Comprehensive Physiology Wiley Online Library

Electrical Transmission: A Functional Analysis and Comparison to Chemical Transmission

Full Article on Wiley Online Library



Abstract

The sections in this article are:

1 Electrically Mediated Synaptic Transmission
1.1 Electrotonic Synapses
1.2 Rectifying Electrotonic Synapses
1.3 Electrical Inhibition
1.4 Electrical Interactions Across Extracellular Space
2 Functions of Electrotonic Transmission
2.1 Short Latency in Through‐conducting Systems
2.2 Reciprocity and Short Latency in Highly Synchronized Systems
2.3 Synchronization in Relay Nuclei and Effector Organs
2.4 Asynchronous Activity and Reciprocal Excitation
2.5 Pathways of Electrotonic Coupling in Synchronization
2.6 Synaptic Control of Degree of Coupling
2.7 Cellular Control of Electrotonic Junctions
2.8 Functions of Electrotonic Junctions in Nonelectrical Communication
3 Some “Unusual” Properties of Chemically Mediated Transmission
3.1 Tonic Release of Transmitter
3.2 PSP's Involving a Conductance Decrease and Cytoplasmic Messengers
3.3 Dual‐ and Multiple‐action Synapses
4 Functional Considerations in Mode of Transmission
4.1 Input‐Output Relations of Electrically Excitable Membrane
4.2 Identification of Mode of Transmission
4.3 Properties with Clear Advantages to Either Mode
4.4 Properties with at Most a Modest Advantage to Either Mode
4.5 Evaluation and Prospects
Figure 1. Figure 1.

Gap junctions at electrotonic synapses formed by club endings on the Mauthner cell lateral dendrite in the goldfish. A: in the central region the unit membranes approach each other closely, but over much of the region of apposition a gap of about 20 Å remains. B: in this plane of section, spots of increased density appear in the centers of the apposed membranes along much of the junction. These spots have a periodicity of about 90 Å and are in register in the two membranes. C: lanthanum injected intraventricularly fills the extracellular spaces at either end of the gap junction and penetrates the gap to the center of the junction. The staining in the gap is interrupted, perhaps periodically. D: in tangential section, lanthanum staining in a gap junction appears as a more‐or‐less hexagonal lattice. Center‐to‐center spacing is about 90 Å. A few of the clear regions outlined by the hexagonal lattice have a central dense spot (arrow). Vertical bar, 500 Å for A, C, and D and 600 Å for B.

From Brightman & Reese 47
Figure 2. Figure 2.

Electron micrographs of tight junctions between endothelial cells of cerebral capillaries in the mouse. In A‐C the vessel lumen is at the top and the basal lamina runs along the bottom. A: unit membranes contact each other and occlude the intercellular cleft at 3 points (arrow indicates the uppermost point of occlusion). Cleft is open the remainder of the distance to the basal lamina. B: intravascularly injected lanthanum fills the vessel lumen and the intercellular cleft down to a point of contact between the endothelial cell membranes. C: intraventricularly injected horseradish peroxidase permeates the basal lamina and passes up the intercellular cleft as far as a point of membrane contact. Aldehyde fixation. Vertical bars, 500 Å.

From Brightman & Reese 47
Figure 3. Figure 3.

Gap junction as seen in freeze‐fracture preparation from mouse liver. At lower left the fracture plane reveals the cytoplasmic leaflet (A face) of the cell that lay away from the viewer; at upper right the outer leaflet (B face) of the near cell is revealed. In the center is the gap junction, which appears as a large aggregate of particles in the cytoplasmic leaflet and as complementary pits in the outer leaflet. In some small regions the particles form a regular hexagonal array, but the domains of close packing are not large. At large ridges in the top center and bottom right the fracture plane broke across the extracellular space between the 2 cells. The ridge in the gap junction between leaflets showing particles and pits is much lower and reflects the close apposition of the cell membranes in this region. As is commonly but not universally found in cells, the cytoplasmic leaflet of the nonjunctional membrane contains dispersed particles that are much more numerous than those in the outer leaflet. Inset shows a region of the cytoplasmic leaflet of the junction at higher magnification. Some of the particles appear to have a small pit at their centers. Vertical bar, 0.1 μm; 0.05 μm for inset

Micrograph provided by N. B. Gilula; cf. 115
Figure 4. Figure 4.

Freeze‐fracture preparation of gap junctions between neurons. The neuron is a spinal electromotor neuron of the gymnotid electric fish Sternarchus. In sectioned material only axosomatic gap junctions are seen 252. There are 4 aggregates of particles typical of gap junctions in other structures (gj 1–4). The somatic location of the junctions is indicated by the nucleus (Nc) and the nuclear pores (p), which were exposed when the fracture plane left the surface membrane to cross the cytoplasm (Cy). The exposed surface of the neuron is the cytoplasmic leaflet (A face). Aldehyde‐fixed tissue.

C. Sandri, K. Akert, and M. V. L. Bennett, unpublished observations
Figure 5. Figure 5.

Occluding junction as seen in a freeze‐fracture preparation from epithelial cells of the rat small intestine. Microvilli protruding into the lumen point downward. The occluding junction runs along the center and appears as a meshwork of ridges and grooves. The surface at top is the outer leaflet (B face) of the cell that lay toward the reader. This leaflet contains the grooves of the fractured occluding junction. In the center and to the right of the junction are regions where the fracture plane broke through to the underlying cell exposing the cytoplasmic leaflet (B face), which contains the ridges. The membrane leaflets bulge toward the viewer between the grooves in the outer faces and away from the viewer between the ridges of the cytoplasmic faces; these bulges reflect the separation of the apposed membranes between the lines of contact. Vertical bar, 0.1 μm.

From Gilula 115
Figure 6. Figure 6.

Effects of fixation on the appearance of gap junctions in thin section. All junctions are between ependymal cells near their apices. A: osmic acid fixation followed by uranyl acetate before dehydration (en bloc staining); the characteristic gap is seen to separate the 2 unit membranes. B: permanganate fixation; there is a central dark line in the junction, but no gap is apparent. C: osmic acid fixation, but no en bloc uranyl staining. The outer lamella of the unit membranes is unstained. There is no central staining in some regions of the junction. In other regions there is a periodic row of dots with a separation of ca. 100 Å (see also Fig. 24). Vertical bar, 500 Å

Micrographs supplied by M. W. Brightman and T. S. Reese; cf. 47
Figure 7. Figure 7.

Gap junctional membranes isolated from rat liver negatively stained with phosphotungstic acid. In many regions, the junctional fragments show a regular hexagonal lattice that corresponds to the lattice seen in thin sections of lanthanum preparations. A dense spot is present in the centers of the clear areas outlined by the lattice. The periodicity of the lattice is about 100 Å. Vertical bar, 500 Å.

From Gilula 115
Figure 8. Figure 8.

Intercellular passage of the dye Procion yellow between axonal segments of crayfish septate axon. Dye was injected into the posterior segment (Ap) through a recording microelectrode that allowed monitoring of axonal impulses. The dye passed into the anterior segment (Ap) across the septum, which runs between arrows. Dye was visible in the anterior segment within 1 h, but the preparation was allowed to stand for about 15 h before it was photographed through a fluorescent microscope. (The lower molecular weight and more fluorescent dye fluorescein visibly crosses the septum within a few minutes.) Some axoplasm had coagulated in the posterior segment and was more brightly fluorescent. Ganglion, connectives, and peripheral nerves are faintly discernible through their autofluorescence. Maximum axon diameter is about 150 μm. The anterior axon narrows sharply caudal to the septum and can be seen to run ventromedially, which is toward its soma on the contralateral side

From Bennett 26
Figure 9. Figure 9.

Diagrammatic representation of a small area of gap junction. Two apposed membranes are shown in perspective as in the horizontal plane. At right the junction is split and the upper membrane is peeled back to reveal the external aspect of the lower membrane. The gap between the apposed membranes and the structures that bridge it are indicated. The membranes each contain subunits that form hemichannels and correspond to the particles seen in freeze‐fracture. Hemichannels are shown as separated by lipid in the plane of the membrane. Two aligned hemichannels form a complete channel connecting the cells interiors. Separated hemichannels are probably not open to the exterior, but no gating mechanism is shown.

N. B. Gilula, unpublished diagram
Figure 10. Figure 10.

Electrical model for an electrotonic synapse. A: 2 cells that would ordinarily be effectively isopotential. Voltage and current electrodes are indicated in each cell with measured voltages (V1 and V2) and applied currents (i1 and i2). The indifferent, or ground, electrode in the bathing medium is omitted. B: an equivalent circuit of the cells. C1 and C2 are cell capacities; r1, r2, and rc are cell and junctional (or coupling) resistances. Resistances of cytoplasm and external medium are neglected. Junctional capacity is generally negligible and is omitted.

From Bennett 18
Figure 11. Figure 11.

Electrotonic coupling of supramedullary neurons of the puffer fish. Upper trace, polarizing current. Middle and lower traces, intracellular recordings from adjacent neurons, cells 1 and 2, respectively. A: multispike response to spinal stimulation. The same number of responses occur in each cell, although the somas are not invaded during the 2nd and 3rd responses and only axon spikes are recorded. B: a depolarizing pulse which is adequate to evoke a spike, is applied to cell 2. A somewhat delayed spike is observed in cell 1, indicating propagation across the electrotonic synapses. There is also appreciable electrotonic spread of the maintained depolarization. Higher gain recording in cell 1. C and D: electrotonic spread of hyperpolarization when current is passed in cells 1 and 2, respectively. Higher gain recording in the unpolarized cell. The electrotonically spread potentials are attenuated and slowed.

From Bennett 18
Figure 12. Figure 12.

Current‐voltage relations for electrotonic coupling of the cells in Fig. 11. Cell 1 and Cell 2 correspond to the cells of the middle and lower traces, respectively. Hyperpolarization in the lower left quadrant. Larger units on the ordinate are for potentials in the polarized cell (•); smaller units are for potentials in the other cell (×). Input resistances (r11 and r22) of the 2 cells are linear for moderate hyperpolarization and differ slightly. Over the same range, the transfer resistances, that is, the ratio of potential in one cell resulting from current in the other (r12 and r21), are approximately equal, and the same resistance line is drawn for both sets of data. For larger hyperpolarizations and for depolarizations, input resistances become nonlinear, but the amount of electrotonic spread is not changed in proportion to change in resistance.

From Bennett 18
Figure 13. Figure 13.

Equivalent circuits of a “leaky” junction. A: circuit where the current path out of the middle of the junction into the extracellular medium is represented by a resistance. B: circuit equivalent to that in A obtained by a π‐T transformation of the central 3 resistors. The “leak” is electrically equivalent to additional resistances in parallel with the pre‐ and postjunctional resistances. Abbreviations as in Fig. 10.

From Bennett 18
Figure 14. Figure 14.

Calculated PSP's for an impulse into the model electrotonic synapse (Fig. 10B). The input function is V1 = (t/A)2 exp (−2t/A + 2), where t is time and A is time to peak; its impulselike shape is indicated by the curve labeled x = 0; x is the ratio of the postsynaptic time constant (including junctional resistance) to the time to peak of the impulse. The amplitude of the PSP (V2) normalized by the steady‐state coupling coefficient k from V1 to V2 is plotted against time measured in terms of rise times of the impulse, t/A. If the rise time of the presynaptic impulse is 1 ms and the postsynaptic time constant is 6.4 ms, x is 6.4 and the observed delay is about 0.3 ms. If the rise time is 0.5 ms, x is 12.8 and the delay is about 0.4 times the rise time or 0.2 ms.

From Bennett 18
Figure 15. Figure 15.

Delays at electrotonic synapses. A: transmission across the septum of the septate axon; R and C indicate impulses in the rostral and caudal segments. Propagation is in the rostrocaudal direction. Delay is less than 0.05 ms. B: transmission from giant fiber to motoneuron in the hatchetfish; the delay is about 0.05 ms; several superimposed sweeps. [From Auerbach & Bennett 9.] C: sonic muscle motoneurons in the toadfish, delays in transmission between motoneurons. Threshold stimulation of the axon of the penetrated cell, about 10 superimposed sweeps. The middle trace is the intracellular record. The upper trace is a record of the antidromic volley at the edge of the spinal cord. The lower trace is recorded later just outside the cell at the same stimulus strength and should be subtracted from the intracellular record to give transmembrane potentials. When the antidromic spike is evoked, the potential rises rapidly off screen. When the axon is not stimulated a positive negative field potential preceeds a slower depolarization of the cell membrane. This depolarization occurs at a latency of about 0.3 ms from the shortest latency antidromic spike and from the peak of the extracellular positivity generated by excitation of the neighboring cells. D: as in C but recorded at a slower sweep speed at different stimulus strengths. The antidromic depolarization is graded with the magnitude of the antidromic volley. [From Bennett 18.] E: delays in transmission between supramedullary neurons. Threshold stimulation of one cell shows that the postsynaptic component associated with the spike is delayed by about 2 ms. Voltage calibrations: 50 and 5 mV for high and low gain respectively. Time calibrations: 1 ms for A, C, and D; 0.2 ms for B; and 5 ms for E. [From Bennett et al. 35.]

From Deschěnes & Bennett 73
Figure 16. Figure 16.

Transmission of impulses across a rectifying electronic synapse, the giant motor synapse of the crayfish. A: an antidromic impulse in the postsynaptic (post) motor axon causes only a very small potential in the presynaptic (pre) fiber. B: a directly evoked impulse in the presynaptic fiber causes a large PSP in the motor axon.

From Furshpan & Potter 107
Figure 17. Figure 17.

Effect of hyperpolarization on the PSP at a rectifying electrotonic synapse, the giant fiber‐motoneuron synapse of the hatchetfish. Inset: upper trace, polarizing current; lower trace, recording in a motoneuron. Hyperpolarization augmented the PSP evoked by spinal stimulation, which activates the giant fibers. There is little change in PSP shape but a doubling in amplitude (3 superimposed records). Graph, relation of PSP amplitude and hyperpolarization; data from the same experiment as the inset.

From Auerbach & Bennett 9
Figure 18. Figure 18.

Electrically mediated inhibition at the Mauthner cell initial segment. A: recording inside and just outside the initial segment (inset).A positivity (arrow) is recorded outside about 1 ms after the large negative potential generated by the Mauthner cell axon; at the same time only a small positivity is recorded internally. Subtraction of the records (black line on white) gives the transmembrane potential; the membrane is hyperpolarized during the external positivity. The later long‐lasting positivity recorded intracellularly is an IPSP that is inverted because of leak of Cl from the recording electrode. B: stimulation of one of the electrical inhibitory neurons (upper trace) generates a small extracellular positivity of ca. 0.4 mV in the initial segment region of the Mauthner cell (lower trace). C: activation of the Mauthner cell by spinal stimulation produces a negative potential in an inhibitory neuron (upper trace), which has the same latency and time course as the negativity outside the initial segment of the Mauthner cell, although it is smaller. This response is much smaller just outside the inhibitory interneuron (lower trace), so that the membrane is hyperpolarized. D: this hyperpolarization decreases cell excitability to direct stimulation (superimposed traces with and without Mauthner cell stimulation, which is indicated by •). Time calibration same for C and D. [B‐D from Korn & Faber 188.]

