Membrane potential oscillations can be induced in molluscan neurones under a variety of artificial conditions. In the so-called ‘burster’ neurones oscillations are generated even in isolated cells. A likely mechanism for ‘bursting’ involves the following ionic currents:

  1. A transient inward current carried by Na+ and Ca2+. This current is responsible for the upstroke of the action potentials.

  2. A delayed outward current carried by K+. This current is voltage-sensitive and is responsible for the downstroke of the action potential during the early part of the burst. It becomes progressively inactivated during the burst. Its amplitude depends on the intracellular pH.

  3. A rapidly developing outward current carried by K+ which is in-activated at potentials close to action potential threshold. This current tends to hold the membrane in the hyperpolarized state and is involved in spacing the action potentials.

  4. A prolonged inward current which may not inactivate. It is probably carried by both Na+ and Ca2+. This current is responsible for the depolarizing phase of the burst but also contributes to the action potential.

  5. A slowly developing outward current, carried by K+. This current appears as a result of a slow increase in intracellular ionized calcium and is responsible for the hyperpolarizing phase of the burst. Note that a transient increase in this current may also contribute to the falling phase of the action potential during the later stages of the burst. It is also sensitive to intracellular pH.

One of the more significant features of this system of producing membrane potential oscillations is that the frequency of the bursts depends on the rate at which the intracellular ionized calcium returns to its resting level. This process depends on the metabolic state of the animal which can thereby exert a considerable influence on the electrical activity of burster neurones.

The timing element or oscillator responsible for a centrally generated pattern of behaviour could be a network of cells or a single rate determining neurone. This review presents evidence for an ionic mechanism by which bursts of action potentials are generated in an exceptionally regular fashion in certain molluscan ‘burster’ neurones. The function of this particular oscillation is unknown but it may have some value as a model system.

Cell R15 in the abdominal ganglion of Aplysia is readily identifiable by its size, colour and position (Frazier et al. 1967). In isolated ganglia, it has a distinctive electrical rhythm which consists of trains, or bursts, of action potentials separated by periods of hyperpolarization (Chalazonitis, 1961; Strumwasser, 1965). The inter-burst hyperpolarization gives way to a slowly developing depolarization which culminates in the next parabolic burst of action potentials. During the burst, the action potentials at first increase in frequency. They reach a maximum rate and then the rate of firing declines. There is a general increase in the duration and size of the action potentials throughout the burst and there are changes in the action potential undershoot. Cells apparently homologous to R15 have been described in Helix aspersa (cell A; Kerkut & Meech, 1966) and Otala lactea (cell 11; Gainer, 1972a). Fig. 1 shows intracellular recordings from each of these cells showing typical patterns of activity. The oscillation appears to be an endogenous rhythm which persists even when the cell is totally isolated (Alving, 1968; Gainer, 1972b; Chen, von Baumgarten & Takeda, 1971).

Fig. 1.

Patterns of electrical activity in ‘burster’ neurones. (A) Helix aspersa, cell A; intracellular record from an active snail. (B) Aplysia californien, R16; superimposed action potentials from a single burst, intracellular record. (From Strumwasser, 1967.) (C) Otala lactea, cell 11; pen-recording of electrical activity. The action potentials are slightly attenuated. Constant current hyperpolarizing pulses between bursts 2–5 indicate change in membrane resistance. Note the low membrane resistance at the end of each burst (from Gainer, 1972c).

Fig. 1.

Patterns of electrical activity in ‘burster’ neurones. (A) Helix aspersa, cell A; intracellular record from an active snail. (B) Aplysia californien, R16; superimposed action potentials from a single burst, intracellular record. (From Strumwasser, 1967.) (C) Otala lactea, cell 11; pen-recording of electrical activity. The action potentials are slightly attenuated. Constant current hyperpolarizing pulses between bursts 2–5 indicate change in membrane resistance. Note the low membrane resistance at the end of each burst (from Gainer, 1972c).

Whether or not the oscillation has any physiological significance is not known. Typical ‘bursting’ is recorded in R15 only in isolated ganglia (Stinnakre & Tauc, 1969) and the cell fires irregularly when the connexion to the branchial ganglion remains intact. (But see Kater & Kaneko, 1972.)

The cells themselves appear to have a secretory function and each has an axon which terminates in the heart. It has been suggested that they are ideally situated for the release of an agent having a general effect on the animal. In Aplysia activation of osmoreceptors in the osphradium by dilute sea water inhibits R15 (Stinnakre & Tauc, 1969) and the cell is known to contain one or more polypeptide hormones apparently involved in ionic regulation or water balance. Injection of an R16 homogenate into the haemocele of Aplysia produces a 3–10% weight increase in 90 min even in 5% hyperosmotic sea water. Control injections of other neurones have no effect and the action of the R15 homogenate is destroyed by proteolytic digestion (Kupfermann & Weiss, 1976).

The rate of firing of spontaneous action potentials in R15 reaches a maximum at about ‘dawn’ each day. The time of this peak depends on the time of ‘dawn’ to which the animal has been conditioned (Strumwasser, 1965). In cell 11 of Otala the pattern of electrical activity changes from bursting to continuous firing for 2–3 h near 12 a.m. (Gainer, 1972b) and in Helix a similar change in pattern occurs in cell A at about 12 p.m. (Meech, 1969). However, in neither of these latter cases has it been possible to change the time of the transition period by altering the environment of the snails.

In Helix, cell A is spontaneously active for most of the year. However, during the summer this spontaneous activity is often absent and the membrane potential reaches a maximum level of about −63 mV (Kerkut & Meech, 1967). At this time the animals themselves are quiescent and are said to be in diapause. Animals in this state can be recognized by the calcarious epiphragm covering the shell aperture. Cell 11 from specimens of Otala in diapause have a lower membrane resistance and a higher potassium permeability than actively ‘bursting’ cells. The current-voltage relationship of the cell membrane is linear rather than anomalously rectifying (Gainer, 1972b). Synthesis of a polypeptide specific to cell 11 is also inhibited.

During membrane potential oscillations of the type recorded from ‘burster’ neurones (Fig. 1), there are repeated changes in the current flow across the neuronal membrane. A slowly developing inward current depolarizes the cell to its action potential threshold and produces a train of action potentials while a slowly developing outward current tends to terminate the burst. This simplified description implies that if the membrane voltage is held at a fixed potential, it should be possible to identify a membrane current rising and falling with the same frequency as the bursts of action potentials seen in the unclamped cell.

