Comprehensive Physiology Wiley Online Library

Cell Biological Studies of Learning in Simple Vertebrate and Invertebrate Systems

Full Article on Wiley Online Library



Abstract

The sections in this article are:

1 Conceptual Issues in Study of Learning
2 Methodological Approaches to Study of Learning
3 Nonassociative Learning
3.1 Introduction
3.2 Simple Vertebrate System
3.3 Invertebrates
4 Associative Learning
4.1 Introduction
4.2 Simple Vertebrate Systems
4.3 Invertebrates
5 Conclusion
5.1 General Principles in Study of Learning
5.2 Coda
Figure 1. Figure 1.

Analysis of habituation and sensitization of hindlimb flexion response in acute spinal cat. A: experimental arrangement. Stimuli can be delivered to skin (SS1 or SS2), to cutaneous afferent nerve (SN1), or to group Ia muscle afferents (SN2). Responses can be recorded from dorsal root (RS), from α‐motoneurons (Rα) or interneurons, from ventral root (RN), or from muscle (RM). B: comparison of changes in excitatory postsynaptic potentials (EPSPs) recorded in deep peroneal motoneuron elicited by stimulation of either cutaneous afferent nerve (1, 2, 3) or muscle afferents (4, 5, 6). Stimulation of the 2 nerves was interpolated. 1, 2, 3, Polysynaptic EPSPs elicited by single‐shock stimuli delivered to posterior femoral cutaneous nerve. 1, Control period, stimulation once every 30 s. 2, Synaptic decrement, stimulation once per second. 3, Recovery, stimulation once every 30 s. 4, 5, 6, Monosynaptic EPSPs elicited by stimulation of deep peroneal nerve at 30‐s intervals during periods corresponding to those in 1, 2, and 3. Note constancy of monosynaptic responses during period of polysynaptic EPSP decrement.

A adapted from Thompson 386; B adapted from Spencer et al. 375.
Figure 2. Figure 2.

Possible role of spinal interneurons in habituation and sensitization of flexion reflex in cat. A: responses of different types of interneurons (nonplastic, type H, and type S) and muscle response during repetitive stimulation producing first sensitization and then habituation of flexion reflex. B: schematic diagram of possible neural substrates of habituation and sensitization. N, nonplastic synapses; H, habituating synapses; S, sensitizing synapses. According to this scheme, cutaneous stimuli can exert their influence on 2 separate systems: reflex system (S‐R pathway) that mediates habituation and “state” system that mediates sensitization.

From Groves and Thompson 181.
Figure 3. Figure 3.

Habituation and sensitization of gill‐ and siphon‐withdrawal responses in Aplysia. A: experimental arrangement for behavioral studies in intact animal showing gill and siphon in relaxed (A1) and contracted (A2) position. Parapodia and mantle shelf are shown retracted to reveal gill and siphon. Gill and siphon withdrawal can be elicited with tactile stimulus to siphon. Electric shock of tail (shock) or head serves as sensitizing stimulus. B: photocell recordings showing habituation and sensitization of gill‐withdrawal response. Tactile stimulus to siphon (bottom trace) elicited gill‐withdrawal response (top trace), which habituated with repeated stimulus deliveries. Application of noxious tail shock (arrow) caused marked sensitization of gill‐withdrawal response. C: neural circuit mediating gill‐ and siphon‐withdrawal responses. Siphon skin is innervated by group of ∼24 mechanoreceptor sensory cells (SN) whose soma are located in LE cluster in abdominal ganglion. Siphon sensory cells project both monosynaptically and through interneurons (INT) to gill and siphon motoneurons (MN), many of which are also located in abdominal ganglion. Effects of sensitizing stimuli in abdominal ganglion are mediated by group of facilitator interneurons (FAC INT) that terminate on sensory cells in synaptic and somatic regions. Habituation of reflex has been shown to involve a decrease in transmitter release from sensory neuron terminals [H(↓)]. Sensitization of reflex has been shown to involve an increase in transmitter release from sensory neurons [S(↑)]. Posttetanic potentiation has also been reported to occur at neuromuscular junction. [Adapted from Kandel and Schwartz 220.]

interstimulus interval (ISI), 1.5 min
Figure 4. Figure 4.