From Furukawa & Furshpan 109
Figure 19. Figure 19.

Diagram and equivalent circuits of the Mauthner cell electrical inhibitory synapse. A: relation between the cells and lines of current flow are indicated on an equivalent circuit. Voltage generators are shown in the initial segment region of the Mauthner (M) cell and in the soma of the inhibitory (I) cell. The resistance of the I cell terminal is r1. The resistance of the M cell soma is rs. The extracellular (access) resistance close around the initial segment and inhibitory terminal is re. The low resistance of the large volume of tissue farther away is neglected. B: the equivalent circuit rearranged to emphasize its symmetry. Activity of the I cell produces current from right to left through re; an extracellular positivity is generated that hyperpolarizes the M cell initial segment. Activity of the M cell causes current to flow from left to right through re; the extracellular space outside the I cell soma is positive compared to that outside its terminal and the soma membrane is hyperpolarized

Adapted from Korn & Faber 188
Figure 20. Figure 20.

Electrical coupling between cells that have a circular region of low‐resistance membranes closely apposed. Left, potential in the gap Vg on the right. Because the gap widens at the edge of the apposition, Vg is assumed to go to zero at this point. V1 and V2 are the potentials in the pre‐ and postappositional cells, respectively. Current is applied in the pre cell setting up V1; the input resistance of nonappositional membrane in the post cell is assumed to be high relative to the resistance of the appositional membrane. With these assumptions and an apposition with a radius of 4 space constants defined in the usual way, the curved line in the graph shows Vg as a function of radial position in the gap. This could occur for appositional membrane with a resistivity of 1 Ωcm2, a gap width of 200 Å, a gap resistivity of 100 Ωcm, and an apposition of 2.8 μm in diameter. The coupling coefficient, V2/V1 is approximately one‐third.

From Bennett 26
Figure 21. Figure 21.

Properties of the giant electromotor neurons of the electric catfish. A: upper and lower traces, recordings from right and left cells, respectively. Brief stimuli of gradually increasing strength are applied to the nearby medulla (several superimposed sweeps; the stimulus artifact occurs near the beginning of the sweep). Depolarizations of successively increasing amplitude are evoked until in one sweep both cells generate spikes. B: 2 electrodes in the right cell, one for passing current (shown on the upper trace) and one for recording; one recording electrode in the left cell. The lower traces, which are from the recording electrodes, start from the same base line. When an impulse is evoked in the right cell by a depolarizing pulse, the left cell also generates a spike after a short delay. C: when a hyperpolarizing current is passed in the right cell, the left cell also becomes hyperpolarized, but more slowly and to a lesser degree (display as in B). D: when organ discharge is evoked by irritating the skin, a depolarization gradually rises up to the threshold of the giant cell and initiates a burst of 3 spikes (lower traces, base line indicated by superimposed sweeps). Each spike produces a response in the organ (upper trace, recorded at high gain and greatly reduced in amplitude because curare, used to prevent movement, also blocked transmission from nerve to electrocyte)

Modified from Bennett et al. 36
Figure 22. Figure 22.

Effect of polarization in a single pacemaker cell of the weakly electric gymnotid Gymnotus. Upper trace, activity in the spinal cord and peripheral nerves leading to the electric organ (recorded by needle electrodes at high gain in a curarized animal); middle trace, intracellular recording from a pacemaker cell; lower trace, current applied through the recording electrode. Two superimposed sweeps in each record, one with and one without applied current. The sweeps are triggered by the spike of the pacemaker cell. Faster sweep in A and B. A and C: a depolarizing pulse that evoked a spike advances the next and subsequent spikes but does not desynchronize or itself cause any descending activity. B and D; a hyerpolarizing pulse retards the next and subsequent spikes but does not desynchronize the descending activity.

From Bennett et al. 38
Figure 23. Figure 23.

Responses of pacemaker and relay cells in a weakly electric gymnotid Gymnotus. Upper traces, activity in the spinal cord and peripheral nerves leading to the electric organ (recorded by needle electrodes at high gain in a curarized animal); lower traces, intracellular recordings in pacemaker (A and C) and relay (B and D) neurons. Faster sweep in C and D where dotted lines indicate the times of firing of the cells in relation to the descending activity.

From Bennett et al. 38
Figure 24. Figure 24.

Electronic coupling by way of dendrodendritic synapses in electromotor relay nuclei of a mormyrid. A: light micrograph of silver‐stained preparation (Romanes' method) of the medullary relay nucleus. A thick bridge appears to connect the 2 cell bodies without an obvious intervening membrane. B: with electron microscopy such bridges are seen to have partitioning membranes (between arrows) that separate the cell bodies (s). Axon terminals (a) occur on the cells and myelinated fibers and blood vessels (bv) are present. C: at a similar region of apposition between spinal relay neurons, higher magnification shows large regions where the membranes are very closely apposed. The membranes show the ordinary extracellular space at the top and bottom of the micrograph. Material was fixed in osmic acid, and sections were stained with uranyl acetate and lead citrate. The central region of the junctions shows dots that may be periodically arranged in the lower part of the figure (see Fig. 6).

From Bennett et al. 37
Figure 25. Figure 25.

Electrotonic coupling by way of presynaptic fibers in the spinal electromotor nucleus of the electric eel. A: an axon (A) appears to synapse with the soma of an electromotor neuron (S2) and with a short dendrite from another cell (SI). B: higher magnification from the same section showing a close apposition (probable gap junction) between A and the dendrite of S1. C: a neighboring section showing a close apposition between A and S2. The axon terminal provides a short pathway between cells that can be expected to contribute to the coupling observed physiologically (dendrodendritic electrotonic synapses are also present in this nucleus). Some presynaptic vesicles are present, although no chemically mediated component is seen in the PSP's.

From Meszler et al. 224
Figure 26. Figure 26.

Synaptic control of electrotonic coupling: neurons controlling pharyngeal expansion in the mollusc Navanax. Upper traces, recording from an M (medium‐sized) cell. Middle traces, recording from an ipsilateral G (giant) cell (higher gain in row of B, B′, D, D′). Lower traces, polarizing current applied to G cell for upper records and in M cell for lower records. Primed letters indicate that a decoupling train of stimuli is applied to the large ipsilateral pharyngeal nerve at the beginning of the sweep. A and B: electrotonic spread of hyperpolarization. C and D: spread of depolarization. Spikes in the polarized cells produce small, somewhat slowed, depolarizing components in the unpolarized cell. These relatively brief potentials are superimposed on a slow depolarization. Irregularity of firing in C is due to activation of an inhibitory interneuron by stimulation in the G cell. A′ and B′: electrotonic spread of hyperpolarization is practically eliminated after decoupling stimuli. Input resistance of the cells is reduced to about one‐half. During hyperpolarization small oscillations become visible which represent variations in the synaptic activity responsible for the decoupling. These oscillations are much smaller at the resting potential because their reversal potential is near this level. C′ and D′: electrotonic spread of depolarization is greatly reduced after decoupling stimuli. Reduction in spread of maintained depolarization is greater than reduction in spread of spikes. Stimulation in the G cell no longer can excite the M cell. The M cell can still be excited by stimuli applied in its own soma, although its excitability is reduced (D′).

From Spira & Bennett 305
Figure 27. Figure 27.

Diagram and equivalent circuits for synaptic control of electrotonic coupling between neurons. A: simplest equivalent. Each cell and the coupling resistance are represented by a single resistor (r1, r2, and re, respectively). Decrease in r1 or r2 by inhibition would decrease coupling, as well as inhibit the cells. B and C: strategic localization of inhibitory synapses along the coupling pathway (filled circle endings in C) could allow effective decoupling without a great degree of inhibition of firing evoked by other inputs. Additional resistances in the equivalent circuit are ri for the inhibitory synapses and rs for the series resistance of the collaterals forming the coupling pathway. As ri becomes zero the coupling between cells vanishes, but excitatory synapses located closer to the soma could depolarize r1 and r2 and cause the cells to fire independently.

From Bennett 27
Figure 28. Figure 28.

Diagram of somatic and dendritic inputs to medial rectus oculomotor neurons of a teleost. Dendritic inputs (left arrow) are activated by stimulation of the ipsilateral eighth nerve. There is no coupling, and movements are smoothly graded in amplitude. The somatic inputs (right arrow) are activated by stimulation of the ophthalmic nerve or contralateral eighth nerve. There is weak coupling between the cell bodies (by way of the presynaptic fibers) and some increase in synchronization of firing results.

From Korn & Bennett 187
Figure 29. Figure 29.

Impulses arising from dendritic or somatic depolarizations in oculomotor neurons of a teleost. Upper and middle traces, intracellular recording at high and low gain from a medial rectus oculomotor neuron. Lower traces, efferent activity in the medial rectus nerve. A1 and A2: threshold stimulation of the ipsilateral eighth nerve, dendritic impulse initiation. Little PSP is seen whether or not an abruptly rising spike occurs. A3: stronger stimulation evokes a multispike discharge. The first 2 spikes arise abruptly. The last 2 are preceded by some PSP. B1 and B2: threshold stimulation of the ipsilateral ophthalmic nerve; impulses arise near the soma. Large slow PSP's appear to initiate the spike. B3: stronger stimulation evokes a more rapidly rising PSP and multiple spikes.

From experiments by M. E. Kriebel and M. V. L. Bennett, cf. 27
Figure 30. Figure 30.

Differential effects of hyperpolarizing current on impulses arising in the dendrites and close to the soma. Upper trace, intracellular recording from a medial rectus oculomotor neuron. Middle trace, current passed through the recording microelec‐trode with a bridge circuit. Lower trace, efferent activity in the medial rectus nerve. A1: spikes initiated in the dendrite arise abruptly from a level base line in response to stimulation of the ipsilateral eighth nerve. A2: when the same stimulus is given during a hyperpolarizing current pulse, the first response is delayed and the number of spikes is reduced, but little PSP is recorded at the times that the first and second spikes arise in A1 (2 superimposed sweeps with and without nerve stimulation). B1: spikes initiated near the soma are preceded by an obvious PSP that is evoked by ipsilateral ophthalmic nerve stimulation. B2: when the same stimulus is given during a hyperpolarizing pulse, spikes are blocked revealing a large PSP (2 superimposed sweeps, with and without nerve stimulation).

From Kriebel et al. 191
Figure 31. Figure 31.

Effects of polarizing currents on EPSP's associated with conductance increases and decreases. A: the fast EPSP evoked by a single stimulus to the presynaptic nerve is associated with a conductance increase. Hyperpolarizing (negative) currents augments the EPSP. Depolarizing (positive) currents decrease and then invert the EPSP. B: the slow EPSP is associated with a conductance decrease. After block of the fast EPSP with nicotine the slow EPSP is evoked by presynaptic stimulation at 100/s for 2 s (indicated under lowest trace). Although the slow EPSP is a depolarizing response, it is augmented by depolarization and decreased by hyperpolarization. The reversal potential for the slow EPSP is close to that for the after hyperpolarization of the spike, which suggests that the slow EPSP is due to a decrease in K conductance.

From Weight & Votava 340
Figure 32. Figure 32.

Dual‐action excitatory‐inhibitory PSP's in the buccal ganglia of Aplysia. The pre‐ and postsynaptic elements are subscripted 4 and 7, respectively. B, buccal; L, left; R, right. A: effects of altering the postsynaptic membrane potential on the PSP's and on the potentials produced by iontophoretic application of ACh to the cell soma. At the resting potential (0, middle traces) PSP shows a slight negative phase and the ACh potential is monophasic. When BL7 is depolarized by 20 mV (upper traces) PSP develops a pronounced hyperpolarizing phase after the initial depolarizing phase, and the ACh potential is also diphasic. When BL7 is hyperpolarized by 20 mV (lower traces) both PSP and ACh potential are monophasic and depolarizing. B: postsynaptic changes in excitability caused by the 2‐component PSP's. 1, when BR7 is at its resting potential, activation of BR4 causes excitation that augments during a high‐frequency burst; 2, when BR7 is depolarized to fire at a slow rate, activation of BR4 causes first inhibition and then excitation. C: pharmacological separation of the 2 components. BL7 is depolarized by 20 mV to emphasize the hyperpolarizing phase seen in seawater. Addition of hexamethonium to the bath largely blocks the initial depolarizing phase, and d‐tubocurarine largely blocks the hyperpolarizing phase. The potential remaining following application of either drug is increased compared to the corresponding component recorded in seawater, presumably because the 2 underlying conductances overlap in time.

From published 110,167 and unpublished work of D. Gardner and E. R. Kandel


Figure 1.

Gap junctions at electrotonic synapses formed by club endings on the Mauthner cell lateral dendrite in the goldfish. A: in the central region the unit membranes approach each other closely, but over much of the region of apposition a gap of about 20 Å remains. B: in this plane of section, spots of increased density appear in the centers of the apposed membranes along much of the junction. These spots have a periodicity of about 90 Å and are in register in the two membranes. C: lanthanum injected intraventricularly fills the extracellular spaces at either end of the gap junction and penetrates the gap to the center of the junction. The staining in the gap is interrupted, perhaps periodically. D: in tangential section, lanthanum staining in a gap junction appears as a more‐or‐less hexagonal lattice. Center‐to‐center spacing is about 90 Å. A few of the clear regions outlined by the hexagonal lattice have a central dense spot (arrow). Vertical bar, 500 Å for A, C, and D and 600 Å for B.

From Brightman & Reese 47


Figure 2.

Electron micrographs of tight junctions between endothelial cells of cerebral capillaries in the mouse. In A‐C the vessel lumen is at the top and the basal lamina runs along the bottom. A: unit membranes contact each other and occlude the intercellular cleft at 3 points (arrow indicates the uppermost point of occlusion). Cleft is open the remainder of the distance to the basal lamina. B: intravascularly injected lanthanum fills the vessel lumen and the intercellular cleft down to a point of contact between the endothelial cell membranes. C: intraventricularly injected horseradish peroxidase permeates the basal lamina and passes up the intercellular cleft as far as a point of membrane contact. Aldehyde fixation. Vertical bars, 500 Å.

From Brightman & Reese 47


Figure 3.

Gap junction as seen in freeze‐fracture preparation from mouse liver. At lower left the fracture plane reveals the cytoplasmic leaflet (A face) of the cell that lay away from the viewer; at upper right the outer leaflet (B face) of the near cell is revealed. In the center is the gap junction, which appears as a large aggregate of particles in the cytoplasmic leaflet and as complementary pits in the outer leaflet. In some small regions the particles form a regular hexagonal array, but the domains of close packing are not large. At large ridges in the top center and bottom right the fracture plane broke across the extracellular space between the 2 cells. The ridge in the gap junction between leaflets showing particles and pits is much lower and reflects the close apposition of the cell membranes in this region. As is commonly but not universally found in cells, the cytoplasmic leaflet of the nonjunctional membrane contains dispersed particles that are much more numerous than those in the outer leaflet. Inset shows a region of the cytoplasmic leaflet of the junction at higher magnification. Some of the particles appear to have a small pit at their centers. Vertical bar, 0.1 μm; 0.05 μm for inset

Micrograph provided by N. B. Gilula; cf. 115


Figure 4.