In reality, if the membrane is held steady at some slightly depolarized potentials long-lasting inward current develops as shown in Fig. 2 A (Gola, 1974b; Eckert & Lux, 1975; see also Wilson & Wachtel, 1974). At slightly higher potentials, an outward current eventually develops which opposes the inward current (Gola, 1974b). This current, which is responsible for the post-burst hyperpolarization, slowly declines when the membrane is repolarized to its resting level.

Fig. 2.

Voltage-clamp records from cell RJ5, Aplysia fasciata. (A) Slowly developing inward current (lower trace) generated in response to a step depolarization lasting 3 s; holding potential, −47 mV. (B) Membrane currents generated in response to a ramp change in membrane voltage from − 82 to −9 mV and back to −82 mV over a period of 62 s. (C) Lower trace: typical electrical activity recorded intracellularly from R15. Upper trace:-membrane currents generated by a ramp change in membrane voltage. The currents during the depolarizing ramp have been superimposed on those recorded during the repolarizing ramp. See text for explanation. (From Gola, 1974b.)

Fig. 2.

Voltage-clamp records from cell RJ5, Aplysia fasciata. (A) Slowly developing inward current (lower trace) generated in response to a step depolarization lasting 3 s; holding potential, −47 mV. (B) Membrane currents generated in response to a ramp change in membrane voltage from − 82 to −9 mV and back to −82 mV over a period of 62 s. (C) Lower trace: typical electrical activity recorded intracellularly from R15. Upper trace:-membrane currents generated by a ramp change in membrane voltage. The currents during the depolarizing ramp have been superimposed on those recorded during the repolarizing ramp. See text for explanation. (From Gola, 1974b.)

Since the changes in membrane current are bound up with changes in membrane potential a useful approach is to slowly depolarize and repolarize the cell in a ramp fashion (Gola, 1974b). Fig. 2B shows that the inward current develops as the membrane potential of the cell is slowly changed from − 82 to − 9 mV, but that it does not reappear as the cell is repolarized. Thus, if the membrane voltage is slowly changed back and forth over a sufficient range of potential, the net membrane current is at first inward and then outward.

A model presented by Gola (1974b) which summarizes these findings and which accounts for the potential oscillation is shown in Fig. 2C. The lower trace is the typical bursting pattern of action potentials. During the burst position 1 is the potential at which the net membrane current is zero (see upper trace). The membrane is stable at position 1 for a short time but as the characteristics of the membrane change the stable position moves to a more negative potential (position 2). The current voltage relationship then slowly returns to its former state and the membrane potential moves back from position 2 to position 1. Position 3 is an intermediate state.

Identification of currents

There is good evidence that the post-burst hyperpolarization is caused by an increase in the potassium conductance of the cell membrane. Fig. iC shows that the conductance of the cell membrane (as judged by the voltage displacement produced by constant current hyperpolarizing pulses) declines to a minimum value just before the start of the next burst (Gainer, 1972c). In R15Junge and Stephens 1973) were able to demonstrate that the value of the reversal potential of the post-burst hyperpolarization depends on the concentration of potassium in the bathing medium. This increase in potassium conductance appears to be a direct result of an increase in the level of ionized calcium in the cell cytoplasm. There is a considerable increase in the level of ionized calcium during a train of impulses in R15 (Stinnakre & Tauc, 1973) and in bursting neurones, the calcium level increases during the burst and declines during the hyperpolarizing phase (see Fig. 4; Thomas & Gorman, 1977; Gorman & Thomas, 1978). If cells are injected with the calcium-chelating agent EGTA so that the increase in intracellular ionized calcium is prevented, they lose their post-burst hyperpolarization (see Fig. 5; Meech, 1974a). Injection of calcium, on the other hand, leads to a hyperpolarization of the cell membrane (see Fig. 3; Meech & Strumwasser, 1970; Meech, 1972; Meech, 1978). The abrupt initiation of the hyperpolarizing phase of the burst (see Fig. 1) is accounted for at least in part by the steep relationship between potassium conductance and intracellular calcium (Meech & Thomas, unpublished).

Fig. 3.

Pen recording of an experiment to show the effect of calcium injection on the membrane potential of cell A (upper trace) and on intracellular pH (lower trace). The duration of each calcium injection is shown below the membrane potential trace (from Meech & Thomas, 1977).

Fig. 3.

Pen recording of an experiment to show the effect of calcium injection on the membrane potential of cell A (upper trace) and on intracellular pH (lower trace). The duration of each calcium injection is shown below the membrane potential trace (from Meech & Thomas, 1977).

Fig. 4.

Changes in free intracellular calcium and membrane potential in cell R15 (Aplysia califomica) during bursts of activity. Changes in Arsenazo III dye absorbance (upper trace) indicate changes in the level of ionized calcium. (A) Effect of a 20 nA depolarizing current passed during a single burst. The membrane hyperpolarization is larger and the next burst delayed until the absorbance has declined almost to its normal baseline value. (B) Effect of period under voltage clamp. The dye absorbance declines below its normal base-line value at the time when the next burst is expected. The next burst occurs immediately upon release from voltage-clamp (from Thomas & Gorman, 1977).

Fig. 4.

Changes in free intracellular calcium and membrane potential in cell R15 (Aplysia califomica) during bursts of activity. Changes in Arsenazo III dye absorbance (upper trace) indicate changes in the level of ionized calcium. (A) Effect of a 20 nA depolarizing current passed during a single burst. The membrane hyperpolarization is larger and the next burst delayed until the absorbance has declined almost to its normal baseline value. (B) Effect of period under voltage clamp. The dye absorbance declines below its normal base-line value at the time when the next burst is expected. The next burst occurs immediately upon release from voltage-clamp (from Thomas & Gorman, 1977).

Fig. 5.

Post-burst hyperpolarization in cell R15 (Aplysia califomica). Pen recording of an experiment to show the effect of injecting calcium/EGTA. Constant current hyperpolarizing pulses provide an estimate of membrane resistance. The burst was generated by a 5 s depolarizing current (2 × 10−3 A) injected at the point indicated by the arrow, (a) Control; (b) after injection of a small volume of calcium/EGTA buffer containing 3 × 10−6 M free calcium ion at pH 7·26; (c) after 1 h to allow for sequestration of the injected calcium, the remaining EGTA buffers the calcium influx during the burst; (d) the post-burst hyperpolarization after injection of a further 5 × 10−101 (approx.) calcium/EGTA solution. See Meech (1974c) for composition of solutions. (From Meech, 1974a.)

Fig. 5.