Electrophysiological analyses of habituation and sensitization of gill‐withdrawal response. A: homosynaptic depression and presynaptic facilitation at synapses from a siphon sensory cell onto L7, a gill motoneuron. Abdominal ganglion was dissected from animal and was maintained in vitro in artificial seawater (ASW). Intracellular stimulation of sensory neuron triggered action potential (bottom trace), which elicited monosynaptic EPSP in gill motoneuron (top trace). (Sharp negative and positive deflections that precede and follow action potential in sensory neuron are stimulus artifacts from depolarizing current pulse.) With repeated stimulation, action potential elicited successively smaller EPSPs in gill motoneuron. Electrical stimulation of left connective (arrow), which carries input to abdominal ganglion from head and tail, caused a marked increase in EPSP elicited in gill motoneuron. Synaptic depression and facilitation parallel and contribute to behavioral habituation and sensitization of gill‐withdrawal reflex. B: connective stimulation (B1) and application of 0.2 mM serotonin (B2) increase action‐potential duration in siphon sensory cells. Abdominal ganglion was maintained in artificial seawater containing 0.1 M tetraethylammonium (TEA) (K+ channel blocker) to facilitate measurement of action‐potential duration and maximize effects of facilitation. Similar though smaller changes also occur in normal seawater. Because Ca2+ entry is voltage dependent, broadening of action potential is believed to contribute to increase in Ca2+ influx and subsequent transmitter release during facilitation. C: comparison of changes in action‐potential duration in siphon sensory cell (bottom trace) and amplitude of EPSPs elicited in gill motoneuron L7 (top trace) in 1 preparation. With repeated stimulation, action‐potential duration and EPSP amplitude decrease concomitantly. Weak stimulation of pleuroabdominal connective (left vertical dashed line) causes a transient increase in action‐potential duration and EPSP amplitude. Stronger stimulation (right vertical dashed line) produces larger effects. Experiment conducted in 0.1 M TEA.

A adapted from Castellucci and Kandel 81; B from Klein and Kandel 233; C adapted from Kandel 218.
Figure 5. Figure 5.

Biophysical analyses of ionic changes underlying homosynaptic depression and presynaptic facilitation in siphon sensory cells. A: serotonin, a putative facilitatory transmitter, depresses an outward K+ current in siphon sensory cells. A1: voltage‐clamp experiment, performed in normal seawater. Voltage step from −35 to +15 mV (bottom trace) elicits initial inward current (downward deflection, top trace) carried largely by Na+ and Ca2+, followed by slower outward current (upward deflection, top trace) carried largely by K+. With repeated stimulation, inward current decreases, paralleling habituation. Application of serotonin (arrow) causes marked decrease in outward current. A2: voltage‐clamp experiment in presence of K+ channel blockers (TEA, tetraethylammonium; 4‐AP, 4‐aminopyridine). Same preparation as in A1. With K+ currents blocked, depolarizing step (bottom trace) elicits primarily an inward current. Application of serotonin (arrow) had no effect on current elicited by voltage step, indicating that serotonin normally exerts its action by depression of outward K+ current. B: effect of serotonin (5‐HT) on single‐channel currents recorded from siphon sensory cells in Aplysia. Left ordinate, number of open channels (n); right ordinate, current magnitude. Individual current steps are 2.6 pA. 1, Current in absence of serotonin. Current record was well fit by binomial distribution assuming that 5 channels are active in the patch and that each channel opens with probability of 0.84. 2, Current recorded 2 min after addition of 30 μM serotonin to bath. 3, Current recorded 1 min after addition of further dose of serotonin, raising total concentration to 60 μM. The 2 traces are a continuous recording. Downward current deflection at end of bottom trace is from intracellular current pulse used to monitor cell resistance. Serotonin produced decrease in average number of open channels but had no effect on unit current. 4, Current recorded ∼5 min after superfusion with serotonin‐free artificial seawater (ASW). Note increase in number of open channels.

A adapted from Klein and Kandel 234; B adapted from Siegelbaum et al. 364.
Figure 6. Figure 6.

Biochemical analyses of presynaptic facilitation in siphon sensory cells. A: injection of cAMP‐dependent protein kinase produces presynaptic facilitation. In base‐line control period, intracellular stimulation of sensory neuron (SN, bottom trace) evoked EPSP in a follower neuron (FN, top trace). Injection of catalytic subunit of cAMP‐dependent protein kinase led to increase in duration of sensory neuron action potential and increase in amplitude of EPSP evoked in follower cell. Experiment conducted in 100 mM tetraethylammonium (TEA). B: injection of kinase inhibitor blocks serotonin‐induced spike broadening. Recordings of action potentials in 2 different sensory neurons before (left) and after (right) application of serotonin. Serotonin produced marked spike broadening in control sensory cell but relatively little spike broadening in sensory neuron that had been injected with Walsh inhibitor, which blocks activation of cAMP‐dependent protein kinase. Experiment conducted in TEA. C: proposed sequence of steps underlying presynaptic facilitation in siphon sensory cells.

A from Castellucci et al. 82; B from Castellucci et al. 83.
Figure 7. Figure 7.