Freeze‐fracture preparation of gap junctions between neurons. The neuron is a spinal electromotor neuron of the gymnotid electric fish Sternarchus. In sectioned material only axosomatic gap junctions are seen 252. There are 4 aggregates of particles typical of gap junctions in other structures (gj 1–4). The somatic location of the junctions is indicated by the nucleus (Nc) and the nuclear pores (p), which were exposed when the fracture plane left the surface membrane to cross the cytoplasm (Cy). The exposed surface of the neuron is the cytoplasmic leaflet (A face). Aldehyde‐fixed tissue.

C. Sandri, K. Akert, and M. V. L. Bennett, unpublished observations


Figure 5.

Occluding junction as seen in a freeze‐fracture preparation from epithelial cells of the rat small intestine. Microvilli protruding into the lumen point downward. The occluding junction runs along the center and appears as a meshwork of ridges and grooves. The surface at top is the outer leaflet (B face) of the cell that lay toward the reader. This leaflet contains the grooves of the fractured occluding junction. In the center and to the right of the junction are regions where the fracture plane broke through to the underlying cell exposing the cytoplasmic leaflet (B face), which contains the ridges. The membrane leaflets bulge toward the viewer between the grooves in the outer faces and away from the viewer between the ridges of the cytoplasmic faces; these bulges reflect the separation of the apposed membranes between the lines of contact. Vertical bar, 0.1 μm.

From Gilula 115


Figure 6.

Effects of fixation on the appearance of gap junctions in thin section. All junctions are between ependymal cells near their apices. A: osmic acid fixation followed by uranyl acetate before dehydration (en bloc staining); the characteristic gap is seen to separate the 2 unit membranes. B: permanganate fixation; there is a central dark line in the junction, but no gap is apparent. C: osmic acid fixation, but no en bloc uranyl staining. The outer lamella of the unit membranes is unstained. There is no central staining in some regions of the junction. In other regions there is a periodic row of dots with a separation of ca. 100 Å (see also Fig. 24). Vertical bar, 500 Å

Micrographs supplied by M. W. Brightman and T. S. Reese; cf. 47


Figure 7.

Gap junctional membranes isolated from rat liver negatively stained with phosphotungstic acid. In many regions, the junctional fragments show a regular hexagonal lattice that corresponds to the lattice seen in thin sections of lanthanum preparations. A dense spot is present in the centers of the clear areas outlined by the lattice. The periodicity of the lattice is about 100 Å. Vertical bar, 500 Å.

From Gilula 115


Figure 8.

Intercellular passage of the dye Procion yellow between axonal segments of crayfish septate axon. Dye was injected into the posterior segment (Ap) through a recording microelectrode that allowed monitoring of axonal impulses. The dye passed into the anterior segment (Ap) across the septum, which runs between arrows. Dye was visible in the anterior segment within 1 h, but the preparation was allowed to stand for about 15 h before it was photographed through a fluorescent microscope. (The lower molecular weight and more fluorescent dye fluorescein visibly crosses the septum within a few minutes.) Some axoplasm had coagulated in the posterior segment and was more brightly fluorescent. Ganglion, connectives, and peripheral nerves are faintly discernible through their autofluorescence. Maximum axon diameter is about 150 μm. The anterior axon narrows sharply caudal to the septum and can be seen to run ventromedially, which is toward its soma on the contralateral side

From Bennett 26


Figure 9.

Diagrammatic representation of a small area of gap junction. Two apposed membranes are shown in perspective as in the horizontal plane. At right the junction is split and the upper membrane is peeled back to reveal the external aspect of the lower membrane. The gap between the apposed membranes and the structures that bridge it are indicated. The membranes each contain subunits that form hemichannels and correspond to the particles seen in freeze‐fracture. Hemichannels are shown as separated by lipid in the plane of the membrane. Two aligned hemichannels form a complete channel connecting the cells interiors. Separated hemichannels are probably not open to the exterior, but no gating mechanism is shown.

N. B. Gilula, unpublished diagram


Figure 10.

Electrical model for an electrotonic synapse. A: 2 cells that would ordinarily be effectively isopotential. Voltage and current electrodes are indicated in each cell with measured voltages (V1 and V2) and applied currents (i1 and i2). The indifferent, or ground, electrode in the bathing medium is omitted. B: an equivalent circuit of the cells. C1 and C2 are cell capacities; r1, r2, and rc are cell and junctional (or coupling) resistances. Resistances of cytoplasm and external medium are neglected. Junctional capacity is generally negligible and is omitted.

From Bennett 18


Figure 11.

Electrotonic coupling of supramedullary neurons of the puffer fish. Upper trace, polarizing current. Middle and lower traces, intracellular recordings from adjacent neurons, cells 1 and 2, respectively. A: multispike response to spinal stimulation. The same number of responses occur in each cell, although the somas are not invaded during the 2nd and 3rd responses and only axon spikes are recorded. B: a depolarizing pulse which is adequate to evoke a spike, is applied to cell 2. A somewhat delayed spike is observed in cell 1, indicating propagation across the electrotonic synapses. There is also appreciable electrotonic spread of the maintained depolarization. Higher gain recording in cell 1. C and D: electrotonic spread of hyperpolarization when current is passed in cells 1 and 2, respectively. Higher gain recording in the unpolarized cell. The electrotonically spread potentials are attenuated and slowed.

From Bennett 18


Figure 12.

Current‐voltage relations for electrotonic coupling of the cells in Fig. 11. Cell 1 and Cell 2 correspond to the cells of the middle and lower traces, respectively. Hyperpolarization in the lower left quadrant. Larger units on the ordinate are for potentials in the polarized cell (•); smaller units are for potentials in the other cell (×). Input resistances (r11 and r22) of the 2 cells are linear for moderate hyperpolarization and differ slightly. Over the same range, the transfer resistances, that is, the ratio of potential in one cell resulting from current in the other (r12 and r21), are approximately equal, and the same resistance line is drawn for both sets of data. For larger hyperpolarizations and for depolarizations, input resistances become nonlinear, but the amount of electrotonic spread is not changed in proportion to change in resistance.

From Bennett 18


Figure 13.

Equivalent circuits of a “leaky” junction. A: circuit where the current path out of the middle of the junction into the extracellular medium is represented by a resistance. B: circuit equivalent to that in A obtained by a π‐T transformation of the central 3 resistors. The “leak” is electrically equivalent to additional resistances in parallel with the pre‐ and postjunctional resistances. Abbreviations as in Fig. 10.

From Bennett 18


Figure 14.

Calculated PSP's for an impulse into the model electrotonic synapse (Fig. 10B). The input function is V1 = (t/A)2 exp (−2t/A + 2), where t is time and A is time to peak; its impulselike shape is indicated by the curve labeled x = 0; x is the ratio of the postsynaptic time constant (including junctional resistance) to the time to peak of the impulse. The amplitude of the PSP (V2) normalized by the steady‐state coupling coefficient k from V1 to V2 is plotted against time measured in terms of rise times of the impulse, t/A. If the rise time of the presynaptic impulse is 1 ms and the postsynaptic time constant is 6.4 ms, x is 6.4 and the observed delay is about 0.3 ms. If the rise time is 0.5 ms, x is 12.8 and the delay is about 0.4 times the rise time or 0.2 ms.

From Bennett 18


Figure 15.

Delays at electrotonic synapses. A: transmission across the septum of the septate axon; R and C indicate impulses in the rostral and caudal segments. Propagation is in the rostrocaudal direction. Delay is less than 0.05 ms. B: transmission from giant fiber to motoneuron in the hatchetfish; the delay is about 0.05 ms; several superimposed sweeps. [From Auerbach & Bennett 9.] C: sonic muscle motoneurons in the toadfish, delays in transmission between motoneurons. Threshold stimulation of the axon of the penetrated cell, about 10 superimposed sweeps. The middle trace is the intracellular record. The upper trace is a record of the antidromic volley at the edge of the spinal cord. The lower trace is recorded later just outside the cell at the same stimulus strength and should be subtracted from the intracellular record to give transmembrane potentials. When the antidromic spike is evoked, the potential rises rapidly off screen. When the axon is not stimulated a positive negative field potential preceeds a slower depolarization of the cell membrane. This depolarization occurs at a latency of about 0.3 ms from the shortest latency antidromic spike and from the peak of the extracellular positivity generated by excitation of the neighboring cells. D: as in C but recorded at a slower sweep speed at different stimulus strengths. The antidromic depolarization is graded with the magnitude of the antidromic volley. [From Bennett 18.] E: delays in transmission between supramedullary neurons. Threshold stimulation of one cell shows that the postsynaptic component associated with the spike is delayed by about 2 ms. Voltage calibrations: 50 and 5 mV for high and low gain respectively. Time calibrations: 1 ms for A, C, and D; 0.2 ms for B; and 5 ms for E. [From Bennett et al. 35.]

From Deschěnes & Bennett 73


Figure 16.

Transmission of impulses across a rectifying electronic synapse, the giant motor synapse of the crayfish. A: an antidromic impulse in the postsynaptic (post) motor axon causes only a very small potential in the presynaptic (pre) fiber. B: a directly evoked impulse in the presynaptic fiber causes a large PSP in the motor axon.

From Furshpan & Potter 107


Figure 17.

Effect of hyperpolarization on the PSP at a rectifying electrotonic synapse, the giant fiber‐motoneuron synapse of the hatchetfish. Inset: upper trace, polarizing current; lower trace, recording in a motoneuron. Hyperpolarization augmented the PSP evoked by spinal stimulation, which activates the giant fibers. There is little change in PSP shape but a doubling in amplitude (3 superimposed records). Graph, relation of PSP amplitude and hyperpolarization; data from the same experiment as the inset.

From Auerbach & Bennett 9


Figure 18.

Electrically mediated inhibition at the Mauthner cell initial segment. A: recording inside and just outside the initial segment (inset).A positivity (arrow) is recorded outside about 1 ms after the large negative potential generated by the Mauthner cell axon; at the same time only a small positivity is recorded internally. Subtraction of the records (black line on white) gives the transmembrane potential; the membrane is hyperpolarized during the external positivity. The later long‐lasting positivity recorded intracellularly is an IPSP that is inverted because of leak of Cl from the recording electrode. B: stimulation of one of the electrical inhibitory neurons (upper trace) generates a small extracellular positivity of ca. 0.4 mV in the initial segment region of the Mauthner cell (lower trace). C: activation of the Mauthner cell by spinal stimulation produces a negative potential in an inhibitory neuron (upper trace), which has the same latency and time course as the negativity outside the initial segment of the Mauthner cell, although it is smaller. This response is much smaller just outside the inhibitory interneuron (lower trace), so that the membrane is hyperpolarized. D: this hyperpolarization decreases cell excitability to direct stimulation (superimposed traces with and without Mauthner cell stimulation, which is indicated by •). Time calibration same for C and D. [B‐D from Korn & Faber 188.]

From Furukawa & Furshpan 109


Figure 19.

Diagram and equivalent circuits of the Mauthner cell electrical inhibitory synapse. A: relation between the cells and lines of current flow are indicated on an equivalent circuit. Voltage generators are shown in the initial segment region of the Mauthner (M) cell and in the soma of the inhibitory (I) cell. The resistance of the I cell terminal is r1. The resistance of the M cell soma is rs. The extracellular (access) resistance close around the initial segment and inhibitory terminal is re. The low resistance of the large volume of tissue farther away is neglected. B: the equivalent circuit rearranged to emphasize its symmetry. Activity of the I cell produces current from right to left through re; an extracellular positivity is generated that hyperpolarizes the M cell initial segment. Activity of the M cell causes current to flow from left to right through re; the extracellular space outside the I cell soma is positive compared to that outside its terminal and the soma membrane is hyperpolarized

Adapted from Korn & Faber 188


Figure 20.

Electrical coupling between cells that have a circular region of low‐resistance membranes closely apposed. Left, potential in the gap Vg on the right. Because the gap widens at the edge of the apposition, Vg is assumed to go to zero at this point. V1 and V2 are the potentials in the pre‐ and postappositional cells, respectively. Current is applied in the pre cell setting up V1; the input resistance of nonappositional membrane in the post cell is assumed to be high relative to the resistance of the appositional membrane. With these assumptions and an apposition with a radius of 4 space constants defined in the usual way, the curved line in the graph shows Vg as a function of radial position in the gap. This could occur for appositional membrane with a resistivity of 1 Ωcm2, a gap width of 200 Å, a gap resistivity of 100 Ωcm, and an apposition of 2.8 μm in diameter. The coupling coefficient, V2/V1 is approximately one‐third.

From Bennett 26


Figure 21.

Properties of the giant electromotor neurons of the electric catfish. A: upper and lower traces, recordings from right and left cells, respectively. Brief stimuli of gradually increasing strength are applied to the nearby medulla (several superimposed sweeps; the stimulus artifact occurs near the beginning of the sweep). Depolarizations of successively increasing amplitude are evoked until in one sweep both cells generate spikes. B: 2 electrodes in the right cell, one for passing current (shown on the upper trace) and one for recording; one recording electrode in the left cell. The lower traces, which are from the recording electrodes, start from the same base line. When an impulse is evoked in the right cell by a depolarizing pulse, the left cell also generates a spike after a short delay. C: when a hyperpolarizing current is passed in the right cell, the left cell also becomes hyperpolarized, but more slowly and to a lesser degree (display as in B). D: when organ discharge is evoked by irritating the skin, a depolarization gradually rises up to the threshold of the giant cell and initiates a burst of 3 spikes (lower traces, base line indicated by superimposed sweeps). Each spike produces a response in the organ (upper trace, recorded at high gain and greatly reduced in amplitude because curare, used to prevent movement, also blocked transmission from nerve to electrocyte)

Modified from Bennett et al. 36


Figure 22.

Effect of polarization in a single pacemaker cell of the weakly electric gymnotid Gymnotus. Upper trace, activity in the spinal cord and peripheral nerves leading to the electric organ (recorded by needle electrodes at high gain in a curarized animal); middle trace, intracellular recording from a pacemaker cell; lower trace, current applied through the recording electrode. Two superimposed sweeps in each record, one with and one without applied current. The sweeps are triggered by the spike of the pacemaker cell. Faster sweep in A and B. A and C: a depolarizing pulse that evoked a spike advances the next and subsequent spikes but does not desynchronize or itself cause any descending activity. B and D; a hyerpolarizing pulse retards the next and subsequent spikes but does not desynchronize the descending activity.

From Bennett et al. 38


Figure 23.