Post-burst hyperpolarization in cell R15 (Aplysia califomica). Pen recording of an experiment to show the effect of injecting calcium/EGTA. Constant current hyperpolarizing pulses provide an estimate of membrane resistance. The burst was generated by a 5 s depolarizing current (2 × 10−3 A) injected at the point indicated by the arrow, (a) Control; (b) after injection of a small volume of calcium/EGTA buffer containing 3 × 10−6 M free calcium ion at pH 7·26; (c) after 1 h to allow for sequestration of the injected calcium, the remaining EGTA buffers the calcium influx during the burst; (d) the post-burst hyperpolarization after injection of a further 5 × 10−101 (approx.) calcium/EGTA solution. See Meech (1974c) for composition of solutions. (From Meech, 1974a.)

In summary, the available evidence suggests the following mechanism for the oscillations in membrane potential seen in burster neurones.

  1. A slowly developing inward current depolarizes the cell membrane. It is probably at least partly carried by calcium ions and contributes to the slow increase in intracellular calcium which occurs at the same time.

  2. The increase in intracellular calcium causes an increase in the potassium conductance of the cell membrane which hyperpolarizes the cell and terminates the burst.

  3. In the hyperpolarized state there is no longer an inward current and the level of ionized calcium returns to its resting level. This process is dependent on metabolism.

This description accounts for the oscillations in a general fashion, but the burst of action potentials themselves have complex characteristics and so a more details examination of the properties of the cell membrane is required. Five different current components appear to be involved.

The importance of calcium ions in the generation of action potentials in molluscan neurones was first demonstrated by Oomura, Ozaki & Maéno (1961). Subsequently action potentials in many neurones of both Helix and Aplysia were shown to depend on both sodium and calcium ions (Kerkut & Gardner, 1967; Junge, 1967). Voltage-clamp experiments show that depolarization produces a transitory inward current which can be carried by calcium ions in sodium-free saline and sodium ions in calcium-free saline (Geduldig & Gruener, 1970; Krishtal & Magura, 1970; Standen, 1974).

By analogy with the early and late inward currents in squid axons (Baker, Hodgkin & Ridgway, 1971) it seemed likely that the transient inward current referred to above and the prolonged inward current described by Gola (1974b) and Eckert & Lux (1975) – see Fig. 2 – passed through two different populations of channels. In perfused Helix neurones, the calcium current decay is made up of at least two exponential processes, but it is apparently not possible to separate the currents into early and late components (Kostyuk & Krishtal, 1977). Akaike, Lee & Brown (1977) have suggested that the inward current inactivates only slowly during a small depolarization and that it is not necessary to postulate two distinct conductances for the transient and persistent channels (but see Connor, 1979).

The prolonged inward current studied by Eckert & Lux (1976) in cell A remains essentially unaffected until 95% or more of the NaCl in the medium is replaced by glucose. On the other hand, substitution of 30 mM-CoCl2 for 10 mM-CaCl2 abolishes the inward current in 2–5 min. These observations suggest that the prolonged inward current is primari’y carried by calcium. However, the calcium blocking agent D-600 does not abolish the inward current in normal saline (Barker & Smith, 1978), although it is effective when NaCl and CaCl2 are both replaced by BaCl2. These conflicting observations may arise from species differences, but it is also possible that the prolonged inward current like the transient inward current, is a mixed current, both sodium and calcium ions being involved.

As shown Fig. 1, successive action potentials during a burst change in shape. One possible reason for this is that the characteristics of the inward current change in some way during the train. For example, it may be activated more rapidly with succeeding action potentials. A facilitating calcium current has been reported on the basis of aequroin experiments in Helix (Lux & Heyer, 1977) and Anisodoris (Eckert, Tillotson & Ridgway, 1977) and on the basis of an apparent discrepancy between outward currents and potassium flux in cell A (Heyer & Lux, 1976b), but this has not been confirmed with experiments with arsenazo III (Gorman & Thomas, 1978; Ahmed & Connor, 1979) or by experiments on perfused snail neurones (Akaike et al. 1978).

Another possible reason for the increase in the duration of the action potentials during a train is that the outward potassium current responsible for the repolarizing phase may become progressively inactivated. As shown in Fig. 6, inactivation of the outward current either during a depolarizing pulse or with successive depolarizing pulses is a well-established phenomenon in molluscan neurones (Gola & Romey, 1971; Leicht, Neves & Wellhöner, 1971; Neher & Lux, 1971; Connor & Stevens, 1971a) and the gradual inactivation of the delayed outward current during a train of impulses seems likely (see also Aldrich, Getting & Thompson, 1979).

Fig. 6.

Inactivation of outward current in cell R15, Aplysia fasciata. (A) Arrowed trace: traitaient inward and delayed outward current recorded under voltage-clamp in response to a 50 mV step depolarization. Superimposed records show the currents recorded in response to subsequent 50 mV pulses at a frequency of 1·5 c/s. Note the depression in outward current. Holding potential, −54 mV. (B) Depression in outward currents with repetitive depolarizing pulses shown on a longer time scale. Holding potential, −48 mV. (C) Outward current recorded in response to a prolonged 55 mV depolarizing pulse. Note that the outward current reaches a maximum within 0·1 s and then declines to a new steady level (A, B from Gola, 1974a; C from Gola & Romey, 1971).

Fig. 6.

Inactivation of outward current in cell R15, Aplysia fasciata. (A) Arrowed trace: traitaient inward and delayed outward current recorded under voltage-clamp in response to a 50 mV step depolarization. Superimposed records show the currents recorded in response to subsequent 50 mV pulses at a frequency of 1·5 c/s. Note the depression in outward current. Holding potential, −54 mV. (B) Depression in outward currents with repetitive depolarizing pulses shown on a longer time scale. Holding potential, −48 mV. (C) Outward current recorded in response to a prolonged 55 mV depolarizing pulse. Note that the outward current reaches a maximum within 0·1 s and then declines to a new steady level (A, B from Gola, 1974a; C from Gola & Romey, 1971).

In molluscan neurones, a depolarizing pulse generates two different delayed outward currents. Both are carried by potassium ions but one depends on an increase in intracellular calcium during the pulse, while the other appears to be simply dependent on the voltage across the cell membrane (see Meech, 1978, for review). The two components are demonstrated by the experiment on cell A shown in Fig. 7 A. The graph shows the outward current measured at the end of a 55 ms pulse plotted against the potential of the membrane during the pulse. Under normal conditions the curve is ‘N-shaped’ (Meech & Standen, 1975) and in Fig. 7A, this is shown by the filled circles. The open circles show that the effect of injecting EGTA into the cell is to reduce the potassium currents for depolarizing pulses up to about +150 mV. The remaining current, which is primarily the voltage-sensitive component, can be subtracted from the control curve leaving the calcium-dependent component (see Fig. 7B). A similar effect is produced by washing the cell in calcium-free saline (Meech & Standen, 1975). The current-voltage relationship for the calcium-dependent component closely resembles the relationship between membrane potential and light output measured in aequorin-injected cells (Lux & Heyer, 1977). The amplitude of the calcium-dependent component is thus directly related to the level of intracellular calcium.

