granular layer in turtle cerebellum with emphasis on large glomeruli. J. Neurophysiol. 37: 1–29, 1974.
 181. Mugnaini, E., and A. L. Dahl. Mode of distribution of aminergic fibers in the cerebellar cortex of the chicken. J. Comp. Neurol. 162: 417–432, 1975.
 182. Murphy, J. T., and N. H. Sabah. Spontaneous firing of cerebellar Purkinje cells in decerebrate and barbiturate anesthetized cats. Brain Res. 17: 515–519, 1970.
 183. Nacimiento, A. C., Spontaneous and evoked discharges of cerebellar Purkinje cells in the frog. In: Neurobiology of Cerebellar Evolution and Development, edited by R. Llinás. Chicago: Am. Med. Assoc., 1969, p. 373–395.
 184. Narahashi, T., J. W. Moore, and W. R. Scott. Tetrodotoxin blockage of sodium conductance increase in lobster giant axons. J. Gen. Physiol. 47: 965–974, 1964.
 185. Nicholson, C., G. ten Bruggencate, R. Steinberg, and H. Stoeckle. Calcium modulation in brain extracellular microenvironment demonstrated with ion‐selective micropipette. Proc. Natl. Acad. Sci. USA 74: 1287–1290, 1977.
 186. Nicholson, C., G. ten Bruggencate, H. Stöckle, and R. Steinberg. Calcium and potassium changes in extracellular microenvironment of cat cerebellar cortex. J. Neurophysiol. 41: 1026–1039, 1978.
 187. Nicholson, C., and J. A. Freeman. Theory of current source‐density analysis and determination of conductivity tensor for anuran cerebellum. J. Neurophysiol. 38: 356–368, 1975.
 188. Nicholson, C., and R. Llinás. Field potentials in the alligator cerebellum and theory of their relationship to Purkinje cell dendritic spikes. J. Neurophysiol. 34: 509–531, 1971.
 189. Nicholson, C., and R. Llinás. Real time current source‐density analysis using multi‐electrode array in cat cerebellum. Brain Res. 100: 418–424, 1975.
 190. Nicholson, C., R. Llinás, and W. Precht. Neural elements of the cerebellum in elasmobranch fishes: structural and functional characteristics. In: Neurobiology of Cerebellar Evolution and Development, edited by R. Llinás. Chicago: Am. Med. Assoc., 1969, p. 215–243.
 191. Nieuwenhuys, R., Comparative anatomy of the cerebellum. In Progress in Brain Research. The Cerebellum, edited by C. A. Fox and R. S. Snider. Amsterdam: Elsevier, 1967, vol. 25, p. 1–83.
 192. Nieuwenhuys, R., and C. Nicholson. A survey of the general morphology of the fiber connections and the possible functional significance of the giganto‐cerebellum of mormyrid fishes. In: Neurobiology of Cerebellar Evolution and Development, edited by R. Llinás. Chicago: Am. Med. Assoc., 1969, p. 107–134.
 193. Nieuwenhuys, R., and C. Nicholson. Aspects of the histology of the cerebellum of mormyrid fishes. In: Neurobiology of Cerebellar Evolution and Development, edited by R. Llinás. Chicago: Am. Med. Assoc., 1969, p. 135–169.
 194. Obata, K., M. Ito, R. Ochi, and N. Sato. Pharmacological properties of the postsynaptic inhibition by Purkinje cell axons and the action of γ‐aminobutyric acid on Deiters' neurones. Exp. Brain Res. 4: 43–57, 1967.
 195. Oscarsson, O., The sagittal organization of the cerebellar anterior lobe as revealed by the projection patterns of the climbing fiber system. In: Neurobiology of Cerebellar Evolution and Development, edited by R. Llinás. Chicago: Am. Med. Assoc., 1969, p. 525–537.
 196. Oscarsson, O., Functional organization of spinocerebellar paths. In: Handbook of Sensory Physiology. Somatosensory System, edited by A. Iggo. New York: Springer‐Verlag, 1973, vol. II, p. 339–380.