Responses of pacemaker and relay cells in a weakly electric gymnotid Gymnotus. Upper traces, activity in the spinal cord and peripheral nerves leading to the electric organ (recorded by needle electrodes at high gain in a curarized animal); lower traces, intracellular recordings in pacemaker (A and C) and relay (B and D) neurons. Faster sweep in C and D where dotted lines indicate the times of firing of the cells in relation to the descending activity.

From Bennett et al. 38


Figure 24.

Electronic coupling by way of dendrodendritic synapses in electromotor relay nuclei of a mormyrid. A: light micrograph of silver‐stained preparation (Romanes' method) of the medullary relay nucleus. A thick bridge appears to connect the 2 cell bodies without an obvious intervening membrane. B: with electron microscopy such bridges are seen to have partitioning membranes (between arrows) that separate the cell bodies (s). Axon terminals (a) occur on the cells and myelinated fibers and blood vessels (bv) are present. C: at a similar region of apposition between spinal relay neurons, higher magnification shows large regions where the membranes are very closely apposed. The membranes show the ordinary extracellular space at the top and bottom of the micrograph. Material was fixed in osmic acid, and sections were stained with uranyl acetate and lead citrate. The central region of the junctions shows dots that may be periodically arranged in the lower part of the figure (see Fig. 6).

From Bennett et al. 37


Figure 25.

Electrotonic coupling by way of presynaptic fibers in the spinal electromotor nucleus of the electric eel. A: an axon (A) appears to synapse with the soma of an electromotor neuron (S2) and with a short dendrite from another cell (SI). B: higher magnification from the same section showing a close apposition (probable gap junction) between A and the dendrite of S1. C: a neighboring section showing a close apposition between A and S2. The axon terminal provides a short pathway between cells that can be expected to contribute to the coupling observed physiologically (dendrodendritic electrotonic synapses are also present in this nucleus). Some presynaptic vesicles are present, although no chemically mediated component is seen in the PSP's.

From Meszler et al. 224


Figure 26.

Synaptic control of electrotonic coupling: neurons controlling pharyngeal expansion in the mollusc Navanax. Upper traces, recording from an M (medium‐sized) cell. Middle traces, recording from an ipsilateral G (giant) cell (higher gain in row of B, B′, D, D′). Lower traces, polarizing current applied to G cell for upper records and in M cell for lower records. Primed letters indicate that a decoupling train of stimuli is applied to the large ipsilateral pharyngeal nerve at the beginning of the sweep. A and B: electrotonic spread of hyperpolarization. C and D: spread of depolarization. Spikes in the polarized cells produce small, somewhat slowed, depolarizing components in the unpolarized cell. These relatively brief potentials are superimposed on a slow depolarization. Irregularity of firing in C is due to activation of an inhibitory interneuron by stimulation in the G cell. A′ and B′: electrotonic spread of hyperpolarization is practically eliminated after decoupling stimuli. Input resistance of the cells is reduced to about one‐half. During hyperpolarization small oscillations become visible which represent variations in the synaptic activity responsible for the decoupling. These oscillations are much smaller at the resting potential because their reversal potential is near this level. C′ and D′: electrotonic spread of depolarization is greatly reduced after decoupling stimuli. Reduction in spread of maintained depolarization is greater than reduction in spread of spikes. Stimulation in the G cell no longer can excite the M cell. The M cell can still be excited by stimuli applied in its own soma, although its excitability is reduced (D′).

From Spira & Bennett 305


Figure 27.

Diagram and equivalent circuits for synaptic control of electrotonic coupling between neurons. A: simplest equivalent. Each cell and the coupling resistance are represented by a single resistor (r1, r2, and re, respectively). Decrease in r1 or r2 by inhibition would decrease coupling, as well as inhibit the cells. B and C: strategic localization of inhibitory synapses along the coupling pathway (filled circle endings in C) could allow effective decoupling without a great degree of inhibition of firing evoked by other inputs. Additional resistances in the equivalent circuit are ri for the inhibitory synapses and rs for the series resistance of the collaterals forming the coupling pathway. As ri becomes zero the coupling between cells vanishes, but excitatory synapses located closer to the soma could depolarize r1 and r2 and cause the cells to fire independently.

From Bennett 27


Figure 28.

Diagram of somatic and dendritic inputs to medial rectus oculomotor neurons of a teleost. Dendritic inputs (left arrow) are activated by stimulation of the ipsilateral eighth nerve. There is no coupling, and movements are smoothly graded in amplitude. The somatic inputs (right arrow) are activated by stimulation of the ophthalmic nerve or contralateral eighth nerve. There is weak coupling between the cell bodies (by way of the presynaptic fibers) and some increase in synchronization of firing results.

From Korn & Bennett 187


Figure 29.

Impulses arising from dendritic or somatic depolarizations in oculomotor neurons of a teleost. Upper and middle traces, intracellular recording at high and low gain from a medial rectus oculomotor neuron. Lower traces, efferent activity in the medial rectus nerve. A1 and A2: threshold stimulation of the ipsilateral eighth nerve, dendritic impulse initiation. Little PSP is seen whether or not an abruptly rising spike occurs. A3: stronger stimulation evokes a multispike discharge. The first 2 spikes arise abruptly. The last 2 are preceded by some PSP. B1 and B2: threshold stimulation of the ipsilateral ophthalmic nerve; impulses arise near the soma. Large slow PSP's appear to initiate the spike. B3: stronger stimulation evokes a more rapidly rising PSP and multiple spikes.

From experiments by M. E. Kriebel and M. V. L. Bennett, cf. 27


Figure 30.

Differential effects of hyperpolarizing current on impulses arising in the dendrites and close to the soma. Upper trace, intracellular recording from a medial rectus oculomotor neuron. Middle trace, current passed through the recording microelec‐trode with a bridge circuit. Lower trace, efferent activity in the medial rectus nerve. A1: spikes initiated in the dendrite arise abruptly from a level base line in response to stimulation of the ipsilateral eighth nerve. A2: when the same stimulus is given during a hyperpolarizing current pulse, the first response is delayed and the number of spikes is reduced, but little PSP is recorded at the times that the first and second spikes arise in A1 (2 superimposed sweeps with and without nerve stimulation). B1: spikes initiated near the soma are preceded by an obvious PSP that is evoked by ipsilateral ophthalmic nerve stimulation. B2: when the same stimulus is given during a hyperpolarizing pulse, spikes are blocked revealing a large PSP (2 superimposed sweeps, with and without nerve stimulation).

From Kriebel et al. 191


Figure 31.

Effects of polarizing currents on EPSP's associated with conductance increases and decreases. A: the fast EPSP evoked by a single stimulus to the presynaptic nerve is associated with a conductance increase. Hyperpolarizing (negative) currents augments the EPSP. Depolarizing (positive) currents decrease and then invert the EPSP. B: the slow EPSP is associated with a conductance decrease. After block of the fast EPSP with nicotine the slow EPSP is evoked by presynaptic stimulation at 100/s for 2 s (indicated under lowest trace). Although the slow EPSP is a depolarizing response, it is augmented by depolarization and decreased by hyperpolarization. The reversal potential for the slow EPSP is close to that for the after hyperpolarization of the spike, which suggests that the slow EPSP is due to a decrease in K conductance.

From Weight & Votava 340


Figure 32.

Dual‐action excitatory‐inhibitory PSP's in the buccal ganglia of Aplysia. The pre‐ and postsynaptic elements are subscripted 4 and 7, respectively. B, buccal; L, left; R, right. A: effects of altering the postsynaptic membrane potential on the PSP's and on the potentials produced by iontophoretic application of ACh to the cell soma. At the resting potential (0, middle traces) PSP shows a slight negative phase and the ACh potential is monophasic. When BL7 is depolarized by 20 mV (upper traces) PSP develops a pronounced hyperpolarizing phase after the initial depolarizing phase, and the ACh potential is also diphasic. When BL7 is hyperpolarized by 20 mV (lower traces) both PSP and ACh potential are monophasic and depolarizing. B: postsynaptic changes in excitability caused by the 2‐component PSP's. 1, when BR7 is at its resting potential, activation of BR4 causes excitation that augments during a high‐frequency burst; 2, when BR7 is depolarized to fire at a slow rate, activation of BR4 causes first inhibition and then excitation. C: pharmacological separation of the 2 components. BL7 is depolarized by 20 mV to emphasize the hyperpolarizing phase seen in seawater. Addition of hexamethonium to the bath largely blocks the initial depolarizing phase, and d‐tubocurarine largely blocks the hyperpolarizing phase. The potential remaining following application of either drug is increased compared to the corresponding component recorded in seawater, presumably because the 2 underlying conductances overlap in time.