Analysis of habituation of crayfish escape response. A: neural circuit mediating escape response and proposed sites of plasticity. Tactile stimulation of posterior parts of crayfish produces forward somersault that is mediated by pair of neurons called lateral giants (LG). The LG have rectifying electrical synapse onto motor giants (MO G) and electrical synapse onto fast flexor motoneurons (FF MN). The LG receive direct electrical connections from sensory neurons (SN) as well as di‐ or trisynaptic input through heterogeneous group of interneurons (INT). These connections are believed to mediate the α‐ and β‐components, respectively, of complex EPSP evoked in LG by tactile stimuli. Habituation has been shown to involve a decrease in transmitter release at chemical synapses from primary sensory neurons onto interneurons. Synaptic depression can also occur at motor giant neuromuscular junction (↓) and posttetanic potentiation has been observed at neuromuscular junction of fast flexor motoneuron (↑); however, contribution of these forms of plasticity to behavioral learning is not clearly established. B: example of decrement in responses recorded from lateral giant fibers in response to repeated stimulation (1/min) of second root of abdominal nerve cord. Initial stimuli evoke action potential that fails as the β‐component of complex EPSP grows smaller with successive stimuli. Note that α‐component of EPSP remains relatively stable. C: example of decrement of monosynaptic EPSP in tactile interneuron. Stimulation of second root was adjusted to recruit single tactile afferent that elicited monosynaptic EPSP in tactile interneuron. With repeated stimulation, amplitude of EPSP decreases. C1, C2, C3: first, fourth, and seventh responses to stimulation at 0.5 Hz. [A and C adapted from Zucker 450,451; B adapted from Krasne 239.]

H(↓)
Figure 8. Figure 8.

Habituation and sensitization of crayfish defense response to tactile stimuli. A: proposed neural circuit and possible sites of plasticity. During habituation, neural responses elicited by tactile stimulus to back or tail decrease in interneurons (INT) and in excitor motoneuron (E) but do not change in inhibitor motoneuron (I) [H(0)]; these results suggest that habituation is due to changes in excitatory reflex pathway and not to buildup of peripheral inhibition. Sensitization has been proposed to involve increase in responses of interneurons and excitor motoneurons [S(↑)]; posttetanic potentiation (PTP) at neuromuscular junction of excitor motoneurons [S(↑)]; and decrease in responses of inhibitor motoneuron [S(↓)]. SN, sensory neuron; CN, command neuron. B: changes in number of excitor and inhibitor spikes accompanying sensitization. Note increase of excitor responses and decrease of inhibitor responses. C: PTP at excitor neuromuscular junction. Excitatory junction potentials recorded during test trains of excitor stimulation (30 Hz for 0.5 s) before (top trace) and 5 s after (bottom trace) conditioning train of excitor stimulation (10 Hz for 10 s). [B and C from Hawkins and Bruner 189.]

H(↓)
Figure 9. Figure 9.

Neural circuit mediating cockroach escape response and proposed sites of plasticity during habituation and sensitization. Primary sensory neurons (SN) make monosynaptic connections onto giant and nongiant interneurons (INT), which in turn project (possibly monosynaptically) to leg motoneurons (MN). Synaptic depression occurs at sensory neuron‐interneuron synapses as well as interneuron‐motoneuron synapses [H(↓)]. Synaptic facilitation [S(↑)] occurs at interneuron‐motoneuron synapses.

Figure 10. Figure 10.

Pigeon heart‐rate conditioning. A: schematic illustration of major ascending visual pathways mediating pigeon heart‐rate conditioning to visual stimuli. Acquisition of conditioned response can be prevented by lesions of retina, as indicated by L1; by combined lesions of principle optic nucleus, nucleus rotundus, and tectofugal fibers to nucleus dorsolateralis posterior of thalamus, as indicated by L2; and by combined lesions of striate and extrastriate cortical areas, as indicated by L3. During conditioning, output of retinal ganglion cells appears to remain constant, as indicated by (0). Increases and decreases in neuronal responses have been recorded in principal optic nucleus and nucleus rotundus (↑, ↓). B: effects of visual‐system lesions on acquisition of conditioned response. L1, L2, L3, performance of animals with combined lesions as indicated in A. Solid line, performance of control (CONT) animals. Each point, mean heart‐rate change in beats per minute (BPM) between 6‐s conditioned‐stimulus (CS) period and an immediately preceding 6‐s control period. C: training‐induced modifications of “type I” geniculate (principal optic nucleus) neurons as function of their response to training stimuli. I/I, cells that increase their discharge at CS onset and unconditioned‐stimulus (US) onset. I/II, cells that increase their discharge at CS onset but decrease their discharge at US onset. COND, units studied during associative training; SENS, units studied during nonassociative training. Curves, mean percentage change from response to light prior to training for phasic responses at CS onset. Note that type I/II cells showed marked response enhancement during associative training, while other cells showed only response decrement. [A adapted from Cohen 97; B adapted from Cohen 95; C from Cohen 96.]