Fig. 7.

(A) The relationship between outward membrane current and membrane potential in cell A, before (filled circles) and after (open circles) injection with EGTA. Abscissa: command potential; ordinate: membrane current measured 55 ms after the beginning of the command pulse. For EGTA solution and voltage-clamp circuit see Meech (1974c). Cell bathed in normal Tris buffered saline, see Meech & Standen (1975). Holding potential − 5 mV; cell diameter 220 μm; temperature 23 °C (replotted from Meech, 1978). (B) The calcium-dependent component of outward current, the difference between the two curves shown in (A).

Fig. 7.

(A) The relationship between outward membrane current and membrane potential in cell A, before (filled circles) and after (open circles) injection with EGTA. Abscissa: command potential; ordinate: membrane current measured 55 ms after the beginning of the command pulse. For EGTA solution and voltage-clamp circuit see Meech (1974c). Cell bathed in normal Tris buffered saline, see Meech & Standen (1975). Holding potential − 5 mV; cell diameter 220 μm; temperature 23 °C (replotted from Meech, 1978). (B) The calcium-dependent component of outward current, the difference between the two curves shown in (A).

Except for small depolarizations (see Fig. 2) the inward current responsible for the increase in intracellular calcium is concealed by the much larger outward current. However, in EGTA-injected cells the outward current inactivates sufficiently during a prolonged pulse for the maintained inward current to be seen (see Fig. 8). Such EGTA-injected cells produce prolonged action potentials which have a plateau on their falling phase (Meech, 1974b).

Fig. 8.

Currents recorded from an EGTA-injected cell A under voltage clamp. Three super-imposed records are shown; membrane voltage at the top and membrane current at the bottom. Pulse duration 1 s. As in Fig. 6, the outward current reaches a maximum in less than 0·1 s and then declines. For 50 and 60 mV command pulses the final current is inward. This long-lasting inward current has a reversal potential at about +15 mV. Cell bathed in normal Tris buffered saline (Meech & Standen, 1975). Holding potential − 50 mV; cell diameter 200 μm; temperature 21–23 °C.

Fig. 8.

Currents recorded from an EGTA-injected cell A under voltage clamp. Three super-imposed records are shown; membrane voltage at the top and membrane current at the bottom. Pulse duration 1 s. As in Fig. 6, the outward current reaches a maximum in less than 0·1 s and then declines. For 50 and 60 mV command pulses the final current is inward. This long-lasting inward current has a reversal potential at about +15 mV. Cell bathed in normal Tris buffered saline (Meech & Standen, 1975). Holding potential − 50 mV; cell diameter 200 μm; temperature 21–23 °C.

The calcium-dependent potassium current inactivates slowly if at all (Meech & Standen, unpublished; Aldrich et al. 1979). However, the inactivation undergone by the voltage-dependent component (see Fig. 8) is sufficient to account for the broadening of the action potentials during a train.

An alternative hypothesis has been presented by Eckert & Lux (1977), who have identified a voltage-dependent inactivation of the outward current which has the appearance of being calcium-dependent. They find that a 1 s prepulse has the effect of depressing currents produced by a second pulse presented 1 s later. The effect of the prepulse is to abolish the ‘N-shape’ of the IV curve in a similar fashion to the effect of injecting EGTA shown in Fig. 7. Eckert and Lux conclude that it is primarily the calcium-dependent currents which are depressed and that an influx of calcium is responsible for this depression. The hypothesis is supported by the observation that injection of calcium into cell A can reduce the size of the time-dependent outward currents recorded under voltage-clamp (Heyer & Lux, 1976b).

In this scheme the calcium dependency of the outward current arises from the influx of calcium across the membrane rather than from the increase in intracellular calcium. Consequently, Eckert and Lux propose that the influx of calcium during a train of impulses leads to a progressive reduction in the amplitude of the outward current and to a prolongation of the action potential. This proposal seems unlikely for the following reasons.

  1. In cell A, inactivation of outward current occurs in EGTA-injected cells (see Fig. 8) and in calcium-free saline (Meech & Standen, unpublished).

  2. Injection of a small volume of calcium chloride into cell A has the effect of reducing the net inward current measured during a depolarizing pulse. It has no effect on the net outward current (see Fig. 9) and so this should shorten rather than prolong the action potentials in a train.

Fig. 9.

Effect of calcium injection on membrane currents in cell A. (A) Pen-recording of the complete experiment. Except for the periods when voltage-clamp was imposed, hyperpolarizing constant-current pulses were injected into the cell to provide an estimate of membrane resistance. At the point indicated by the filled circle the cell was pressure-injected with calcium chloride (see Meech, 1974c for conditions). About 5 s later the cell was voltage-clamped and then briefly depolarized 100 mV by a 37·5 ms command pulse. Cell bathed in Tris buffered saline (Meech & Standen, 1975). Holding potential − 50 mV; cell diameter 200 µm temperature 21–23 °C. (B) Pen-recording of the membrane currents generated by the sequence of four 50 mV depolarizing command pulses shown in A. The trace was made by replaying the stored current records at a slow speed. (C) Superimposed oscilloscope records of the traces shown in B. Note that the net inward current during the depolarizing pulse is smaller after the calcium injection while the net outward current is approximately the same.

Fig. 9.

Effect of calcium injection on membrane currents in cell A. (A) Pen-recording of the complete experiment. Except for the periods when voltage-clamp was imposed, hyperpolarizing constant-current pulses were injected into the cell to provide an estimate of membrane resistance. At the point indicated by the filled circle the cell was pressure-injected with calcium chloride (see Meech, 1974c for conditions). About 5 s later the cell was voltage-clamped and then briefly depolarized 100 mV by a 37·5 ms command pulse. Cell bathed in Tris buffered saline (Meech & Standen, 1975). Holding potential − 50 mV; cell diameter 200 µm temperature 21–23 °C. (B) Pen-recording of the membrane currents generated by the sequence of four 50 mV depolarizing command pulses shown in A. The trace was made by replaying the stored current records at a slow speed. (C) Superimposed oscilloscope records of the traces shown in B. Note that the net inward current during the depolarizing pulse is smaller after the calcium injection while the net outward current is approximately the same.