From published 110,167 and unpublished work of D. Gardner and E. R. Kandel
References
 1. Albe‐Fessard, D., and H. Martins‐Ferreira. Role de la commande nerveuse dans la synchronisation du fonctionnement des éléments de l'organe electrique du Gymnote, Electrophorus electricus L. J. Physiol. Paris 45: 533–546, 1953.
 2. Albertini, D. F., and E. Anderson. Structural modifications of lutein cell gap junctions during pregnancy in the rat and the mouse. Anat. Record 181: 171–194, 1975.
 3. Albuquerque, E., J. E. Warnick, J. R. Tasse, and F. M. Sansone. Effects of vinblastine and colchicine on neural regulation of the fast and slow skeletal muscles of the rat. Exptl. Neurol. 37: 607–634, 1972.
 4. Anderson, C. R., and C. F. Stevens. Voltage clamp analysis of acetylcholine produced end‐plate current fluctuations at frog neuromuscular junction. J. Physiol. London 235: 655–691, 1973.
 5. Arvanitaki, A. Effects evoked in an axon by the activity of a contiguous one. J. Neurophysiol. 5: 89–108, 1942.
 6. Arvanitaki, A., and N. Chalazonitis. Interactions électriques entre le soma géant A et les somata immédiatement contigus (Ganglion pleuro‐branchial d'Aplysia). Bull. Inst. Oceanogr. 1143: 1–30, 1959.
 7. Asada, Y., and M. V. L. Bennett. Experimental alteration of coupling resistance at an electrotonic synapse. J. Cell Biol. 49: 159–172, 1971.
 8. Ascher, P. Inhibitory and excitatory effects of dopamine of Aplysia neurones. J. Physiol. London 225: 173–209, 1972.
 9. Auerbach, A. A., and M. V. L. Bennett. Chemically mediated transmission at a giant fiber synapse in the central nervous system of a vertebrate. J. Gen. Physiol. 53: 183–210, 1969.
 10. Auerbach, A. A., and M. V. L. Bennett. A rectifying synapse in the central nervous system of a vertebrate. J. Gen. Physiol. 53: 211–237, 1969.
 11. Azarnia, R., and W. R. Loewenstein. Intercellular communication and tissue growth. V. A cancer cell strain that fails to make permeable membrane junctions with normal cells. J. Membrane Biol. 6: 368–385, 1971.
 12. Baker, R., and R. Llinás. Electrotonic coupling between neurons in the rat mesencephalic nucleus. J. Physiol. London 212: 45–63, 1971.
 13. Barker, J. L., and H. Gainer. Studies on bursting pacemaker potential activity in molluscan neurons. I. Membrane properties and ionic contributions. Brain Res. 84: 461–477, 1975.
 14. Barker, J. L., M. S. Ifshin, and H. Gainer. Studies on bursting pacemaker potential activity in molluscan neurons. III. Effects of hormones. Brain Res. 84: 501–513, 1975.
 15. Barr, L., W. Berger, and M. M. Dewey. Electrical transmission at the nexus between smooth muscle cells. J. Gen. Physiol. 51: 347–369, 1968.
 16. Barr, L., M. M. Dewey, and W. Berger. Propagation of action potentials and the structure of the nexus in cardiac muscle. J. Gen. Physiol. 48: 797–823, 1965.
 17. Baylor, D. A., M. G. F. Fuortes, and P. M. O'Bryan. Receptive fields of cones in the retina of the turtle. J. Physiol. London 214: 265–294, 1971.
 18. Baylor, D. A., and J. G. Nicholls. After‐effects of nerve impulses on signalling in the central nervous system of the leech. J. Physiol. London 203: 571–589, 1969.
 19. Benedetti, E. L., I. Dunia, and H. Broemendal. Development of junctions during differentiation of lens fibers. Proc. Nat. Acad. Sci. US 71: 5073–5077, 1974.
 20. Bennett, M. V. L. Physiology of electrotonic junctions. Ann. NY Acad. Sci. 137: 509–539, 1966.
 21. Bennett, M. V. L. Electrical connections between supramedullary neurons. Federation Proc. 19: 282, 1960.
 22. Bennett, M. V. L. Neural control of electric organs. In: The Central Nervous System and Fish Behavior, edited by D. Ingle. Chicago: Univ. of Chicago Press, 1968, p. 147–169.
 23. Bennett, M. V. L. Similarities between chemically and electrically mediated transmission. In: Physiological and Biophysical Aspects of Nervous Integration, edited by F. D. Carlson Englewood Cliffs, N.J.: Prentice‐Hall, 1968, p. 73–128.
 24. Bennett, M. V. L. Electric organs. In: Fish Physiology, edited by W. S. Hoar and D. J. Randall. New York: Academic, 1971, vol. 5, p. 347–491.
 25. Bennett, M. V. L. Electroreception. In: Fish Physiology, edited by W. S. Hoar and D. J. Randall. New York: Academic, 1971, vol. 5, p. 493–574.
 26. Bennett, M. V. L. Electrolocation in fish. Ann. NY Acad. Sci. 188: 242–269, 1971.
 27. Bennett, M. V. L. A comparison of electrically and chemically mediated transmission. In: Structure and Function of Synapses, edited by G. D. Pappas and D. P. Purpura. New York: Raven, 1972, p. 221–256.
 28. Bennett, M. V. L. Function of electrotonic junctions in embryonic and adult tissues. Federation Proc. 32: 65–75, 1973.
 29. Bennett, M. V. L. Permeability and structure of electrotonic junctions and intercellular movement of tracers. In: Intracellular Staining Techniques in Neurobiology, edited by S. D. Kater and C. Nicholson. New York: Elsevier, 1973, p. 115–133.
 30. Bennett, M. V. L. Flexibility and rigidity in electrotonically coupled systems. In: Synaptic Transmission and Neuronal Interaction, edited by M. V. L. Bennett New York: Raven, 1974, p. 153–178.
 31. Bennett, M. V. L., E. Aljure, Y. Nakajima, and G. D. Pappas. Electrotonic junctions between teleost spinal neurons: electrophysiology and ultrastructure. Science 141: 262–264, 1963.
 32. Bennett, M. V. L., and A. A. Auerbach. Calculation of electrical coupling of cells separated by a gap. Anat. Record 163: 152, 1969.
 33. Bennett, M. V. L., S. M. Crain, and H. Grundfest. Electrophysiology of supramedullary neurons in Spheroides maculatus. III. Organization of the supramedullary neurons. J. Gen. Physiol. 43: 221–250, 1959.
 34. Bennett, M. V. L., N. Feder, T. S. Reese, and W. Stewart. Movement during fixation of peroxidases injected into the crayfish septate axon. J. Gen. Physiol. 61: 254–255, 1973.
 35. Bennett, M. V. L., A. R. Freeman, and P. Thaddeus. “Reversal” of postsynaptic potentials in non‐isopotential systems. Ann. Meeting Biophysical Soc., 10th, Abstr. FD11, 1966.
 36. Bennett, M. V. L., and N. B. Gilula. Membranes and junctions in developing Fundulus embryos, freeze fractures and electrophysiology. J. Cell Biol. 63: 21a, 1974.
 37. Bennett, M. V. L., and D. A. Goodenough. Electrotonic junctions. Neurosci. Res. Program Bull. In press.
 38. Bennett, M. V. L., B. Hille, and S. Obara. Voltage threshold in excitable cells depends on stimulus form. J. Neurophysiol. 33: 585–594, 1970.
 39. Bennett, M. V. L., Y. Nakajima, and G. D. Pappas. Physiology and ultrastructure of electrotonic junctions. I. Supramedullary neurons. J. Neurophysiol. 30: 161–179, 1967.
 40. Bennett, M. V. L., Y. Nakajima, and G. D. Pappas. Physiology and ultrastructure of electrotonic junctions. III. Giant electromotor neurons of Malapterurus electricus. J. Neurophysiol. 30: 209–235, 1967.
 41. Bennett, M. V. L., G. D. Pappas, E. Aljure, and Y. Nakajima. Physiology and ultrastructure of electrotonic junctions. II. Spinal and medullary electromotor nuclei in Mormyrid fish. J. Neurophysiol. 30: 180–208, 1967.
 42. Bennett, M. V. L., G. D. Pappas, M. Giménez, and Y. Nakajima. Physiology and ultrastructure of electrotonic junctions. IV. Medullary electromotor nuclei in gymnotid fish. J. Neurophysiol. 30: 236–300, 1967.
 43. Bennett, M. V. L., M. E. Spira, and G. D. Pappas. Properties of electrotonic junctions between embryonic cells of Fundulus. Develop. Biol. 29: 419–435, 1972.
 44. Bennett, M. V. L., and J. P. Trinkaus. Electrical coupling between embryonic cells by way of extracellular space and specialized junctions. J. Cell Biol. 44: 592–610, 1970.
 45. Bennett, M. V. L., M. Würzel, and H. Grundfest. The electrophysiology of electric organs of marine electric fishes. I. Properties of electroplaques of Torpedo nobiliana. J. Gen. Physiol. 44: 757–804, 1961.
 46. Berry, M. S., and G. A. Cottrell. Excitatory, inhibitory and biphasic synaptic potentials mediated by an identified dopamine‐containing neuron. J. Physiol. London 244: 589–612, 1975.
 47. Blackshaw, S. E., and A. E. Warner. Low resistance junctions between mesoderm cells during development of trunk muscles. J. Physiol. London 245: 209–230, 1976.
 48. Blankenship, J. E., H. Wachtel, and E. R. Kandel. Ionic mechanisms of excitatory, inhibitory and dual synaptic actions mediated by an identified interneuron in the abdominal ganglion of Aplysia. J. Neurophysiol. 34: 76–92, 1971.
 49. Bodian, D. The structure of the vertebrate synapse. A study of the axon endings on Mauthner's cell and neighboring centers in the goldfish. J. Comp. Neurol. 68: 117–159, 1938.
 50. Borsellino, A., R. E. Poppele, and C. A. Terzuolo. Transfer functions of the slowly adapting stretch receptor organ of crustacea. Cold Spring Harbor Symp. Quant. Biol. 30: 581–586, 1965.
 51. Brazier, M. A. B. The historical development of neurophysiology. In: Handbook of Physiology. Neurophysiology, edited by John Field. Washington, D.C.: Am. Physiol. Soc, sect. 1, vol. I, 1959, p. 1–58.
 52. Bremer, F. Le tétanos strychnique et le mécanisme de la synchronisation neuronique. Arch. Intern. Physiol. 51: 51–84, 1941.
 53. Brightman, M. W., and T. S. Reese. Junctions between intimately apposed cell membranes in the vertebrate brain. J. Cell Biol. 40: 648–677, 1969.
 54. Brown, M. C., and P. B. C. Matthews. The effect of a muscle twitch on the back response of its motor nerve fibers. J. Physiol. London 150: 332–346, 1960.
 55. Bullock, T. H. The invertebrate neuron junction. Cold Spring Harbor Symp. Quant. Biol. 17: 267–280, 1952.
 56. Bullock, T. H., and G. A. Horridge. Structure and Function in the Nervous Systems of Invertebrates. San Francisco: Freeman, 1965.
 57. Burke, W., and B. L. Ginsborg. The action of the neuromuscular transmitter on the slow fiber membrane. J. Physiol. London 132: 599–610, 1956.
 58. Byzov, A. L., and Y. A. Trifonov. The response to electric stimulation of horizontal cells in the carp retina. Vision Res. 8: 817–822, 1968.
 59. Calvin, W. H. Generation of spike trains in CNS neurons. Brain Res. 84: 1–22, 1975.
 60. Calvin, W. H., and C. F. Stevens. Synaptic noise and other sources of randomness in motoneuron interspike intervals. J. Neurophysiol. 31: 574–587, 1968.
 61. Cantino, D., and E. Mugnaini. The structural basis for electrotonic coupling in the avian ciliary ganglion. J. Neurocytol. 4: 505–536, 1975.
 62. Cass, A., A. Finkelstein, and V. Krespi. The ion permeability induced in thin lipid membranes by the polyene antibiotics nystatin and amphotericin B. J. Gen. Physiol. 56: 100–124, 1970.
 63. Cavoto, F. V., and B. A. Flaxman. Low‐resistance pathways between mitotic and interphase epidermal cells in vitro. J. Cell Biol. 58: 223–225, 1973.
 64. Cervetto, L., and M. Piccolino. Synaptic transmission between photoreceptor and horizontal cells in the turtle retina. Science 183: 417–419, 1974.
 65. Claude, P., and D. A. Goodenough. Fracture faces of zonulae occludentes from “tight” and “leaky” epithelia. J. Cell Biol. 58: 390–400, 1973.
 66. Clusin, W., D. C. Spray, and M. V. L. Bennett. Activation of a voltage insensitive conductance by inward calcium current. Nature 256: 425–427, 1975.
 67. Coggeshall, R. E. Gap junctions between identified glial cells in the leech. J. Neurobiol. 5: 463–468, 1974.
 68. Cohen, M. I. Discharge patterns of brain‐stem respiratory neurons in relation to carbon dioxide tension. J. Neurophysiol. 31: 142–165, 1968.
 69. Cohen, M. I. Synchronization of discharge, spontaneous and evoked, between inspiratory neurons. Acta Neurobiol. Exptl. 33: 189–218, 1973.
 70. Connor, J. A., and C. F. Stevens. Inward and delayed outward membrane currents in isolated neural somata under voltage clamp. J. Physiol. London 213: 1–19, 1971.
 71. Connor, J. A., and C. F. Stevens. Prediction of repetitive firing behavior from voltage clamp data on an isolated neuron soma. J. Physiol. London 213: 31–53, 1971.
 72. Cooke, J. D., and D. M. J. Quastel. Transmitter release by mammalian motor nerve terminals in response to focal polarization. J. Physiol. London 228: 377–405, 1973.
 73. Cox, R. P., M. R. Krauss, M. E. Balis, and J. Dancis. Metabolic cooperation in cell culture. In: Cell Communication, edited by R. P. Cox. New York: Wiley, 1974, p. 67–95.
 74. Curtis, D. R., and R. W. Ryall. The acetylcholine receptors of Renshaw cells. Exptl. Brain Res. 2: 66–80, 1966.
 75. Decker, R. S. Hormal regulation of gap junction differentiation. J. Cell Biol. 69: 669–684, 1976.
 76. Decker, R. S., and D. S. Friend. Assembly of gap junctions during amphibian neurulation. J. Cell Biol. 62: 32–47, 1974.
 77. Del Castillo, J., and B. Katz. The membrane change produced by the neuromuscular transmitter. J. Physiol. London 125: 546–565, 1954.
 78. Délèze, J. The recovery of resting potential and input resistance in sheep heart injured by knife or laser. J. Physiol. London 208: 547–562, 1970.
 79. De Mello, W. C., G. E. Motta, and M. Chapeau. A study on the healing‐over of myocardial cells of toads. Circulation Res. 24: 475–487, 1964.
 80. Dennis, M. J., A. J. Harris, and S. W. Kuffler. Synaptic transmission and its duplication by focally applied acetylcholine in parasympathetic neurons in the heart of the frog. Proc. Roy. Soc. London Ser. B 177: 509–539, 1971.
 81. Deschěnes, M., and M. V. L. Bennett. A qualification to the use of TEA as a tracer for monosynaptic pathways. Brain Res. 77: 169–172, 1974.
 82. Dewey, M. M., and L. Barr. A study of structure and distribution of the nexus. J. Cell Biol. 23: 553–585, 1964.
 83. Diamond, J. The Mauthner cell. In: Fish Physiology, edited by W. S. Hoar and D. J. Randall. New York: Academic, 1971, vol. 5, p. 265–346.
 84. Diamond, J. M. Tight and leaky junctions of epithelia. Federation Proc. 33: 2220–2224, 1974.
 85. DiCaprio, R. A., A. S. French, and E. J. Sanders. Dynamic properties of electrotonic coupling between cells of early Xenopus embryos. Biophys. J. 14: 387–411, 1974.
 86. Dixon, J. S., and J. R. Cronly‐Dillon. The fine structure of the developing retina in Xenopus laevis. J. Embryol. Exptl. Morphol. 28: 659–666, 1972.
 87. Dowling, J. E. Synaptic organization of the frog retina: an electron microscopic analysis comparing the retinas of frogs and primates. Proc. Roy. Soc. London Ser. B 170: 205–228, 1968.
 88. Dowling, J. E., and B. B. Boycott. Neural connections of the retina: fine structure of the inner plexiform layer. Cold Spring Harbor Symp. Quant. Biol. 30: 393–402, 1965.
 89. Dowling, J. E., and H. Ripps. Effect of magnesium on horizontal cell activity in the skate retina. Nature 242: 101–103, 1973.