Data for cells that decrease their discharge at CS onset (type II cells) are not shown.
Figure 11. Figure 11.

Eye‐blink conditioning in cat. A: proposed neural circuit. Unconditioned eye‐blink response (UR) to glabella tap unconditioned stimulus (US) is mediated by facial nucleus (VII), which receives input directly and/or indirectly from trigeminal nucleus (V). Conditioned response (CR) pathway has been proposed to involve coronal pericruciate sensorimotor cortex (SENS‐MOTOR CX). Lesions (L) of sensorimotor cortex block acquisition of CR. Increases in neural excitability occurring in sensorimotor cortex and facial nucleus (↑) have been proposed to be intrinsic to these regions and to contribute to response specificity during learning. Increases in neural responses to conditioned stimulus (CS) have been recorded in auditory association cortex (AUD ASSOC CX) and have been proposed to be due to changes in afferent or presynaptic elements and to contribute to stimulus specificity during conditioning. Lesions of caudal cortical areas, including auditory association cortex, do not prevent acquisition, and essential afferent pathways for click CS are not known. B: effects of bilateral lesions of cortical motor areas on acquisition of conditioned eye blink. Average percent conditioned responses (± SD) of normal cats (N) and cats with lesions of rostral cortex (L) during training of conditioned eye blink, n = 5 and n = 5, respectively. One additional animal (not included) whose lesion was confined to coronal pericruciate region reached 35% performance levels. C: mean ± SD threshold currents required to discharge “both” projective neurons of coronalpericruciate cortex with intracellular stimulation for 6 behavioral groups. Cortical cells were classified as “both” cells if their intracellular or extracellular stimulation evoked electromyogram responses in nose and eye musculature. Stippled bars, groups of cats that received paired CS‐US presentations, including extinction (US‐CS) group for whom order of stimuli was reversed during testing phase. Open bars, groups that received only CS‐alone trials or US‐alone trials. Delay groups were tested 3–28 days (DEL CS‐US) or 25–100 days (DEL US) after training and received no stimuli during testing phase. Asterisks, significant decreases (P < 0.05) in threshold compared with CS‐only group. US‐only group exhibited only moderate decreases in thresholds (P < 0.10). D: average changes in membrane resistance in cells of coronal pericruciate cortex given either iontophoresis of acetylcholine (ACh) plus depolarizing current sufficient to repeatedly discharge cell (ACh + DISCHARGE), iontophoresis of ACh alone (ACh ONLY), or current‐induced discharge only (DISCHARGE ONLY). Iontophoresis of ACh occurred from period of Ach to 0 as indicated on abscissa.

B from Woody et al. 437; C from Brons and Woody 57; D adapted from Woody et al. 435.
Figure 12. Figure 12.

Behavioral and electrophysiological changes during classical conditioning of rabbit nictitating membrane response for animals that received paired training (open circles) or explicitly unpaired training (closed circles). Each point, ∼23 training trials. Top: changes in magnitude of behavioral conditioned nictitating membrane response. Middle: changes in number of hippocampal multiple‐unit discharges occurring between conditioned stimulus onset and unconditioned stimulus onset. Bottom: changes in amplitude of population spike elicited in dentate granule cell layer by stimulation of perforant path. LTP, long‐term potentiation.

From Teyler and Discenna 385.
Figure 13. Figure 13.