 197. Oscarsson, O., Functional organization of olivary projection to the cerebellar anterior lobe. In: The Inferior Olivary Nucleus: Anatomy and Physiology, edited by J. Courville, C. de Montigny, and Y. Lamarre. New York: Raven, 1980, p. 279–289.
 198. Oscarsson, O., and B. Sjölund. The ventral spino‐olivocerebellar system in the cat. I. Identification of five paths and their termination in the cerebellar anterior lobe. Exp. Brain Res. 28: 469–486, 1977.
 199. Oscarsson, O., and B. Sjölund. The ventral spino‐olivocerebellar system in the cat. III. Functional characteristics of the five paths. Exp. Brain Res. 28: 505–520, 1977.
 200. Palay, S. L., and V. Chan‐Palay. Cerebellar Cortex; Cytology and Organization. Berlin: Springer‐Verlag, 1974.
 201. Palkovits, M., P. Magyar, and J. Szentágothai. Quantitative histological analysis of the cerebellar cortex in the cat. II. Cell numbers and densities in the granular layer. Brain Res. 32: 15–30, 1971.
 202. Palkovits, M., P. Magyar, and J. Szentágothai. Quantitative histological analysis of the cerebellar cortex in the cat. III. Structural organization of the molecular layer. Brain Res. 34: 1–18, 1971.
 203. Paul, D. H. Electrical activity in the cerebellum of the spiny dogfish (Squalus acanthias). J. Physiol. London 191: 68P–70P, 1967.
 204. Paul, D. H., Electrophysiological studies on parallel fibers of the corpus cerebelli of the dogfish Scyliorhinus canicula. In: Neurobiology of Cerebellar Evolution and Development, edited by R. Llinás. Chicago: Am. Med. Assoc., 1969, p. 245–249.
 205. Pearson, A. A. The acustico‐lateral centers and the cerebellum, with fiber connections, of fishes. J. Comp. Neurol. 65: 201–294, 1936.
 206. Pellet, J., M. F. Tardy, F. Harlay, S. Dubrocar, and J. C. Gilhodes. Activité spontanée des cellules de Purkinje chez le chat chronique: étude statistique des spikes complexes. Brain Res. 81: 75–95, 1974.
 207. Pellionisz, A. Cerebellar control theory. Recent Dev. Neurobiol. Hung. 8: 211–243, 1978.
 208. Pellionisz, A., and R. Llinás. Brain modeling by tensor network theory and computer simulation. The cerebellum: parallel processor for predictive coordination. Neuroscience 4: 323–348, 1979.
 209. Pellionisz, A., and R. Llinás. Tensorial approach to the geometry of brain function: cerebellar coordination via metric tensor. Neuroscience 5: 1125–1136, 1980.
 210. Pellionisz, A., R. Lliná and D. H. Perkel. A computer model of the cerebellar cortex of the frog. Neuroscience 2: 19–36, 1977.
 211. Peterson, R. Electrical response of the goldfish cerebellum. I. Response to parallel fiber and peduncle stimulation. Brain Res. 47: 67–79, 1972.
 212. Pompeiano, O. Responses to electrical stimulation of the intermediate part of the cerebellar anterior lobe in the decerebrate cat. Arch. Ital. Biol. 96: 330–360, 1958.
 213. Precht, W. Neuronal Operations in the Vestibular System. Berlin: Springer‐Verlag, 1978.
 214. Precht, W., and R. Llinás. Comparative aspects of the vestibular input to the cerebellum. In: Neurobiology of Cerebellar Evolution and Development, edited by R. Llinás. Chicago: Am. Med. Assoc., 1969, p. 677–702.
 215. Ramón Y Cajal, S. Histologic du système nerveux de l'homme et des vertébrés. Paris: Maloine, 1909–1911, vols. I and II.
 216. Rolando, L. Experience sur les fonctions du système nerveux. J. Physiol. Exp. 3: 113–114, 1823.
 217. Rubia, F. J., U. Hoeppener, and H. Langhof. Lateral inhibition of Purkinje cells through climbing fiber afferents. Brain Res. 70: 153–156, 1974.
 218. Rubia, F. J., and W. Lange. Lateral inhibition of Purkinje cells via climbing fibres. Pfluegers Arch. 347: R51, 1974.