 90. Dreifuss, J. J., L. Girardier, and W. G. Forssmann. Etude de la propagation de l'excitation dans le ventricule de rat au moyen de solutions hypertoniques. Pfluegers Arch. European J. Physiol. 292: 13–33, 1966.
 91. Eccles, J. C. The Physiology of Nerve Cells. Baltimore: Johns Hopkins Press, 1957.
 92. Eccles, J. C. The Physiology of Synapses. Berlin: Springer Verlag, 1964.
 93. Eckert, R., and H. D. Lux. A non‐inactivating inward current recorded during small depolarizing voltage steps in snail pacemaker neurons. Brain Res. 83: 486–489, 1975.
 94. Elliot, T. R. On the action of adrenalin. J. Physiol. London 31: xx–xxi, 1904.
 95. Elliot, T. R. The action of adrenalin. J. Physiol. London 32: 401–467, 1905.
 96. Engberg, I., and K. C. Marshall. Mechanism of noradrenaline hyperpolarization in spinal cord motoneurones of the cat. Acta Physiol. Scand. 83: 142–144, 1971.
 97. Epstein, M., and J. Sheridan. Formation of low‐resistance junctions in the absence of protein synthesis and metabolic energy production. J. Cell Biol. 63: 95a, 1974.
 98. Evans, D. H. L., and E. M. Evans. The membrane relationships of smooth muscles: an electron microscopic study. J. Anat. 98: 37–46, 1964.
 99. Fain, G. L., G. H. Gold, and J. E. Dowling. Receptor coupling in the toad retina. Cold Spring Harbor Symp. Quant. Biol. 40: 547–561, 1975.
 100. Farquhar, M. G., and G. E. Palade. Junctional complexes in various epithelia. J. Cell Biol. 17: 375–412, 1963.
 101. Farquhar, M. G., and G. E. Palade. Cell junctions in amphibian skin. J. Cell Biol. 26: 263–291, 1965.
 102. Fatt, P. Biophysics of junctional transmission. Physiol. Rev. 34: 674–710, 1954.
 103. Fischbach, G. D., M. P. Henkart, S. A. Cohen, A. C. Breuer, J. Whysner, and F. M. Neal. Studies on the development of neuromuscular junctions in cell culture. In: Synaptic Transmission and Neuronal Interaction. New York: Raven, 1974.
 104. Fitzhugh, R. Thresholds and plateaus in the Hodgkin‐Huxley nerve equations. J. Gen. Physiol. 43: 867–896, 1960.
 105. Flaxman, B. A., and F. V. Cavoto. Low‐resistance junctions in epithelial outgrowths from normal and cancerous epidermis in vitro. J. Cell Biol. 58: 219–223, 1973.
 106. Frank, K., and M. G. F. Fuortes. Stimulation of spinal motoneurones with intracellular electrodes. J. Physiol. London 134: 451–470, 1956.
 107. Frank, K., and J. M. Sprague. Direct contralateral inhibition in the lower sacral spinal cord. Exptl. Neurol. 1: 28–43, 1959.
 108. Frankenhaeuser, B., and A. L. Hodgkin. The after‐effects of impulses in the giant nerve fibres of Loligo. J. Physiol. London 131: 341–376, 1956.
 109. Frankenhaeuser, B., and A. F. Huxley. The action potential in the myelinated nerve fiber of Xenopus laevis as computed on the basis of voltage clamp data. J. Physiol. London 171: 302–315, 1964.
 110. Friend, D. S., and D. W. Fawcett. Membrane differentiations in freeze fractured mammalian sperm. J. Cell Biol. 63: 641–664, 1974.
 111. Frömter, E., and J. Diamond. Route of passive ion permeation in epithelia. Nature New Biol. 235: 9–13, 1972.
 112. Fuortes, M. G. F., K. Frank, and M. C. Becker. Steps in the production of motoneuron spikes. J. Gen. Physiol. 40: 735–752, 1957.
 113. Fuortes, M. G. F., and F. Mantegazzini. Interpretation of the repetitive firing of nerve cells. J. Gen. Physiol. 45: 1163–1179, 1961.
 114. Furshpan, E. J. “Electrical transmission” at an excitatory synapse in a vertebrate brain. Science 144: 878–880, 1964.
 115. Furshpan, E. J., and T. Furukawa. Intracellular and extracellular responses of the several regions of the Mauthner cell of the goldfish. J. Neurophysiol. 25: 732–771, 1962.
 116. Furshpan, E. J., and D. D. Potter. Transmission at the giant motor synapses of the crayfish. J. Physiol. London 145: 289–325, 1959.
 117. Furshpan, E. J., and D. D. Potter. Low‐resistance junctions between cells in embryos and tissue culture. In: Current Topics in Developmental Biology, edited by A. A. Moscona and A. Monroy. New York: Academic, 1968, vol. 3, p. 95–127.
 118. Furukawa, T., and E. J. Furshpan. Two inhibitory mechanisms in the Mauthner neurons of goldfish. J. Neurophysiol. 26: 140–176, 1963.
 119. Futamachi, K. J., and T. A. Pedley. Glial cells and extracellular potassium: their relation in mammalian cortex. Brain Res., 109: 285–310, 1976.
 120. Gardner, D., and E. R. Kandel. Diphasic postsynaptic potential: a chemical synapse capable of mediating conjoint excitation and inhibition. Science 176: 675–678, 1972.
 121. Gerschenfeld, H. M. Chemical transmission in invertebrate central nervous systems and neuromuscular junctions. Physiol. Rev. 53: 1–119, 1973.
 122. Gerschenfeld, H. M., and D. Paupardin‐Tritsch. Ionic mechanisms and receptor properties underlying the responses of molluscan neurones to 5‐hydroxtryptamine. J. Physiol. London 243: 427–456, 1974.
 123. Gerschenfeld, H. M., and D. Paupardin‐Tritsch. On the transmitter function of 5‐hydroxytryptamine at excitatory and inhibitory synaptic junctions. J. Physiol. London 243: 457–482, 1974.
 124. Getting, P., and A. O. D. Willows. Modification of neuron properties by electrotonic synapses. II. Burst formation by electrotonic synapses. J. Neurophysiol. 37: 858–868, 1974.
 125. Gilula, N. B. Junctions between cells. In: Cell Communication, edited by R. P. Cox. New York: Wiley, 1974, p. 1–29.
 126. Gilula, N. B., O. R. Reeves, and A. Steinbach. Metabolic coupling, ionic coupling and cell contacts. Nature 235: 262–265, 1972.
 127. Gilula, N. B., and P. Satir. Septate and gap junctions in molluscan gill epithelium. J. Cell Biol. 51: 869–872, 1971.
 128. Globus, A., H. D. Lux, and P. Schubert. Transfer of amino acids between neuroglial cells and neurons in the leech ganglion. Exptl. Neurol. 40: 104–113, 1973.
 129. Globus, A., P. Schubert, and H. D. Lux. Somatodendritic spread of intracellularly injected tritiated glycine in cat spinal motoneurones. Brain Res. 11: 440–445, 1968.
 130. Gogan, P., J. P. Gueritaud, G. Horscholle‐Bossavit, and S. Tyc‐Dumon. Electrotonic coupling between motoneurons in the abducens nucleus of the cat. Exptl. Brain Res. 21: 139–154, 1974.
 131. Goldstein, S. S., and W. Rall. Changes of action potential shape and velocity for changing core conductor geometry. Biophys. J. 14: 731–757, 1974.
 132. Goodenough, D. A. Bulk isolation of mouse hepatocyte gap junctions. Characterization of the principal protein, connexin. J. Cell Biol. 61: 557–563, 1974.
 133. Goodenough, D. A., D. L. D. Caspar, L. Makowski, and W. O. Phillips. X‐ray diffraction of isolated gap junctions. J. Cell Biol. 63: 114a, 1974.
 134. Goodenough, D. A., and N. B. Gilula. The splitting of hepatocyte gap junctions and zonulae occludentes with hypertonic disaccharides. J. Cell Biol. 61: 575–590, 1974.
 135. Goodenough, D. A., and J. P. Revel. The permeability of isolated and in situ mouse hepatic gap junctions studied with enzymatic tracers. J. Cell Biol. 50: 81–91, 1971.
 136. Goodenough, D. A., and W. Stoeckenius. The isolation of mouse hepatocyte gap junctions. J. Cell Biol. 54: 646–656, 1972.
 137. Gorman, A. L. F., and M. F. Marmor. Contributions of the sodium pump and ionic gradients to the membrane potential of a molluscan neurone. J. Physiol. London 210: 897–917, 1970.
 138. Goshima, K. Formation of nexuses and electrotonic transmission between myocardial and FL cells in monolayer culture. Exptl. Cell Res. 63: 124–130, 1970.
 139. Granit, R., and C. R. Skoglund. Facilitation, inhibition and depression at the “artificial synapse” formed by the cut end of a mammalian nerve. J. Physiol. London 103: 435–448, 1945.
 140. Grinnell, A. D. Electrical interaction between antidromically stimulated frog motoneurones and dorsal root afferents: enhancement by gallamine and TEA. J. Physiol. London 210: 17–43, 1970.
 141. Grobstein, C. Cell contact in relation to embryonic induction. Exptl. Cell Res. Suppl. 8: 234–245, 1961.
 142. Grossman, Y., M. E. Spira, and I. Parnas. Differential flow of information into branches of a single axon. Brain Res. 64: 379–386, 1973.
 143. Grundfest, H. Synaptic and ephaptic transmission. In: Handbook of Physiology. Neurophysiology, edited by J. Field. Washington, D.C.: Am. Physiol. Soc, 1959, sect. 1, vol. I, p. 147–197.
 144. Grundfest, H., and J. Magnes. Excitability changes in dorsal roots produced by electrotonic effects from adjacent afferent activity. Am. J. Physiol. 164: 502–508, 1951.
 145. Hagins, W., R. D. Penn, and S. Yoshikami. Dark current and photocurrent in retinal rods. Biophys. J. 10: 380–412, 1970.
 146. Hagiwara, S., and H. Morita. Electrotonic transmission between two nerve cells in leech ganglion. J. Neurophysiol. 25: 721–731, 1962.
 147. Hagiwara, S., and I. Tasaki. A study on the mechanism of impulse transmission across the giant synapse of the squid. J. Physiol. London 143: 114–137, 1958.
 148. Hagiwara, S., A. Watanabe, and N. Saito. Potential changes in syncytial neurons of lobster cardiac ganglion. J. Neurophysiol. 22: 554–572, 1959.
 149. Hama, K. Some observations on the fine structure of the giant nerve fibers of the earthworm, Eisenia foetida. J. Biophys. Biochem. Cytol. 6: 61–66, 1959.
 150. Hama, K. Some observations on the fine structure of the giant fibers of the crayfishes (Cambarus virilus and Cambarus clarkii) with special reference to the submicroscopic organization of the synapses. Anat. Record 141: 275–294, 1961.
 151. Hamilton, D. W. The calyceal synapse of Type I vestibular hair cells. J. Ultrastruct. Res. 23: 98–114, 1968.
 152. Hand, A. R., and S. Gopel. The structural organization of the septate and gap junctions in Hydra. J. Cell Biol. 52: 397–408, 1972.
 153. Hartline, H. K., and F. Ratliff. Inhibitory interaction of receptor units in the eye of Limulus. J. Gen. Physiol. 40: 357–376, 1957.
 154. Hatt, H., and D. O. Smith. Axon conduction block: differential channeling of nerve impulses in the crayfish. Brain Res. 87: 85–88, 1975.
 155. Hax, W. M. A., G. E. P. M. Van Venrooij, and J. B. J. Vossenberg. Cell communication: a cyclic‐AMP mediated phenomenon. J. Membrane Biol. 19: 253–266, 1974.
 156. Heppner, D. B., and R. Plonsey. Simulation of electrical interaction of cardiac cells. Biophys. J. 10: 1057–1075, 1970.
 157. Hille, B. Ionic channels in nerve membranes. Progr. Biophys. Mol. Biol. 21: 1–32, 1970.
 158. Hinrichsen, C. F. L., and L. M. H. Larramendi. Synapses and cluster formation of the mouse mesencephalic fifth nucleus. Brain Res. 7: 296–299, 1968.
 159. Howell, W. H. Vagus inhibition of the heart in its relation to the inorganic salts of the blood. Am. J. Physiol. 15: 280–294, 1906.
 160. Hubbard, J. I. Microphysiology of vertebrate neuromuscular transmission. Physiol. Rev. 53: 674–723, 1973.
 161. Hubbard, J. I., and R. F. Schmidt. An electrophysiological investigation of mammalian motor nerve terminals. J. Physiol. London 166: 145–167, 1963.
 162. Hunt, R. K., and M. Jacobson. Neuronal specificity revisited. Current Topics Develop. Biol. 8: 203–259, 1974.
 163. Hyde, A., B. Blondel, A. Matter, J. P. Cheneval, B. Fillous, and L. Girardier. Homo‐ and heterocellular junctions in cell cultures: an electrophysiological and morphological study. Progr. Brain Res. 31: 283–311, 1969.
 164. Imanaga, I. Cell to cell diffusion of Procion yellow in sheep and calf Purkinje fibres. J. Membrane Biol. 16: 381–388, 1974.
 165. Ishii, Y., S. Matsuura, and T. Furukawa. An input‐output relation at the synapse between hair cells and eighth nerve fibers in goldfish. Japan. J. Physiol. 21: 91–98, 1971.
 166. Ito, S., and W. R. Loewenstein. Ionic communication between early embryonic cells. Develop. Biol. 19: 228–243, 1969.
 167. Ito, S., E. Sato, and W. R. Loewenstein. Studies on the formation of a permeable cell membrane junction. II. Evolving junctional conductance and junctional insulation. J. Membrane Biol. 19: 339–355, 1974.
 168. Jaffe, L. F., and R. Nuccitelli. An ultrasensitive vibrating probe for measuring steady extracellular currents. J. Cell Biol. 63: 614–628, 1974.
 169. Jansen, J. K. S., and J. G. Nicholls. Conductance changes, an electrogenic pump and the hyperpolarization of leech neurons following impulses. J. Physiol. London 229: 635–655, 1973.
 170. Johnels, A. G. On the origin of the electric organ in Malapterurus electricus. Quart. J. Microscop. Sci. 97: 455–464, 1956.
 171. Johnson, R., M. Hammer, J. Sheridan, and J.‐P. Revel. Gap junction formation between reaggregated Novikoff hepatoma cells. Proc. Natl. Acad. Sci. US 71: 4536–4540, 1974.
 172. Johnson, R. G., W. Herman, and D. Preus. Homocellular and heterocellular gap junctions in Limulus. J. Ultrastruct. Res. 43: 298–312, 1973.
 173. Johnson, R. G., and J. D. Sheridan. Junctions between cancer cells in culture: ultrastructure and permeability. Science 174: 717–719, 1971.
 174. Jongsma, H. J., and H. E. van Run. Electrotonic spread of current in monolayer cultures of neonatal rat heart cells. J. Membrane Biol. 9: 341–360, 1972.
 175. Junge, D., and C. L. Stevens. Cyclic variation of potassium conductance in a burst‐generating neurone in Aplysia. J. Physiol. London 235: 155–181, 1973.
 176. Kandel, E. R., and D. Gardner. The synaptic actions mediated by the different branches of a single neuron. Res. Publ. Assoc. Res. Nervous Mental Diseases 50: 91–144, 1972.
 177. Kandel, E. R., W. A. Spencer, and F. J. Brinley, Jr. Electrophysiology of hippocampal neurons. I. Sequential invasion and synaptic organization. J. Neurophysiol. 24: 225–242, 1961.
 178. Kaneko, A. Electrical connections between horizontal cells in the dogfish retina. J. Physiol. London 213: 95–105, 1971.
 179. Karlsson, U., and R. L. Schultz. Fixation of the central nervous system for electron microscopy by aldehyde perfusion. J. Ultrastruct. Res. 12: 160–206, 1965.
 180. Katz, B., and R. Miledi. Propagation of electric activity in motor nerve terminals. Proc. Roy. Soc. London Ser. B 161: 433–482, 1965.
 181. Katz, B., and R. Miledi. The measurement of synaptic delay, and the time course of acetylcholine release at the neuromuscular junction. Proc. Roy. Soc. London Ser. B 161: 483–495, 1965.
 182. Katz, B., and R. Miledi. The effect of temperature on the synaptic delay at the neuromuscular junction. J. Physiol. London 181: 656–670, 1965.
 183. Katz, B., and R. Miledi. The release of acetylcholine from nerve endings by graded electric pulses. Proc. Roy. Soc. London Ser. B 167: 23–28, 1967.
 184. Katz, B., and R. Miledi. Spontaneous and evoked activity of motor nerve endings in calcium Ringer. J. Physiol. London 203: 689–706, 1969.