Rabbit nictitating membrane conditioning. A: proposed neural circuitry underlying rabbit nictitating membrane and eyelid responses. Unconditioned‐response (UR) pathway is a di‐ or trisynaptic arc involving fifth sensory ganglion (VG), fifth sensory nuclei (VS), and motoneurons mediating response in accessory abducens, abducens, and facial nuclei (VIA, VI, and VIII). Secondary pathway through reticular formation (RF) has also been proposed. Lesions (L) of middle cerebellar peduncle (MCP), dentate‐interpositus nuclei (DENT INT), superior cerebellar peduncle (SCP) before and after point of decussation, and red nucleus (RED N) all abolish conditioned responses (CR) to tone or light but have little or no effect on unconditioned responses to corneal air puff, indicating that these areas are essential components of CR pathway (shaded structures). At present, role of cerebellar cortex is controversial. Conditioned‐stimulus (CS) pathway has been proposed to involve auditory input from cochlear nucleus (VIII) projecting ultimately to dorsolateral pontine nuclei (PONT N) and possibly lateral reticular nucleus, and their mossy fiber projections to cerebellum via middle cerebellar peduncle. Unconditioned‐stimulus (US) pathway has been proposed to involve somatosensory projections to inferior olive, which in turn projects to cerebellum via climbing fibers in inferior cerebellar peduncle (ICP). Training‐dependent increases in neuronal activity evoked by tone CS have been recorded in deep cerebellar nuclei (↑). Similar responses have also been recorded in cerebellar cortex and red nucleus. B: effects of unilateral lesions of dentate‐interpositus cerebellar nuclei on mean peak amplitude of CR and UR nictitating membrane (NM) responses. CR were measured in 250‐ms period between tone CS onset and air‐puff US onset; UR were measured in the 250‐ms period following US onset. Data presented in 4 periods of training trials per session, ∼27 trials per period. Animals received 3 days of training (P1–3) with US delivered to left eye, while movement of left NM was monitored (n = 14). Following lesions of left cerebellar nuclei, animals were trained for 4 more days (L1–4) to test for retention and recovery of CR. CR were almost completely abolished by lesion, but UR amplitude was unaffected. On fifth postlesion day (L5), US was switched to right (nonlesioned) side, and movement of right NM was monitored; training was then returned to left eye (n = 13). Right (nonlesioned) side learned quickly, controlling for nonspecific lesion effects, while conditioned responding on left side showed essentially no recovery. C: example of unusually robust change in neuronal unit activity recorded from medial dentate lateral interpositus nuclei during explicitly unpaired training (C1) and paired training (C2). Top trace, NM extension; bottom trace, peristimulus histogram (9‐ms bins) of multiple‐unit activity recorded from dentateinterpositus (DI) nuclei. Left vertical line, CS tone onset; right vertical line, US corneal air‐puff onset. C1: average of behavioral and neural responses to tone CS on first day of training in which tone CS and air‐puff US were presented in explicitly unpaired fashion. C2: responses to tone CS on second day of paired training. Onset of unit response preceded behavioral NM response within a trial by 36–58 ms. Note that pattern of unit discharge appears to parallel topography of conditioned behavioral response more closely than unconditioned behavioral response evoked by air puff. Stimulation through recording electrode elicited ipsilateral eyelid closure and NM extension. D: comparison of changes in conditioned NM response and neuronal activity recorded from dentate‐interpositus in second half of CS period over course of training (n = 7). Magnitude of conditioned NM response (dashed line) was measured as area under curve described by amplitude time course of NM response in millimeter milliseconds. Standard scores of unit activity were calculated by comparing neuronal response in CS period to pre‐CS background period according to following equation: (CS half‐block – pre‐CS half‐block)/(SD pre‐CS session). Note that neuronal responses of dentate‐interpositus nuclei increase in close relation to size of CR (r = 0.90).

A adapted from Thompson 387; B from Clark et al. 91; C and D adapted from McCormick and Thompson 298.
Figure 14. Figure 14.

Cat leg‐flexion conditioning. A: experimental arrangement and proposed neural circuit. Conditioned stimulus (CS) is train of pulses to cerebral peduncle (CP); unconditioned stimulus (US) is shock to forelimb. Cerebral peduncle is cut caudal to red nucleus to restrict corticofugal output to red nucleus. Red nucleus neurons receive monosynaptic excitatory synaptic input onto distal dendritic regions from cerebral cortex and from nucleus interpositus of cerebellum (IP) onto proximal dendritic and somatic regions. Red nucleus projects via interneurons (INT) to flexor motoneurons innervating biceps brachii muscle controlling leg flexion. Conditioning in this paradigm has been proposed to involve sprouting of new corticorubral synapses onto proximal dendritic and somatic regions of cells in red nucleus (dashed connection, ↑). B: example of acquisition and extinction of conditioned leg flexion in 1 cat. During acquisition, subject received paired CS‐US training for 10 sessions; during extinction, animal received backward US‐CS trails. C: physiological evidence for sprouting of new corticorubral synapses following conditioning. EPSPs elicited in red nucleus neurons by stimulation of cerebral peduncle in conditioned cat (C1) exhibit component with relatively fast rise time, whereas EPSPs elicited by stimulation of cerebral peduncle in control cat that had received no training (C2) exhibit relatively slow rise time. Frequency histograms of time to peak for EPSP elicited by stimulation of cerebral peduncle in conditioned cat (C3) and nontrained control cat (C4). Ordinate, number of cells; abscissa, time to peak for EPSPs.

A from Tsukahara 395; B from Tsukahara et al. 402; C from Tsukahara and Oda 401.
Figure 15. Figure 15.