Injection of calcium into cell A increases the potassium conductance of the cell membrane and hyperpolarizes the cell (see Fig. 3). Under voltage-clamp the injection has the effect of increasing the instantaneous or time-independent current measured during a step change in voltage. This is not obvious in Fig. 9 because of the small amount of calcium injected. The effect of a higher calcium injection is shown in Figure. 10. These voltage-clamp records show (on the right) the currents generated by 50 mV depolarizing pulses and (on the left) 50 mV hyperpolarizing pulses. After injection (middle records) the inward current generated by a hyperpolarizing pulse is larger than in the control (top records) because of the increase in the resting potassium conductance of the membrane. This increase in the resting conductance also accounts for the large immediate step change in the outward current generated by the depolarizing pulse. It appears to behave in a simple ohmic fashion with an equilibrium potential at about − 75 mV. Subtraction of this time-independent current shows that calcium injection produces a significant reduction in the time-dependent delayed outward current while leaving the transient inward current unaffected.

Fig. 10.

Voltage-clamp experiment on cell A to show the effect of calcium injection on membrane currents. Depolarizing command pulses are shown on the right and hyperpolarizing pulses are shown on the left. The voltage traces are shown above the recorded currents. The membrane currents recorded before injection are shown at the top. After calcium injection (middle records) there is a large increase in the instantaneous current which has largely disappeared 6 min later (lower record). Cell bathed in saline containing 23 HIM HCO3 equilibrated with 2·5% CO2. Holding potential − 50 mV; cell diameter 200 μm; temperature 21–23 °C.

Fig. 10.

Voltage-clamp experiment on cell A to show the effect of calcium injection on membrane currents. Depolarizing command pulses are shown on the right and hyperpolarizing pulses are shown on the left. The voltage traces are shown above the recorded currents. The membrane currents recorded before injection are shown at the top. After calcium injection (middle records) there is a large increase in the instantaneous current which has largely disappeared 6 min later (lower record). Cell bathed in saline containing 23 HIM HCO3 equilibrated with 2·5% CO2. Holding potential − 50 mV; cell diameter 200 μm; temperature 21–23 °C.

Another way in which it is possible to investigate the effect of injected calcium is to examine ‘tail’ currents. These currents represent the closing of potassium channels following a depolarizing pulse. In Fig. 11 the membrane was first depolarized by 100 mV to +50 mV for 45 ms and then repolarized to − 20 mV. The tail of outward current shown is made up of two exponentially declining components which can be separated as in Fig. 12. The fast component is known to be unaffected by calcium-free saline or EGTA injection, while the slower component is abolished by EGTA injection and greatly reduced in calcium-free saline. It presumably represents the closing of the calcium-dependent potassium channels. Analysis of ‘tail’ currents after CaCl2 injection shows that both fast and slow components are affected.

Fig. 11.

‘Tail’ current recorded from cell A. Upper trace, membrane voltage; lower trace, membrane current. Cell depolarized to + 50 mV for 45 ms and repolarizcd to − 20 mV. The currents generated during the first 100 mV step are not shown but there is a slowly declining outward current ‘tail’ during the second step. Cell bathed in normal Tris buffered saline (Meech & Standen, 1975). Holding potential, − 50 mV; cell diameter 200 μm, temperature 21–23 °C.

Fig. 11.

‘Tail’ current recorded from cell A. Upper trace, membrane voltage; lower trace, membrane current. Cell depolarized to + 50 mV for 45 ms and repolarizcd to − 20 mV. The currents generated during the first 100 mV step are not shown but there is a slowly declining outward current ‘tail’ during the second step. Cell bathed in normal Tris buffered saline (Meech & Standen, 1975). Holding potential, − 50 mV; cell diameter 200 μm, temperature 21–23 °C.

Fig. 12.

Effect of HC1 injection on ‘tail’ currents in cell A. Tail generated as shown in Fig. 11. Abscissa, time after repolarization; ordinate, membrane current (logarithmic scale). •, Total current; ○ fast component obtained by subtraction of extrapolated slow component from total current. (A) Control; (B) after brief (approx. 1 s) injection of solution containing 100 mM-HCl and 450 mM-KCl. Injection pipette diameter 1·5 μ m; (C) 23 min after injection. Cell bathed in normal Tris saline (Meech & Standen, 1975). Holding potential − 50 mV; cell diameter 200 μ tm; temperature 21–23 °C.

Fig. 12.

Effect of HC1 injection on ‘tail’ currents in cell A. Tail generated as shown in Fig. 11. Abscissa, time after repolarization; ordinate, membrane current (logarithmic scale). •, Total current; ○ fast component obtained by subtraction of extrapolated slow component from total current. (A) Control; (B) after brief (approx. 1 s) injection of solution containing 100 mM-HCl and 450 mM-KCl. Injection pipette diameter 1·5 μ m; (C) 23 min after injection. Cell bathed in normal Tris saline (Meech & Standen, 1975). Holding potential − 50 mV; cell diameter 200 μ tm; temperature 21–23 °C.

The effect of calcium injection on the delayed outward current appears to be brought about by a fall in intracellular pH (pH1). Fig. 3 shows that calcium injection not only hyperpolarizes the cell membrane but also reduces the intracellular pH of the cell cytoplasm. In fact since each injected calcium is exchanged for a proton, the effect of injecting CaCl2 on pH, is equivalent to injecting HC1 (see Meech & Thomas, 1977). Note, however, that HC1 injection does not hyperpolarize the cell membrane, i.e. the increase in potassium conductance does not depend on the fall in pHPFig. 12 shows that injection of HC1 into cell A like CaCl2 injection reduces the amplitude of the ‘tail’ currents. A comparison of the effects of HC1 and CaCl2 injection on the ‘tail’ currents recorded from cell A is shown in Fig. 13. In part A, the recovery of the ‘tail’ over a period of 25 min is shown. In B the fast component has been separated and C shows the recovery of the slow component. The filled

Fig. 13.

Recovery of ‘tail’ currents after injection of CaCl2 (open circles) and HC1 (filled circles) into cell A. (a) Abscissa, amplitude of tail current 1 msec. After repolarization expressed as a proportion of control; ordinate, time after brief (approx, 1 s) injection of, (•) solution containing 100 mM-HCl and 450 mM-KCl, (○) solution containing 100 mM-CaCl2 and 450 mM-KCl. (b, c) Recovery of tail current components separated as shown in Fig. 12. Rapidly declining component (b) and slowly declining component (c). Two different cells were used for the two experiments. Both cells bathed in normal Tris buffered saline (Meech & Standen, 1975). Holding potential in each case, −50 mV; cells’ diameter approximately 200 μm; temperature approximately 22 °C.