 185. Katz, B., and R. Miledi. The statistical nature of the acetylcholine potential and its molecular components. J. Physiol. London 224: 665–699, 1972.
 186. Katz, B., and O. H. Schmitt. Electric interaction between two adjacent nerve fibres. J. Physiol. London 97: 471–488, 1940.
 187. Katz, B., and S. Thesleff. A study of the ‘desensitization’ produced by acetylcholine at the motor end‐plate. J. Physiol. London 138: 63–80, 1957.
 188. Keeter, J. S., M. Deschěnes, G. D. Pappas, and M. V. L. Bennett. Fine structure and permeability studies of a rectifying electrotonic synapse (Abstract). Biol. Bull. 147: 485–486, 1974.
 189. Kehoe, J. S. Ionic mechanism of a two‐component cholinergic inhibition in Aplysia neurons. J. Physiol. London 225: 85–114, 1972.
 190. Kelly, A. M., and S. J. Zacks. The fine structure of motor endplate morphogenesis. J. Cell Biol. 42: 154–169, 1969.
 191. Kennedy, D., A. I. Selverston, and M. P. Remler. Analysis of restricted neural networks. Science 164: 1488–1496, 1969.
 192. King, J. L., and T. H. Jukes. Non‐Darwinian evolution. Science 164: 788–798, 1969.
 193. Kohno, K. Symmetrical axo‐axonic synapses in the axon cap of the goldfish Mauthner cell. Brain Res. 23: 255–258, 1970.
 194. Kolb, H., and E. V. Famiglietti. Rod and cone pathways in the inner plexiform layer of cat retina. Science 186: 47–49, 1974.
 195. Korn, H., and M. V. L. Bennett. Vestibular nystagmus and teleost oculomotor neurons: functions of electrotonic coupling and dendritic impulse initiation. J. Neurophysiol. 38: 430–451, 1975.
 196. Korn, H., and D. S. Faber. An electrically mediated inhibition in goldfish medulla. J. Neurophysiol. 38: 452–471, 1975.
 197. Korn, H., C. Sotelo, and F. Crepel. Electrotonic coupling between neurons in the rat lateral vestibular nucleus. Exptl. Brain Res. 16: 225–275, 1973.
 198. Kriebel, M. E. Electrical characteristics of tunicate heart cell membranes and nexuses. J. Gen. Physiol. 52: 46–59, 1968.
 199. Kriebel, M. E., M. V. L. Bennett, S. G. Waxman, and G. D. Pappas. Oculomotor neurons in fish: electrotonic coupling and multiple sites of impulse initiation. Science 166: 520–524, 1969.
 200. Krnjević, K. Chemical nature of synaptic transmission in vertebrates. Physiol. Rev. 54: 418–540, 1974.
 201. Krnjević, K., and A. Lisiewiscz. Injections of calcium ions into spinal motoneurones. J. Physiol. London 225: 363–390, 1972.
 202. Krnjević, K., R. Pumain, and L. Renaud. The mechanism of excitation by acetylcholine in the cerebral cortex. J. Physiol. London 215: 247–268. 1971.
 203. Kuffler, S. W., and C. Edwards. Mechanism of gamma aminobutyric acid (GABA) action and its relation to synaptic inhibition. J. Neurophysiol. 21: 589–610, 1958.
 204. Kuffler, S. W. and J. G. Nicholls. The physiology of neuroglial cells. Ergeb. Physiol. Biol. Chem. Exptl. Pharmakol. 57: 1–90, 1966.
 205. Kuffler, S. W., and D. D. Potter. Glia in the leech central nervous system. Physiological properties and neuron‐glia relationship. J. Neurophysiol. 27: 290–320, 1964.
 206. Kuffler, S. W., and D. Yoshikami. The distribution of acetylcholine sensitivity at the post‐synaptic membrane of vertebrate skeletal twitch muscles: iontophoretic mapping in the micron range. J. Physiol. London 244: 703–730, 1975.
 207. Kuno, M., and R. Llinás. Alterations of synaptic action in chromatolysed motoneurones of the cat. J. Physiol. London 210: 823–838, 1970.
 208. Kusano, K., and H. Grundfest. Circus reexcitation as a cause of repetitive activity in crayfish lateral giant axons. J. Cell Comp. Physiol. 65: 325–336, 1965.
 209. Kusano, K., and M. M. LaVail. Impulse conduction in the shrimp medullated giant fiber with special reference to the structure of functionally excitable areas. J. Comp. Neurol. 142: 481–494, 1971.
 210. Kusano, K., R. Miledi, and J. Stinnakre. Postsynaptic entry of calcium induced by transmitter action. Proc. Roy. Soc. London Ser. B 189: 49–56, 1975.
 211. Landmesser, L., and G. Pilar. The onset and development of transmission in the chick ciliary ganglion. J. Physiol. London 222: 691–713, 1972.
 212. Lane, B. P., and J. A. G. Rhodin. Cellular interrelationships and electrical activity in two types of smooth muscle. J. Ultrastruct. Res. 10: 470–488, 1964.
 213. Lasansky, A. Cell junctions in ommatidia of Limulus. J. Cell Biol. 33: 365–383, 1967.
 214. Lasek, R. J., H. Gainer, and R. J. Przybylski. Transfer of newly synthesized proteins from Schwann cells to the squid giant axon. Proc. Natl. Acad. Sci. US 71: 1188–1192, 1974.
 215. LaVail, J. H., and M. M. LaVail. Retrograde axonal transport in the central nervous system. Science 176: 1416–1417, 1972.
 216. Lentz, T. L., and J. P. Trinkaus. Differentiation of the junctional complex of surface cells in the developing Fundulus blastoderm. J. Cell Biol. 48: 455–472, 1971.
 217. Levitan, H. and L. Tauc. Aectylcholine receptors: topographic distribution and pharmacological properties of two receptor types on a single molluscan neuron. J. Physiol. London 222: 537–558, 1972.
 218. Levitan, H., and L. Tauc. Polyphasic synaptic potentials in the ganglion of the mollusc, Navanax. J. Physiol. London 248: 35–44, 1975.
 219. Lewontin, R. C. The Genetic Basis of Evolutionary Change. New York: Columbia Univ. Press, 1974.
 220. Llinás, R. Motor aspects of cerebellar control. Physiologist 17: 19–46, 1974.
 221. Llinás, R. Electrical synaptic transmission in the mammalian central nervous system. In: Golgi Centennial Symposium: Perspectives in Neurobiology, edited by M. Santini. New York: Raven, 1975, p. 379–386.
 222. Llinás, R., R. Baker, and C. Sotelo. Electrotonic coupling between neurons in cat inferior olive. J. Neurophysiol. 37: 560–571, 1971.
 223. Loewenstein, W. R., and Y. Kanno. Intercellular communication and tissue growth. I. Cancerous growth. J. Cell Biol. 33: 225–234, 1967.
 224. Loewenstein, W. R., and R. D. Penn. Intercellular communication and tissue growth. II. Tissue regeneration. J. Cell Biol. 33: 235–242, 1967.
 225. Loewi, O. The humoral transmission of nervous impulse. Harvey Lectures Ser. 218–233, 1932–1933.
 226. LoPresti, V., E. R. Macagno, and C. Levinthal. Structure and development of neuronal connections in isogenic organism: transient gap junctions between growing optic axons and lamina neuroblasts. Proc. Natl. Acad. Sci. US 71: 1098–1102, 1974.
 227. Machemer, H. Ciliary activity and the origin of metachrony in Paramecium: the effects of increased viscosity. J. Exptl. Biol. 57: 239–259, 1972.
 228. Magleby, K. L., and C. F. Stevens. The effect of voltage on the time course of end‐plate currents. J. Physiol. London 223: 151–171, 1972.
 229. Markin, V. S. Electrical interaction of parallel nonmyelinated nerve fibers. III. Interaction in bundles. Biophysics 18: 324–332, 1973.
 230. Martin, A. R., and G. Pilar. Dual mode of synaptic transmission in the avian ciliary ganglion. J. Physiol. London 168: 443–463, 1963.
 231. Martin, A. R., and G. Pilar. Transmission through the ciliary ganglion of the chick. J. Physiol. London 168: 464–475, 1963.
 232. Martin, A. R., and G. Pilar. An analysis of electrical coupling in the avian ciliary ganglion. J. Physiol. London 171: 454–475, 1964.
 233. McNutt, N. S., and R. S. Weinstein. The ultrastructure of the nexus. A correlated thin section and freeze cleave study. J. Cell Biol. 47: 666–688, 1970.
 234. Meech, R. W., and N. B. Standen. Potassium activation in Helix aspersa neurones under voltage clamp: a component mediated by calcium influx. J. Physiol. London 249: 211–239, 1975.
 235. Meier, S., and E. D. Hay. Stimulation of corneal differentiation by interaction between cell surface and extracellular matrix. I. Morphometric analysis of transfilter induction. J. Cell Biol. 66: 275–291, 1975.
 236. Meszler, R. M., G. D. Pappas, and M. V. L. Bennett. Morphological demonstration of electrotonic coupling of neurons by way of presynaptic fibers. Brain Res. 37: 412–415, 1972.
 237. Meszler, R. M., G. D. Pappas, and M. V. L. Bennett. Morphology of the electromotor system in the spinal cord of the electric eel, Electrophorus electricus. J. Neurocytol. 3: 251–261, 1974.
 238. Michalke, W., and W. R. Loewenstein. Communication between cells of different type. Nature 232: 121–122, 1971.
 239. Miledi, R. Spontaneous synaptic potentials and quantal release of transmitter in the stellate ganglion of the squid. J. Physiol. London 192: 379–406, 1967.
 240. Model, P. G., M. E. Spira, and M. V. L. Bennett. Synaptic inputs to cell bodies of the giant fibers of the hatchetfish. Brain Res. 45: 288–295, 1972.
 241. Mugnaini, E., F. Walburg, and E. Hauglie‐Hansen. Observations on the fine structure of the lateral vestibular nucleus (Deiters' nucleus) in the cat. Exptl. Brain Res. 4: 146–186, 1967.
 242. Muir, A. R. The effect of divalent cations on the ultrastructure of the perfused rat heart. J. Anat. 101: 239–261, 1967.
 243. Mulloney, B. Structure of the giant fibers of earthworms. Science 168: 994–996, 1970.
 244. Mulloney, B., and A. I. Selverston. Organization of the stomatogastric ganglion of the spiny lobster. I. Neurons driving the lateral teeth. J. Comp. Physiol. 91: 1–32, 1974.
 245. Nakajima, S. Analysis of K inactivation and TEA action in the supramedullary cells of puffer. J. Gen. Physiol. 49: 629–640, 1966.
 246. Nakajima, Y. Fine structure of the synaptic endings on the Mauthner cell of the goldfish. J. Comp. Neurol. 156: 375–402, 1974.
 247. Nakajima, S., and K. Kusano. Behavior of delayed current under voltage clamp in supramedullary neurons of puffer. J. Gen. Physiol. 49: 613–628, 1966.
 248. Nastuk, W. L., and R. L. Parsons. Factors in the inactivation of postjunctional membrane receptors of frog skeletal muscle. J. Gen. Physiol. 56: 218–249, 1970.
 249. Neher, E. Two fast transient current components during voltage clamp on snail neurons. J. Gen. Physiol. 58: 36–53, 1971.
 250. Nelson, P. G. Interaction between spinal motoneurons of the cat. J. Neurophysiol. 29: 275–287, 1966.
 251. Nicholls, J. G., and D. Purves. Monosynaptic chemical and electrical connections between sensory and motor cells in the central nervous system of the leech. J. Physiol. London 209: 647–668, 1970.
 252. Nicholls, J. G., and D. Purves. A comparison of chemical and electrical synaptic transmission between single sensory cells and a motoneurone in the central nervous system of the leech. J. Physiol. London 225: 637–656, 1972.
 253. Obara, S. Receptor cell activity at ‘rest’ with respect to the tonic operation of a specialized lateralis receptor. Proc. Japan Acad. 50: 386–391, 1974.
 254. Obara, S., and M. V. L. Bennett. Mode of operation of ampullae of Lorenzini of the skate Raja. J. Gen. Physiol. 60: 534–557, 1972.
 255. Obara, S., and Y. Oomura. Disfacilitation as the basis for the sensory suppression in a specialized lateralis receptor of the marine catfish. Proc. Japan Acad. 49: 213–217, 1973.
 256. Ogden, T. E. Intraretinal slow potentials evoked by brain stimulation in the primate. J. Neurophysiol. 29: 898–908, 1966.
 257. Ohta, T., and M. Kimura. On the constancy of the evolutionary rate of cistrons. J. Mol. Evolution 1: 18–25, 1971.
 258. O'Lague, P., H. Dalen, H. Rubin, and C. Tobias. Electrical coupling: low resistance junctions between mitotic and interphase fibroblasts in tissue culture. Science 170: 464–466, 1970.
 259. Oliveira‐Castro, G. M., and W. R. Loewenstein. Junctional membrane permeability. Effects of divalent cations. J. Membrane Biol. 5: 51–77, 1971.
 260. Paine, R. T. Food recognition and predation on opistho‐branchs by Navanax inermis. Veliger 6: 1–9, 1963.
 261. Pappas, G. D., Y. Asada, and M. V. L. Bennett. Morphological correlates of increased coupling resistance at an electrotonic synapse. J. Cell Biol. 49: 173–188, 1971.
 262. Pappas, G. D., and M. V. L. Bennett. Fine structure of two transmission systems operating toadfish motoneurons. Intern. Congr. Electron Microscopy, 6th, Kyoto, 1966, p. 429–430.
 263. Pappas, G. D., and M. V. L. Bennett. Specialized junctions involved in electrical transmission between neurons. Ann. NY Acad. Sci. 137: 495–508, 1966.
 264. Pappas, G. D., and D. P. Purpura. Distribution of colloidal particles in extracellular space and synaptic cleft substance of mammalian cerebral cortex. Nature 210: 1391–1392, 1966.
 265. Pappas, G. D., S. G. Waxman, and M. V. L. Bennett. Morphology of spinal electromotor neurons and presynaptic coupling pathways in the gymnotid Sternarchus albifrons. J. Neurocytol. 4: 469–478, 1975.
 266. Parnas, I., D. Armstrong, and F. Strumwasser. Prolonged excitatory and inhibitory synaptic modulation of a bursting pacemaker neuron. J. Neurophysiol. 37: 594–608, 1974.
 267. Parnas, I., and F. Strumwasser. Mechanisms of long‐lasting inhibition of a bursting pacemaker neuron. J. Neurophysiol. 37: 609–620, 1974.
 268. Payton, B. W., M. V. L. Bennett, and G. D. Pappas. Temperature‐dependence of resistance at an electrotonic synapse. Science 165: 594–597, 1969.
 269. Payton, B. W., M. V. L. Bennett, and G. D. Pappas. Permeability and structure of junctional membranes at an electrotonic synapse. Science 166: 1641–1643, 1969.
 270. Peracchia, C. Low resistance junctions in crayfish. I. Two arrays of globules in junctional membranes. J. Cell Biol. 57: 54–65, 1973.
 271. Peracchia, C. Low resistance junctions in crayfish. II. Structural details and further evidence for intercellular channels by freeze‐fracture and negative staining. J. Cell Biol. 57: 66–76, 1973.
 272. Pfenninger, K., and C. Rovainen. Stimulation‐ and calcium‐dependence of vesicle attachment sites in the presynaptic membrane; a freeze‐cleave study on the lamprey spinal cord. Brain Res. 72: 1–23, 1974.
 273. Pinching, A. J., and T. P. S. Powell. The neuropil of the glomeruli of the olfactory bulb. J. Cell Sci. 9: 347–377, 1971.
 274. Pitts, J. D. Molecular exchange and growth control in tissue culture. In: Growth Control in Cell Cultures, edited by G. E. W. Wolstenholme and J. Knight. London: Churchill Livingstone, 1971, p. 89–105.
 275. Pitts, J. D. Direct interaction between animal cells. In: Cell Interactions, edited by L. G. Silvestri. Amsterdam: North‐Holland, 1972, p. 277–285.
 276. Politoff, A., and G. D. Pappas. Mechanisms of increase in coupling resistance at electrotonic synapses of the crayfish septate axon. Anat. Record 172: 384–385, 1972.
 277. Politoff, A., G. D. Pappas, and M. V. L. Bennett. Cobalt: a tracer for light and electron microscopy that can cross an electrotonic synapse. J. Cell Biol. 55: 204a, 1972.