A: experimental arrangement for leg‐position learning in locust. Animal given paired training (P) receives electric shock from stimulator (STIM) when leg is lowered below a defined position. Since the 2 animals are arranged in series, yoked control (R) receives same shock irrespective of where it holds its leg. B: effects of up‐training on firing rate of AAdC, a motoneuron controlling leg position. Demand level was set at 15 Hz, and 10 shocks were delivered. 1, Before up‐training, firing rate in high Mg2+‐zero Ca2+ solution was 11.4 Hz. 2, In normal saline, firing rate was similar (11.0 Hz). 3, After up‐training, firing rate of AAdC had increased to 28.5 Hz. 4, After second infusion of high Mg2+‐zero Ca2+, firing rate was 22.1 Hz. The fact that firing rate remained elevated in high Mg2+‐zero Ca2+ solution (which inhibits chemical synaptic transmission) suggests that some of the increase was due to intrinsic changes in AAdC motoneuron. C: effect of up‐training on input resistance of AAdC motoneurons. Intracellular constant current pulse (bottom trace) elicits hyperpolarization of AAdC motoneuron (top trace). Note increase in size of voltage step (sharp downward deflection) after up‐training (C2) relative to voltage step in naive preparation (C1). Note also change in action potentials (small up‐ward deflections) after training.

A adapted from Horridge 204; B adapted from Woollacott and Hoyle 439; C adapted from Hoyle 208.
Figure 16. Figure 16.

Partial neural circuit mediating Pleurobranchaea feeding response and sites of altered neural response following aversive conditioning. Chemosensory neurons (SN) in oral veil produce both indirect excitation and inhibition of command neurons (CN) for feeding. Conditioned food stimuli produce less excitation and more inhibition (↓) in command neurons in animals that have had paired training than in animals that have had unpaired training or naive animals. This effect is partially accounted for by greater inhibition (↓) of interneurons (INT) that excite feeding command cells and greater excitation (↑) of interneurons that inhibit feeding command cells. MN, motoneuron. Triangles, excitatory synapses; circles, inhibitory synapses.

Figure 17. Figure 17.

Classical conditioning in Hermissenda. A: acquisition, retention, and reacquisition of changes in response latencies to enter a test light following different training paradigms (RR, random rotation; RL, random light; ULR, unpaired light and rotation; RLR, random light and rotation; NLR, no light or rotation; PLR, paired light and rotation). Median response ratio compared latency during test (A) with pretraining base‐line response latency (B); values below 0.5 indicate increase in response latency. Group receiving paired light and rotation showed significantly greater suppression of phototactic response during both acquisition and retention phases than did control groups, indicating conditioning had occurred. During reacquisition, the 3 groups shown were subject to paired training. B: proposed neural circuit and sites of plasticity mediating conditioning in Hermissenda. Open circles, inhibitory synapses; open triangles, excitatory synapses. Following pairing of a light CS and a rotation US with caudal orientation [which excites caudal hair cells (CH)], there is an increase in excitability of type B photoreceptors (↑) and a decrease in excitability of type A photoreceptors (↓). Electrically coupled S and E cells of optic ganglion are shown as single cell. [A from Crow and Alkon 108, copyright 1978 by the American Association for the Advancement of Science; B adapted from Alkon 8.]

1 — A/(A + B)
Figure 18. Figure 18.

Electrophysiological analyses of changes in type B photoreceptors as function of conditioning in Hermissenda. A: comparison of input resistance of isolated (cut‐nerve), dark‐adapted type B photoreceptors from animals given paired or random presentations of light and rotation. 1, Representative linear current‐voltage relationship from experimental and random control groups. 2, Example of changes in membrane potential (top trace) to hyperpolarizing current pulses (bottom trace) in paired experimental and random control animals. Change in membrane potential to a given current pulse is larger following paired training relative to random control. B: example of voltage‐dependent outward currents in type B cells isolated from paired, random, and naive animals. Command pulses to 0 mV (bottom trace) elicit an early, inactivating K+ current (IA) and a more slowly inactivating Ca2+‐dependent K+ current () (top trace). Records were chosen to illustrate reduction of IA and for paired animals as compared with random and naive animals. C: intracellular voltage recordings of type B photoreceptors during and after presentation of unpaired light and rotation (1), light alone (2), or paired light and rotation (3). Responses to second of 2 succeeding 30‐s light steps are shown. Square pulse (solid line), light step. Ramped pulse (connected circles), rotational stimulus. Dashed lines, cell's initial resting potential preceding first light step; shaded areas, depolarization above resting level after second light step.

A from Crow and Alkon 109, copyright 1980 by the American Association for the Advancement of Science; B from Alkon et al. 18; C from Alkon 10.
Figure 19. Figure 19.