Fig. 13.

Recovery of ‘tail’ currents after injection of CaCl2 (open circles) and HC1 (filled circles) into cell A. (a) Abscissa, amplitude of tail current 1 msec. After repolarization expressed as a proportion of control; ordinate, time after brief (approx, 1 s) injection of, (•) solution containing 100 mM-HCl and 450 mM-KCl, (○) solution containing 100 mM-CaCl2 and 450 mM-KCl. (b, c) Recovery of tail current components separated as shown in Fig. 12. Rapidly declining component (b) and slowly declining component (c). Two different cells were used for the two experiments. Both cells bathed in normal Tris buffered saline (Meech & Standen, 1975). Holding potential in each case, −50 mV; cells’ diameter approximately 200 μm; temperature approximately 22 °C.

circles are from an HC1 injection experiment and the open circles show the effect of a CaCl2 injection. In both cases the fast component is the most affected and recovers most quickly. Inhibition of the slow component appears to develop with time and recovers more slowly. Brief injections of a solution containing 500 mM-PIPES buffer and 500 mM-KCl at pH 6·8 also give a reversible reduction in ‘tail’ current while control injections of a solution of 500 mM-HEPES and 500 mM-KCl at pH 7·4 give an increase in tail current amplitude.

In summary, the effect of injecting calcium into cell A is to increase the resting potassium conductance of the cell membrane. However, a secondary fall in pH1 caused by calcium/hydrogen exchange in the cell cytoplasm reduces the time-dependent outward currents seen under voltage clamp. The voltage-sensitive component of the outward current is the most affected.

The action potential frequency during a train or burst is greatly influenced by an outward current component which is inactivated at potentials near the action potential threshold (Hagiwara, Kusano & Saito, 1961; Connor & Stevens, 1971b, c; Neher, 1971). When inactivation is removed by hyperpolarization, a depolarizing pulse generates a rapidly developing outward current which opposes the fast inward current responsible for the rising phase of the action potential. Thus, the effect of the fast outward current is to reduce the frequency of firing of the action potentials (Connor & Stevens, 1971c). Early in the burst the action potential undershoot becomes smaller, possibly because of inactivation of the voltage-dependent delayed outward current. This means that the fast outward current is partially inactivated and the frequency of firing of action potentials increases. Later in the burst the increasing level of intracellular calcium tends to oppose the inward current and to slow the rate of firing.

One of the main difficulties in accepting the ionic mechanisms for ‘bursting’ presented above is that oscillations can be recorded in calcium-free saline (Carpenter & Gunn, 1970; Mathew & Roberge, 1971; Strumwasser, 1973), although it is sometimes necessary to apply a constant hyperpolarizing current. Fig. 14 shows an experiment showing oscillations recorded from R15 in calcium-free saline containing TTX (tetrodotoxin). Strumwasser has suggested that the oscillations (which occur in the absence of action potentials) are the result of an electrogenic sodium pump. However, the hyperpolarizing phase of these cycles appear to have a normal reversal potential and they are not abolished by ouabain (Junge & Stephens, 1973). Bursting is abolished if all the divalent cations in the external saline are replaced by 50 mM cobalt chloride (Barker & Gainer, 1975a) and it is likely that the 10−5 M calcium normally present in nominally calcium-free saline is sufficient to support bursting.

Fig. 14.

Intracellular record of membrane potential oscillations in cell R15 (Aplysia californica) in calcium-free tetrodotoxin (TTX) containing saline. The change from light to dark conditions is shown under trace 7. (From Strumwasser, 1974.)

Fig. 14.

Intracellular record of membrane potential oscillations in cell R15 (Aplysia californica) in calcium-free tetrodotoxin (TTX) containing saline. The change from light to dark conditions is shown under trace 7. (From Strumwasser, 1974.)

The effect of inhibitors however, must be treated with caution. Cell R15 can produced bursts in the presence of the calcium blocking agent D600 (Barker & Gainer, 1975a) or in the presence of verapamil (Meech, unpublished), but bursting in the presence of verapamil is unlike that of normal cells because the membrane resistance is highest at the most negative potential.

Bursting may be induced in other molluscan neurones in a variety of ways (Ducreux, 1978). In barium-containing salines, bursts of action potentials are generated which have a prolonged plateau on the falling phase. These bursts are blocked by 10 mM cobalt. In this experimental situation, barium carries the inward current and because it does not itself activate the potassium conductance (Meech & Thomas, unpublished), repolarization is delayed. The drug pentamethylene tetrazol (PTZ) also induces bursting, but in this case both TTX and Na-free saline block the bursts, although cobalt does not.

Although calcium influx appears to be necessary for the post-burst increase in potassium conductance observed in these neurones, it is possible that at least part of the increase in intracellular ionized calcium during the burst results from ‘calcium induced calcium release’ from some intracellular site. An argument against this possibility is that calcium injected into cell A appears to be taken up almost exclusively by mitochondria. Isolated preparations of mitochondria are known to take up Ca2+ and release H+ in a ratio of 1:1. Experiments on cell A in which the effects of injecting measured quantities of CaCl2 or HCl were studied show that the Ca:H ratio is 1:1 under these conditions. It is unlikely, therefore, that calcium injection induces any further release of calcium (see Meech & Thomas, 1977, for references).