 278. Potter, D. D., E. J. Furshpan, and E. S. Lennox. Connections between cells of the developing squid as revealed by electrophysiological methods. Proc. Natl. Acad. Sci. US 55: 328–336, 1966.
 279. Prosser, C. L. Smooth muscle. Ann. Rev. Physiol. 36: 503–537, 1974.
 280. Prosser, C. L., C. L. Ralph, and W. W. Steinberger. Responses to stretch and the effect of pull on propagation in non‐striated muscle of Golfingia (=Phascolosoma) and Mustelus. J. Cellular Comp. Physiol. 54: 135–146, 1959.
 281. Pumphrey, R. J., and J. Z. Young. The rates of conduction of nerve fibres of various diameters in cephalopods. J. Exptl. Biol. 15: 453–466, 1938.
 282. Rall, W., G. M. Shepherd, T. S. Reese, and M. W. Brightman. Dendrodendritic synaptic pathway for inhibition in the olfactory bulb. Exptl. Neurol. 14: 44–56. 1966.
 283. Rash, J. E. and D. Fambrough. Ultrastructural and electrophysiological correlates of cell coupling and cytoplasmic fusion during myogenesis in vitro. Develop. Biol. 30: 166–186, 1973.
 284. Rash, J. E. and L. Staehelin. Freeze‐cleave demonstration of gap junctions between skeletal cells in vivo. Develop. Biol. 36: 455–461, 1974.
 285. Raviola, E., and N. B. Gilula. Gap junctions between photoreceptor cells in the vertebrate retina. Proc. Natl. Acad. Sci. US 70: 1677–1681, 1973.
 286. Raviola, E., and N. B. Gilula. Intramembrane organization of specialized contacts in the outer plexiform layer of the retina. J. Cell Biol. 65: 192–222, 1975.
 287. Reese, T. S., and M. J. Karnovsky. Fine structural localization of a blood brain barrier to exogenous peroxidase. J. Cell Biol. 34: 207–211, 1967.
 288. Remler, M., A. Selverston, and D. Kennedy. Lateral giant fibers of crayfish: location of somata by dye injection. Science 162: 281–283, 1968.
 289. Revel, J. P., and M. J. Karnovsky. Hexagonal array of subunits in intercellular junctions of the mouse heart and liver. J. Cell Biol. 33: C7–C12, 1967.
 290. Revel, J. P., A. G. Yee, and A. J. Hudspeth. Gap junctions between electrotonically coupled cells in tissue culture and in brown fat. Proc. Natl. Acad. Sci. US 68: 2924–2927. 1971.
 291. Rieske, E., P. Schubert, and G. W. Kreutzberg. Transfer of radioactive material between electrically coupled neurons of the leech central nervous system. Brain Res. 84: 365–382, 1975.
 292. Robertson, J. D. Recent electron microscope observations on the ultrastructure of the crayfish median‐to‐motor giant synapse. Exptl. Cell Res. 8: 226–229, 1955.
 293. Robertson, J. D. Ultrastructure of excitable membranes and the crayfish median‐giant synapse. Ann. NY Acad. Sci. 94: 339–389, 1961.
 294. Robertson, J. D. The occurrence of a subunit pattern in the unit membranes of club endings in Mauthner cell synapses in goldfish brains. J. Cell Biol. 19: 201–221, 1963.
 295. Rose, B. Junctional membrane permeability, restoration by repolarizing current. Science 169: 607–609, 1970.
 296. Rose, B. Intercellular communication and some structural aspects of membrane junctions in a simple cell system. J. Membrane Biol. 5: 1–19, 1971.
 297. Rose, B., and W. R. Loewenstein. Junctional membrane permeability. Depression by substitution of Li for extracellular Na, and by long‐term lack of Ca and Mg; restoration by cell repolarization. J. Membrane Biol. 5: 20–50, 1971.
 298. Rose, B., and W. R. Loewenstein. Permeability of cell junction depends on local cytoplasmic calcium activity. Nature 254: 250–252, 1975.
 299. Rose, B., and W. R. Loewenstein. Calcium ion distribution in cytoplasm visualized by aequorin: diffusion in cytosol restricted by energized sequestering. Science 190: 1204–1206, 1975.
 300. Rovainen, C. M. Synaptic interactions of identified nerve cells in the spinal cord of the sea lamprey. J. Comp. Neurol. 154: 189–206, 1974.
 301. Rovainen, C. M. Synaptic interactions of reticulospinal neurons and nerve cells in the spinal cord of the sea lamprey. J. Comp. Neurol. 154: 207–224, 1974.
 302. Satir, B., C. Schooley, and P. Satir. Membrane reorganization during secretion in Tetrahymena. Nature 235: 53–54, 1972.
 303. Saunders, J. W., and M. T. Gasseling. Trans‐filter propagation of apical ectoderm maintenance factor in the chick embryo wing bud. Develop. Biol. 7: 64–78, 1963.
 304. Sheridan, J. D. Low‐resistance junctions between cancer cells in various solid tumors. J. Cell Biol. 45: 91–99, 1970.
 305. Sheridan, J. D. Electrical coupling of cells and cell communication. In: Cell Communication, edited by R. P. Cox. New York: Wiley, 1974, p. 31–42.
 306. Siggins, G. R., A. P. Oliver, B. J. Hoffer, and F. E. Bloom. Cyclic adenosine‐monophosphate and epinephrine: effects on transmembrane properties of cerebellar Purkinje cells. Science 171: 192–194, 1971.
 307. Simon, E. J. Two types of luminosity horizontal cells in the retina of the turtle. J. Physiol. London 230: 199–211, 1973.
 308. Singer, S. J., and G. L. Nicholson. The fluid mosaic model of the structure of cell membranes. Science 175: 720–731, 1972.
 309. Sleigh, M. A. Coordination of the rhythm of beat in some ciliary systems. Intern. Rev. Cytol. 25: 31–54, 1969.
 310. Sloper, J. J. Gap junctions between dendrites in the primate neocortex. Brain Res. 44: 641–646, 1972.
 311. Smith, T. G., J. L. Barker, and H. Gainer. Requirements for bursting pacemaker potential activity in molluscan neurones. Nature 253: 450–452, 1975.
 312. Smith, T. G., and F. Bauman. The functional organization within the ommatidium of the lateral eye of Limulus. Progr. Brain Res. 31: 313–349, 1969.
 313. Smith, T. G., R. B. Wuerker, and K. Frank. Membrane impedance changes during synaptic transmission in cat spinal neurons. J. Neurophysiol. 30: 1072–1096, 1967.
 314. Socolar, S. J. Cell coupling in epithelia. Exptl. Eye Res. 15: 693–698, 1973.
 315. Socolar, S. J., and A. L. Politoff. Uncoupling cell junctions in a glandular epithelium by depolarizing current. Science 172: 492–494, 1971.
 316. Sotelo, C., and R. Llinás. Specialized membrane junctions between neurons in the vertebrate cerebellar cortex. J. Cell Biol. 53: 271–289, 1972.
 317. Sotelo, C., R. Llinás, and R. Baker. Structural study of inferior olivary nucleus of the cat: morphological correlates of electrotonic coupling. J. Neurophysiol. 37: 541–559, 1974.
 318. Sotelo, S., and S. Palay. The fine structure of the lateral vestibular nucleus in the rat. II. Synaptic organization. Brain Res. 18: 93–116, 1970.
 319. Sotelo, C., and J. Taxi. Ultrastructural aspects of electrotonic junctions in the spinal cord of the frog. Brain Res. 17: 137–141, 1970.
 320. Spira, A. W. The nexus in the intercalated disc of the canine heart: quantitative data for an estimation of its resistance. J. Ultrastruct. Res. 34: 409–425, 1971.
 321. Spira, M. E., and M. V. L. Bennett. Synaptic control of electrotonic coupling between neurons. Brain Res. 37: 294–300, 1971.
 322. Srivastava, C. B. L. Morphological evidence for electrical synapses of ‘gap’ junction type in another vertebrate receptor. Experientia 28: 1029–1030, 1972.
 323. Staehelin, L. A. Structure and function of intercellular junctions. Intern. Rev. Cytol. 39: 191–283, 1974.
 324. Staehelin, L. A., T. M. Mukherjee, and A. W. Williams. Freeze‐etch appearance of the tight junctions in the epithelium of small and large intestine of mice. Protoplasma 67: 165–184, 1969.
 325. Stein, R. B. The frequency of nerve action potentials generated by applied currents. Proc. Roy. Soc. London Ser. B 167: 64–86, 1967.
 326. Stein, R. B. Some models of neuronal variability. Biophys. J. 7: 37–68, 1967.
 327. Steinbach, A. B., and M. V. L. Bennett. Effects of divalent ions and drugs on synaptic transmission in phasic electroreceptors in a mormyrid fish. J. Gen. Physiol. 58: 580–598, 1971.
 328. Takeuchi, N. Some properties of conductance changes at the endplate membrane during the action of acetylcholine. J. Physiol. London 167: 128–140, 1963.
 329. Tasaki, I., and S. Hagiwara. Demonstration of two stable potential states in the squid giant axon under tetraethylammonium chloride. J. Gen. Physiol. 40: 859–885, 1957.
 330. Tauc, L. Polyphasic synaptic activity. In: Mechanisms of Synaptic Transmission, edited by K. Akert and P. G. Waser. Amsterdam: Elsevier, 1969, p. 247–257.
 331. Tauc, L. Site of origin and propagation of spike in the giant neuron of Aplysia. J. Gen. Physiol. 45: 1077–1097, 1961.
 332. Terzuolo, C. A., and T. H. Bullock. Measurement of imposed voltage gradient adequate to modulate neuronal firing. Proc. Natl. Acad. Sci. US 42: 687–694, 1956.
 333. Terzuolo, C. A., and Y. Washizu. Relation between stimulus strength, generator potential and impulse frequency in stretch receptor of crustacea. J. Neurophysiol. 25: 56–66, 1962.
 334. Thomas, R. C. Electrogenic sodium pump in nerve and muscle cells. Physiol. Rev. 52: 563–949, 1972.
 335. Tomita, T. Electrical activity of vertebrate photoreceptors. Quart. Rev. Biophys. 3: 179–222, 1970.
 336. Trautwein, W. Membrane currents in cardiac muscle fibers. Physiol. Rev. 53: 793–835, 1973.
 337. Trifonov, Y. A. Study of synaptic transmission between the photoreceptor and the horizontal cell using electrical stimulation of the retina. Biophysics 13: 948–954, 1968.
 338. Tsien, R. W. Mode of action of chronotropic agents in cardiac Purkinje fibers. J. Gen. Physiol. 64: 320–342, 1974.
 339. Van Essen, D. C. The contribution of membrane hyperpolarization to adaptation and conduction block in sensory neurones of the leech. J. Physiol. London 230: 509–534, 1973.
 340. Venter, J. C., J. Ross, Jr., N. O. Kaplan. Lack of detectable change in cyclic AMP during the cardiac inotropic response to isoproterenol immobilized on glass beads. Proc. Natl. Acad. Sci. US 72: 824–829, 1975.
 341. Vergara, J., F. Zambrano, J. D. Robertson, and H. Elrod. Isolation and characterization of luminal membranes from urinary bladder. J. Cell Biol. 61: C83–C94, 1974.
 342. Wachtel, H., and E. R. Kandel. Conversion of synaptic excitation to inhibition at a dual chemical synapse. J. Neurophysiol. 34: 56–68, 1971.
 343. Wald, F. Ionic differences between somatic and axonal action potentials in snail giant neurones. J. Physiol. London 220: 267–281, 1972.
 344. Warner, A. E. The electrical properties of the ectoderm in the amphibian embryo during induction and early development of the nervous system. J. Physiol. 235: 267–286, 1973.
 345. Warner, A. E., and P. A. Lawrence. Electrical coupling across developmental boundaries in insect epidermis. Nature 245: 47–50, 1973.
 346. Washizu, Y. Single spinal neurons excitable from two different antidromic pathways. Japan J. Physiol. 10: 121–131, 1960.
 347. Watanabe, A. The interaction of electrical activity among neurons of lobster cardiac ganglion. Japan J. Physiol. 8: 305–318, 1958.
 348. Watanabe, A., and H. Grundfest. Impulse propagation at the septal and commissural junctions of crayfish lateral giant axons. J. Gen. Physiol. 45: 267–308, 1961.
 349. Watanabe, A., S. Obara, and T. Akiyama. Pacemaker potentials for the periodic burst discharge in the heart ganglion of a stomatopod, Squilla oratorio. J. Gen. Physiol. 50: 839–862, 1967.
 350. Waxman, S. G. Micropinocytotic invaginations in the axolemma of peripheral nerves. Z. Zellforsch. Mikroskop. Anat. 86: 571–573, 1968.
 351. Waxman, S. G., and G. D. Pappas. Pinocytosis at postsynaptic membranes: electron microscopic evidence. Brain Res. 14: 240–244, 1969.
 352. Waxman, S. G., G. D. Pappas, and M. V. L. Bennett. Morphological correlates of functional differentiation of nodes of Ranvier along single fibers in the neurogenic electric organ of the knifefish, Sternarchus. J. Cell Biol. 53: 210–224, 1972.
 353. Waziri, R. Electriral transmission mediated by an identified cholinergic neuron of Aplysia. Life Sci., Part 1 8: 469–476, 1969.
 354. Weidmann, S. The diffusion of radiopotassium across intercalated disks of mammalian cardiac muscle. J. Physiol. London 187: 323–342, 1966.
 355. Weight, F. F., and A. Padjen. Slow synaptic inhibition: evidence for synaptic inactivation of sodium conductance in sympathetic ganglion cells. Brain Res. 55: 219–224, 1973.
 356. Weicht, F. F., and A. Padjen. Acetylcholine and slow synaptic inhibition in frog sympathetic ganglion. Brain Res. 55: 225–228, 1973.
 357. Weight, F. F., G. Petzold, and P. Greengard. Guanosine 3′,5′‐monophosphate in sympathetic ganglia: increase associated with synaptic transmission. Science 186: 942–944, 1974.
 358. Weight, F. F., and J. Votava. Slow synaptic excitation in sympathetic ganglion cells: evidence for synaptic inactivation of potassium conductance. Science 170: 755–758, 1970.
 359. Weingart, R. The permeability to tetraethylammonium ions of the surface membrane and the intercalated discs of sheep and calf myocardium. J. Physiol. London 240: 741–762, 1974.
 360. Werblin, F. S., and J. E. Dowling. Organization of the retina of the mud puppy, Necturus maculosus. II. Intracellular recording. J. Neurophysiol. 32: 339–355, 1969.
 361. Willows, A. O. D., D. A. Dorsett, and G. Hoyle. The neuronal basis of behavior in Tritonia. III. Neuronal mechanism of a fixed action pattern. J. Neurobiol. 4: 255–285, 1973.
 362. Wine, J. J., and F. B. Krasne. The organization of escape behavior in the cray fish. J. Exptl. Biol. 56: 1–18, 1972.
 363. Woodbury, J. W., and W. E. Crill. The potential in the gap between two abutting cardiac muscle cells. A closed solution. Biophys. J. 10: 1084–1089, 1970.
 364. Woody, C. D., and G. Brozek. Gross potential from facial nucleus of cat as an index of neural activity in response to glabellar tap. J. Neurophysiol. 32: 704–716, 1969.
 365. Wylie, R. M. Evidence of electrotonic transmission in the vesitublar nuclei of the rat. Brain Res. 50: 179–183, 1973.
 366. Yee, A. G. Gap junctions between hepatocytes in regenerating rat liver. J. Cell Biol. 55: 294a, 1972.
 367. Zampighi, G., and J. D. Robertson. Fine structure of the synaptic discs separated from the goldfish medulla oblongata. J. Cell Biol. 56: 92–105, 1973.
 368. Zieglgänzberger, W., and C. Reiter. Interneuronal movement of procion yellow in cat spinal neurones. Exptl. Brain Res. 20: 527–530, 1974.
 369. Zucker, R. S. Changes in the statistics of transmitter release during facilitation. J. Physiol. London 229: 787–810, 1973.

Contact Editor

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

* Required Field

How to Cite

M. V. L. Bennett. Electrical Transmission: A Functional Analysis and Comparison to Chemical Transmission. Compr Physiol 2011, Supplement 1: Handbook of Physiology, The Nervous System, Cellular Biology of Neurons: 357-416. First published in print 1977. doi: 10.1002/cphy.cp010111