Differential classical conditioning of siphon‐withdrawal response in Aplysia. A: experimental preparation. Either siphon or mantle shelf could serve as conditioned stimulus (CS). Unconditioned stimulus (US) was an electric shock to tail. For illustrative purposes, parapodia and mantle shelf are shown retracted; however, all experiments were conducted using freely moving animals whose parapodia had been surgically removed. B: experimental paradigm. One group (SIPHON+) received siphon CS (CS+) followed by tail‐shock US and specifically unpaired mantle CS (CS). Other group (MANTLE+) received mantle CS (CS+) followed by tail‐shock US and specifically unpaired siphon CS (CS). Intertrial interval was 5 min. C: results from differential conditioning experiment using paradigm shown in B (n = 12). Testing was carried out 30 min after 15 training trials. SIPHON+ animals showed significantly longer siphon‐withdrawal responses to siphon CS than mantle CS (P < 0.05), while MANTLE+ animals showed significantly longer siphon‐withdrawal responses to mantle CS than to siphon CS (P < 0.01). Data in C and D expressed as means ± SE. D: pooled data from C. Test scores from CS and CS+ pathways following training are compared with their respective test scores obtained before training. Responses to stimuli delivered to CS+ pathway were significantly longer than responses to stimuli to CS pathway (P < 0.001), indicating differential conditioning had occurred.

From Carew et al. 74, copyright 1983 by the American Association for the Advancement of Science.
Figure 20. Figure 20.

Electrophysiological analysis of classical conditioning of siphon‐withdrawal reflex in Aplysia. A: experimental arrangement (A1) and training protocol (A2). Intracellular recordings were obtained in vitro from 2 different sensory neurons (SN) and siphon motoneuron (MN) to which they both projected. EPSP elicited in motoneuron by each sensory neuron was tested before and after training. Training consisted of trials in which train of spikes in 1 SN (paired SN) was followed by electric tail‐shock US; train of spikes in other SN (unpaired SN) was explicitly unpaired. Intertrial interval was 5 min. B: examples of EPSPs produced in siphon MN by paired SN and unpaired SN before (PRE) and 1 h after (POST) training protocol shown in A. Note that there is greater facilitation of EPSP elicited by paired SN than by unpaired SN. C: example of differential broadening of action potential in paired SN relative to an unpaired SN in the same preparation. Action potentials recorded before (PRE) and 3 h after (POST) 15 training trials using protocol shown in A. Experiment conducted in 50 mM TEA. D: removal of extracellular Ca2+ diminishes spike broadening produced by serotonin (5‐HT) paired with spike activity. Trains of action potentials in single SN were paired with brief (‐1 s) applications of serotonin alternately in normal seawater (10−2 M Ca2+) and in Ca2+‐free seawater. Action‐potential duration was measured in normal seawater before and after pairing in each solution. Note that after action potentials were paired with serotonin application in normal seawater, increase in action‐potential duration was significantly greater than after pairing in absence of Ca2+ influx (n = 8, P < 0.001).

A, B, and C from Hawkins et al. 188, copyright 1983 by the American Association for the Advancement of Science; D from Abrams 1.


Figure 1.

Analysis of habituation and sensitization of hindlimb flexion response in acute spinal cat. A: experimental arrangement. Stimuli can be delivered to skin (SS1 or SS2), to cutaneous afferent nerve (SN1), or to group Ia muscle afferents (SN2). Responses can be recorded from dorsal root (RS), from α‐motoneurons (Rα) or interneurons, from ventral root (RN), or from muscle (RM). B: comparison of changes in excitatory postsynaptic potentials (EPSPs) recorded in deep peroneal motoneuron elicited by stimulation of either cutaneous afferent nerve (1, 2, 3) or muscle afferents (4, 5, 6). Stimulation of the 2 nerves was interpolated. 1, 2, 3, Polysynaptic EPSPs elicited by single‐shock stimuli delivered to posterior femoral cutaneous nerve. 1, Control period, stimulation once every 30 s. 2, Synaptic decrement, stimulation once per second. 3, Recovery, stimulation once every 30 s. 4, 5, 6, Monosynaptic EPSPs elicited by stimulation of deep peroneal nerve at 30‐s intervals during periods corresponding to those in 1, 2, and 3. Note constancy of monosynaptic responses during period of polysynaptic EPSP decrement.

A adapted from Thompson 386; B adapted from Spencer et al. 375.


Figure 2.

Possible role of spinal interneurons in habituation and sensitization of flexion reflex in cat. A: responses of different types of interneurons (nonplastic, type H, and type S) and muscle response during repetitive stimulation producing first sensitization and then habituation of flexion reflex. B: schematic diagram of possible neural substrates of habituation and sensitization. N, nonplastic synapses; H, habituating synapses; S, sensitizing synapses. According to this scheme, cutaneous stimuli can exert their influence on 2 separate systems: reflex system (S‐R pathway) that mediates habituation and “state” system that mediates sensitization.

From Groves and Thompson 181.


Figure 3.