Ahmed
,
Z.
&
Connor
,
J. A.
(
1979
).
Measurement of calcium influx under voltage clamp in molluscan neurones using the metallocromic dye arsenazo III
.
J. Physiol., Lond
.
286
,
61
82
.
Akaike
,
N.
,
Lee
,
K. S.
&
Brown
,
A. M.
(
1978
).
The calcium current of Helix neuron
.
J. gen. Physiol
.
71
,
509
531
.
Aldrich
,
R. W.
Jr
.,
Getting
,
R. A.
&
Thompson
,
S. H.
(
1979
).
Inactivation of delayed outward current in molluscan neuron somata
.
J. Physiol., Lond. (in the press)
.
Alving
,
B. O.
(
1968
).
Spontaneous activity in isolated somata of Aplysia pacemaker neurons
.
J. gen. physiol
.
51
,
29
45
.
Baker
,
P. F.
,
Hodgkin
,
A. L.
&
Ridgway
,
E. B.
(
1971
).
Depolarization and calcium entry in squid giant axons
.
J. Physiol., Lond
.
218
,
709
755
.
Barker
,
J. L.
&
Gainer
,
H.
(
1975a
).
Studies on bursting pacemaker potential activity in molluscan neurons. I. Membrane properties and ionic contributions
.
Brain Res
.
84
,
461
477
.
Barker
,
J. L.
&
Gainer
,
H.
(
1975b
).
Studies on bursting pacemaker potential activity in molluscan neurones. II. Regulation by divalent cations
.
Brain Res
.
84
,
479
500
.
Barker
,
J. L.
&
Smith
,
T. C.
Jr
. (
1978
).
Electrophysiological studies of molluscan neurons generating bursting pacemaker potential activity
.
In Abnormal Neuronal Discharges
(ed.
N.
Chalazonitis
and
M.
Boisson
), pp.
359
387
.
New York
:
Raven Press
.
Carpenter
,
D.
&
Gunn
,
R.
(
1970
).
The dependence of pacemaker discharge of Aplysia neurons upon Na+ and Ca++
.
J. Cell. Physiol
.
75
,
121
128
.
Chalazonitis
,
N.
(
1961
).
Chemopotentials in giant nerve cells (Aplysia fasciata
).
In Nervous Inhibition
(ed.
E.
Florey
), pp.
179
193
.
Oxford
:
Pergamon
.
Chen
,
C. F.
,
Von Baumgarten
,
R.
&
Takeda
,
R.
(
1971
).
Pacemaker properties of completely isolated neurones in Aplysia californica
.
Nature, New Biol
.
233
,
27
29
.
Connor
,
J. A.
(
1979
).
Calcium current in molluscan neurones: measurement under conditions which maximize its visibility
,
J. Physiol., Lond
.
286
,
41
60
.
Connor
,
J. A.
&
Stevens
,
C. F.
(
1971a
).
Inward and delayed outward membrane currents in isolated neural somata under voltage clamp
.
J. Physiol., Lond
.
213
,
1
19
.
Connor
,
J. A.
&
Stevens
,
C. F.
(
1971b
).
Voltage clamp studies of a transient outward membrane current in gastropod neural somata
.
J. Physiol., Lond
.
213
,
21
30
.
Connor
,
J. A.
&
Stevens
,
C. F.
(
1971c
).
Prediction of repetitive firing behaviour from voltage clamp data on an isolated neurone soma
.
J. Physiol., Lond
.
213
,
31
53
.
Ducreux
,
C.
(
1978
).
Pharmacological alteration of the ionic currents underlying slow potential wave generation in molluscan neurons
.
In Abnormal Neuronal Discharges
(ed.
N.
Chalazonitis
and
M.
Boisson
), pp.
271
285
.
New York
:
Raven Press
.
Eckert
,
R.
&
Lux
,
H. D.
(
1975
).
A non-inactivating inward current recorded during small depolarizing voltage steps in snail pacemaker neurons
.
Brain Res
.
83
,
486
489
.
Eckert
,
R.
&
Lux
,
H. D.
(
1976
).
A voltage-sensitive persistent calcium conductance in neuronal somata of Helix
.
J. Physiol., Lond
.
254
,
129
151
.
Eckert
,
R.
&
Lux
,
H. D.
(
1977
).
Calcium-dependent depression of a late outward current in snail neurons
.
Science, Lond
.
197
,
472
475
.
Eckert
,
R.
,
Tillotson
,
D.
&
Ridgway
,
E. B.
(
1977
).
Voltage-dependent facilitation of Cas+ entry in voltage-clamped, aequroin-injected molluscan neurons
.
Proc. natn. Acad. Sci. U.S.A
.
74
,
1748
1752
.
Frazier
,
W. T.
,
Kandel
,
E. R.
,
Kupfermann
,
I.
,
Waziri
,
R.
&
Coggeshall
,
R. E.
(
1967
).
Morphological and functional properties of identified neurons in the abdominal ganglion of Aplysia californica
.
J. Neurophysiol
.
30
,
1288
1351
.
Gainer
,
H.
(
1972a
).
Patterns of protein synthesis in individual, identified molluscan neurons
.
Brain Res
.
39
,
369
385
.
Gainer
,
H.
(
1972b
).
Effects of experimentally induced diapause on the electrophysiology and protein synthesis patterns of identified molluscan neurons
.
Brain Res
.
39
,
387
402
.
Gainer
,
H.
(
1972c
).
Electrophysiological behavior of an endogenously active neurosecretory cell
.
Brain Res
.
39
,
403
418
.
Geduldig
,
D.
&
Gruener
,
R.
(
1970
).
Voltage clamp of the Aplysia giant neurone: early sodium and calcium currents
.
J. Physiol
.
211
,
217
244
.
Gola
,
M.
(
1974a
).
Evolution de la forme des potentiels d’action par stimulations répétitives
.
Pflügers Arch
.
346
,
121
140
.
Gola
,
M.
(
1974b
).
Neurones à ondes-salves des mollusques
.
Pflügers Arch
.
352
,
17
36
.
Gola
,
M.
&
Romey
,
G.
(
1971
).
Responses anomales a des courants sous-liminaires de certaines membranes somatiques (neurones géants d’Helix pomatia)
.
Pflügers Arch
.
32
,
7
,
105
131
.
Gorman
,
A. L. F.
&
Thomas
,
M. V.
(
1978
).
Changes in the intracellular concentration of free calcium ions in a pace-maker neurone, measured with the metallochromie indicator dye arsenazo III
.
J. Physiol., Lond
.
275
,
357
376
.
Hagiwara
,
S.
,
Kusano
,
K.
&
Saito
,
N.
(
1961
).
Membrane changes of Onchidium nerve cell in potassium-rich media
.
J. Physiol., Lond
.
155
,
470
489
.
Heyer
,
C. B.
&
Lux
,
H. D.
(
1976a
).
Properties of a facilitating calcium current in pace-maker neurones of the snail, Helix pomatia
.
J. Physiol., Lond
.
262
,
319
348
.
Heyer