 219. Rushmer, D. S., and D. J. Woodward. Inhibition of Purkinje cells in the frog cerebellum. I. Evidence for a stellate cell inhibitory pathway. Brain Res. 33: 83–90, 1971.
 220. Schaper, A. The finer structure of the selachian cerebellum (Mustelus vulgaris) as shown by chrome‐silver preparation. J. Comp. Neurol. 8: 1–20, 1898.
 221. Schlegel, P. A. Perception of objects in weakly electric fish Gymnotus carapo as studied in recordings from rhombencephalic neurons. Exp. Brain Res. 18: 340–354, 1973.
 222. Schnitzlein, H. N., and J. R. Faucette. General morphology of the fish cerebellum. In: Neurobiology of Cerebellar Evolution and Development, edited by R. Llinás. Chicago: Am. Med. Assoc., 1969, p. 77–106.
 223. Schwartzkroin, P. A., and M. Slawsky. Probable calcium spikes in hippocampal neurons. Brain Res. 135: 157–161, 1977.
 224. Shambes, G. M., J. M. Gibson, and W. Welker. Fractured somatotopy in granule cell tactile areas of rat cerebellar hemispheres revealed by micromapping. Brain Behav. Evol. 15: 94–140, 1978.
 225. Shinoda, Y., and K. Yoshida. Neural pathways from the vestibular labyrinths to the flocculus in the cat. Exp. Brain Res. 22: 97–111, 1975.
 226. Siggins, G. R., B. J. Hoffer, A. P. Oliver, and F. E. Bloom. Activation of a central noradrenergic projection to the cerebellum. Nature London 233: 481–483, 1971.
 227. Simpson, J. I., W. Precht, and R. Llinás. Sensory separation in climbing and mossy fiber inputs to cat vestibulocerebellum. Pfluegers Arch. 351: 183–193, 1974.
 228. Smolyaninov, V. A., Some special features of organization of the cerebellar cortex. In: Models of the Structural‐Functional Organization of Certain Biological Systems, edited by I. M. Gel'fand, V. S. Gurfinkel', S. V. Fomin, and M. L. Tsetlin. Cambridge: MIT Press, 1971, p. 251–325. {Transl. from Russian by Carol R. Beard, 1966.}.
 229. Sotelo, C., Ultrastructural aspects of the cerebellar cortex of the frog. In: Neurobiology of Cerebellar Evolution and Development, edited by R. Llinás. Chicago: Am. Med. Assoc., 1969, p. 327–371.
 230. Sotelo, C., D. E. Hillman, A. J. Zamora, and R. Llinás. Climbing fiber deafferentation: its action on Purkinje cell dendritic spines. Brain Res. 98: 574–581, 1975.
 231. Sotelo, C., and R. Llinás. Specialized membrane junctions between neurons in the vertebrate cerebellar cortex. J. Cell Biol. 53: 271–289, 1972.
 232. Sotelo, C., R. Llinás, and R. Baker. Structural study of the inferior olivary nucleus of the cat: morphological correlates of electrotonic coupling. J. Neurophysiol. 37: 541–559, 1974.
 233. Spencer, W. A., The physiology of supraspinal neurons in mammals. In: Handbook of Physiology. The Nervous System, edited by J. M. Brookhart and V. B. Mountcastle. Bethesda, MD: Am. Physiol. Soc., 1977, sect. 1, vol. I, pt. 2, chapt. 26, p. 969–1021.
 234. Spira, M. E., and M. V. L. Bennett. Synaptic control of electrotonic coupling between neurons. Brain Res. 37: 294–300, 1972.
 235. Stensiö, E. The brain and the cranial nerves in fossil, lower craniate vertebrates. Skr. Nor. Vidensk. Akad. Oslo 13: 1–120, 1963.
 236. Sugimori, M., and R. Llinás. Lidocaine differentially blocks fast and slowly inactivating sodium conductance in Purkinje cells: an in vitro study in guinea pig cerebellum using iontophoretic glutamic acid. Soc. Neurosci. Abstr. 6: 468, 1980.