,
C. B.
&
Lux
,
H. D.
(
1976b
).
Control of the delayed outward potassium currents in bursting pace-maker neurones of the snail, Helix pomatia
.
J. Physiol., Lond
.
262
,
349
382
.
Junge
,
D.
(
1967
).
Multi-ionic action potentials in molluscan giant neurones
.
Nature, Lond
.
215
,
546
548
.
Junge
,
D.
&
Stephens
,
C. L.
(
1973
).
Cyclic variation of potassium conductance in a burst-generating neurone in Aplysia
.
J. Physiol., Lond
.
235
,
155
181
.
Kater
,
S. B.
&
Kaneko
,
C. R. S.
(
1972
).
An endogenously bursting neuron in the gastropod mollusc, Heliosoma trivolvis
.
J. comp. Physiol
.
79
,
1
14
.
Kerkut
,
G. A.
&
Gardner
,
D. R.
(
1967
).
The role of calcium ions in the action potentials of Helix aspersa neurones
.
Comp. Biochem. Physiol
.
20
,
147
162
.
Kerkut
,
G. A.
&
Meech
,
R. W.
(
1966
).
The internal chloride concentration of H and D cells in the snail brain
.
Comp. Biochem. Physiol
.
19
,
819
832
.
Kerkut
,
G. A.
&
Meech
,
R. W.
(
1967
).
The effect of ions on the membrane potential of snail neurones
.
Comp. Biochem. Physiol
.
20
,
411
429
.
Kostyuk
,
P. G.
&
Krishtal
,
O. A.
(
1977
).
Separation of sodium and calcium currents in the somatic membrane of mollusc neurones
.
J. Physiol
.,
270
,
545
568
.
Krishtal
,
O. A.
&
Magura
,
I. S.
(
1970
).
Calcium ions as inward current carriers in mollusc neurones
.
Qpmp. Biochem. Physiol
.
35
,
857
866
.
Hermann
,
I.
&
Weiss
,
K. R.
(
1976
).
Water regulation by a presumptive hormone contained in IBentified neurosecretory cell R15 of Aplysia
.
J. gen. Physiol
.
67
,
113
123
.
Leicht
,
R.
,
Meves
,
H.
&
WellhÖner
,
H.-H.
(
1971
).
Slow changes of membrane permeability giant neurones of Helix pomatia
.
Pflügers Arch
.
323
,
63
79
.
Lux
,
H. D.
&
Heyer
,
C. B.
(
1977
).
An aequorin study of a facilitating calcium current in bursting pacemaker neurons of Helix
.
Neuroscience
2
,
585
592
.
Mathieu
,
R. A.
&
Roberge
,
F. A.
(
1971
).
Characteristics of pacemaker oscillations in Aplysia neurons
.
Can. J. Physiol. Pharmacol
.
49
,
787
795
.
Meech
,
R. W.
(
1969
).
Studies on pacemaker-like activity in molluscan neurones
.
In Biology Annual Report
, pp.
113
114
. California Institute of Technology, Pasadena.
Meech
,
R. W.
(
1972
).
Intracellular calcium injection causes increased potassium conductance in Aplysia nerve cells
.
Comp. Biochem. Physiol
.
42 A
.
493
499
.
Meech
,
R. W.
(
1974a
).
Calcium influx induces a post-tetanic hyperpolarization in Aplysia neurones
.
Comp. Biochem. Physiol
.
48A
,
387
395
.
Meech
,
R. W.
(
1974b
).
Prolonged action potentials in Aplysia neurones injected with EGTA
.
Comp. Biochem. Physiol
.
48A
:
397
402
.
Meech
,
R. W.
(
1974c
).
The sensitivity of Helix aspersa neurones to injected calcium ions
.
J. Physiol., Lond
.
237
,
259
277
.
Meech
,
R. W.
(
1978
).
Calcium-dependent potassium activation in nervous tissues
.
Ann. Rev. Biophys. Bioeng
.
7
,
1
18
.
Meech
,
R. W.
&
Standen
,
N. B.
(
1975
).
Potassium activation in Helix aspersa neurones under voltage clamp: a component mediated by calcium influx
.
J. Physiol., Lond
.
249
,
211
239
.
Meech
,
R. W.
&
Strumwasser
,
F.
(
1970
).
Intracellular calcium injection activates potassium conductance in Aplysia nerve cells
.
Fedn Proc
.
29
,
834
.
Meech
,
R. W.
&
Thomas
,
R. C.
(
1977
).
The effect of calcium injection on the intracellular sodium and pH of snail neurones
.
J. Physiol., Lond
.
265
,
867
879
.
Neher
,
E.
(
1971
).
Two fast transient current components during voltage clamp on snail neurons
.
J. gen. Physiol
.
58
,
36
53
.
Neher
,
E.
&
Lux
,
H. D.
(
1971
).
Properties of somatic membrane patches of snail neurons under voltage clamp
.
Pflügers Arch
.
322
,
35
38
.
Oomura
,
Y.
,
Ozaki
,
S.
&
Maéno
,
T.
(
1961
).
Electrical activity of a giant nerve cell under abnormal conditions
.
Nature, Lond
.
191
,
1265
1267
.
Standen
,
N. B.
(
1974
).
Properties of a calcium channel in snail neurones
.
Nature, Lond
.
250
,
340
342
.
Stinnakre
,
J.
&
Tauc
,
L.
(
1969
).
Central neuronal response to the activation of osmoreceptors in the osphradium of Aplysia
.
J. exp. Biol
.
51
,
347
361
.
Stinnakre
,
J.
&
Tauc
,
L.
(
1973
).
Calcium influx in active Aplysia neurones detected by injected aequorin
.
Nature, New Biol
.
242
,
113
115
.
Strumwasser
,
F.
(
1965
).
The demonstration and manipulation of a circadian rhythm in a single neuron
.
In Circadian Clocks
(ed.
J.
Aschoff
), pp.
442
462
.
Amsterdam
:
North-Holland
.
Strumwasser
,
F.
(
1967
).
Types of information stored in single neurons
.
In Invertebrate Nervous Systems
(ed.
C. A. G.
Wiersma
), pp.
291
319
.
University of Chicago Press
.
Strumwasser
,
F.
(
1973
).
Neural and humoral factors in the temporal organization of behavior
.
Physiologist
16
,
9
42
.
Strumwasser
,
F.
(
1974
).
Neuronal principles organizing periodic behaviors
.
In The Neurosciences, Third Study Program
(ed.
F. O.
Schmitt
and
F. G.
Worden
), pp.
459
478
.
Cambridge, Mass., and London
:
MIT Press
.
Thomas
,
M. V.
&
Gorman
,
A. L. F.
(
1977
).
Internal calcium changes in a bursting pacemaker neuron measured with arsenazo III
.
Science, N.Y
.
196
,
531
533
.
Wilson
,
W. A.
&
Wachtel
,
H.
(
1974
).
Negative resistance characteristic essential for the maintenance of slow oscillations in bursting neurons
.
Science
186
,
932
934
.