 237. Szentágothai, J. Structuro‐functional considerations of the cerebellar neuron network. Proc. IEEE 56: 960–968, 1968.
 238. Szentágothai, J., Glomerular synapses, complex synaptic arrangements, and their operational significance. In: The Neurosciences: Second Study Program, edited by F. O. Schmitt. New York: Rockefeller Univ. Press, 1970, p. 427–443.
 239. Szentágothai, J., and K. Rajkovitz. Ueber den Ursprung der Kletterfasern des Kleinhirns. Z. Anat. Entwicklungsgesch. 121: 130–141, 1959.
 240. Terzuolo, C. A. Cerebellar inhibitory and excitatory actions upon spinal extensor motoneurons. Arch. Ital. Biol. 97: 316–339, 1959.
 241. Thach, W. T. Somatosensory receptive fields of single units in cat cerebellar cortex. J. Neurophysiol. 30: 675–696, 1967.
 242. Thach, W. T. Discharge of Purkinje and cerebellar nuclear neurons during rapidly alternating arm movements in the monkey. J. Neurophysiol. 31: 785–797, 1968.
 243. Thach, W. T. Cerebellar output: properties, synthesis and uses. Brain Res. 40: 89–97, 1972.
 244. Tolbert, D. L., H. Bantli, and J. R. Bloedel. Anatomical and physiological evidence for a cerebellar nucleocortical projection in the cat. Neuroscience 1: 205–217, 1976.
 245. Tsukahara, N., Electrophysiological study of cerebellar nucleus neurones in the dogfish, Mustelus canis. In: Neurobiology of Cerebellar Evolution and Development, edited by R. Llinás. Chicago: Am. Med. Assoc., 1969, p. 251–256.
 246. Tsukahara, N., K. Toyama, K. Kosaka, and M. Udo. Disfacilitation of red nucleus neurones. Experientia 21: 544, 1965.
 247. Tuge, H. Studies on cerebellar function in the teleost. II. Is there a cerebello‐tectal path? Marchi method. J. Comp. Neurol. 61: 225–251, 1934.
 248. Voogd, J., The importance of fiber connections in the comparative anatomy of the mammalian cerebellum. In: Neurobiology of Cerebellar Evolution and Development, edited by R. Llinás. Chicago: Am. Med. Assoc., 1969, p. 493–514.
 249. Waespe, W., and V. Henn. The velocity response of vestibular nucleus neurons during vestibular, visual and combined angular acceleration. Exp. Brain Res. 37: 337–347, 1979.
 250. Walsh, J. V., J. C. Houk, and E. Mugnaini. Identification of unitary potentials in turtle cerebellum and correlations with structures in granular layer. J. Neurophysiol. 37: 30–47, 1974.
 251. Whitlock, D. G. A neurohistological and neurophysiological study of afferent fiber tracts and receptive areas of the avian cerebellum. J. Comp. Neurol. 97: 567–635, 1952.
 252. Willis, T. Cerebri anatome; cui accessit nervorum descripto et usus. Amsterdam: Schagen, 1664.
 253. Wilson, V. J., M. Maeda, and J. I. Franck. Input from neck afferents to the cat flocculus. Brain Res. 89: 133–138, 1975.
 254. Wong, R. K. S., D. A. Prince, and A. I. Basbaum. Intradendritic recordings from hippocampal neurons. Proc. Natl. Acad. Sci. USA. 76: 986–990, 1979.
 255. Yarom, Y., and R. Llinás. Electrophysiological properties of mammalian inferior olive neuron in in vitro brain stem slices and in vitro whole brain stem. Soc. Neurosci. Abstr. 5: 109, 1979.
 256. Young, W. Field potential analysis in elasmobranch cerebellum. Brain Res. 199: 101–112, 1980.

Contact Editor

Submit a note to the editor about this article by filling in the form below.

* Required Field

How to Cite

Rodolfo R. Llinás. Electrophysiology of the Cerebellar Networks. Compr Physiol 2011, Supplement 2: Handbook of Physiology, The Nervous System, Motor Control: 831-876. First published in print 1981. doi: 10.1002/cphy.cp010217