Habituation and sensitization of gill‐ and siphon‐withdrawal responses in Aplysia. A: experimental arrangement for behavioral studies in intact animal showing gill and siphon in relaxed (A1) and contracted (A2) position. Parapodia and mantle shelf are shown retracted to reveal gill and siphon. Gill and siphon withdrawal can be elicited with tactile stimulus to siphon. Electric shock of tail (shock) or head serves as sensitizing stimulus. B: photocell recordings showing habituation and sensitization of gill‐withdrawal response. Tactile stimulus to siphon (bottom trace) elicited gill‐withdrawal response (top trace), which habituated with repeated stimulus deliveries. Application of noxious tail shock (arrow) caused marked sensitization of gill‐withdrawal response. C: neural circuit mediating gill‐ and siphon‐withdrawal responses. Siphon skin is innervated by group of ∼24 mechanoreceptor sensory cells (SN) whose soma are located in LE cluster in abdominal ganglion. Siphon sensory cells project both monosynaptically and through interneurons (INT) to gill and siphon motoneurons (MN), many of which are also located in abdominal ganglion. Effects of sensitizing stimuli in abdominal ganglion are mediated by group of facilitator interneurons (FAC INT) that terminate on sensory cells in synaptic and somatic regions. Habituation of reflex has been shown to involve a decrease in transmitter release from sensory neuron terminals [H(↓)]. Sensitization of reflex has been shown to involve an increase in transmitter release from sensory neurons [S(↑)]. Posttetanic potentiation has also been reported to occur at neuromuscular junction. [Adapted from Kandel and Schwartz 220.]

interstimulus interval (ISI), 1.5 min


Figure 4.

Electrophysiological analyses of habituation and sensitization of gill‐withdrawal response. A: homosynaptic depression and presynaptic facilitation at synapses from a siphon sensory cell onto L7, a gill motoneuron. Abdominal ganglion was dissected from animal and was maintained in vitro in artificial seawater (ASW). Intracellular stimulation of sensory neuron triggered action potential (bottom trace), which elicited monosynaptic EPSP in gill motoneuron (top trace). (Sharp negative and positive deflections that precede and follow action potential in sensory neuron are stimulus artifacts from depolarizing current pulse.) With repeated stimulation, action potential elicited successively smaller EPSPs in gill motoneuron. Electrical stimulation of left connective (arrow), which carries input to abdominal ganglion from head and tail, caused a marked increase in EPSP elicited in gill motoneuron. Synaptic depression and facilitation parallel and contribute to behavioral habituation and sensitization of gill‐withdrawal reflex. B: connective stimulation (B1) and application of 0.2 mM serotonin (B2) increase action‐potential duration in siphon sensory cells. Abdominal ganglion was maintained in artificial seawater containing 0.1 M tetraethylammonium (TEA) (K+ channel blocker) to facilitate measurement of action‐potential duration and maximize effects of facilitation. Similar though smaller changes also occur in normal seawater. Because Ca2+ entry is voltage dependent, broadening of action potential is believed to contribute to increase in Ca2+ influx and subsequent transmitter release during facilitation. C: comparison of changes in action‐potential duration in siphon sensory cell (bottom trace) and amplitude of EPSPs elicited in gill motoneuron L7 (top trace) in 1 preparation. With repeated stimulation, action‐potential duration and EPSP amplitude decrease concomitantly. Weak stimulation of pleuroabdominal connective (left vertical dashed line) causes a transient increase in action‐potential duration and EPSP amplitude. Stronger stimulation (right vertical dashed line) produces larger effects. Experiment conducted in 0.1 M TEA.

A adapted from Castellucci and Kandel 81; B from Klein and Kandel 233; C adapted from Kandel 218.


Figure 5.

Biophysical analyses of ionic changes underlying homosynaptic depression and presynaptic facilitation in siphon sensory cells. A: serotonin, a putative facilitatory transmitter, depresses an outward K+ current in siphon sensory cells. A1: voltage‐clamp experiment, performed in normal seawater. Voltage step from −35 to +15 mV (bottom trace) elicits initial inward current (downward deflection, top trace) carried largely by Na+ and Ca2+, followed by slower outward current (upward deflection, top trace) carried largely by K+. With repeated stimulation, inward current decreases, paralleling habituation. Application of serotonin (arrow) causes marked decrease in outward current. A2: voltage‐clamp experiment in presence of K+ channel blockers (TEA, tetraethylammonium; 4‐AP, 4‐aminopyridine). Same preparation as in A1. With K+ currents blocked, depolarizing step (bottom trace) elicits primarily an inward current. Application of serotonin (arrow) had no effect on current elicited by voltage step, indicating that serotonin normally exerts its action by depression of outward K+ current. B: effect of serotonin (5‐HT) on single‐channel currents recorded from siphon sensory cells in Aplysia. Left ordinate, number of open channels (n); right ordinate, current magnitude. Individual current steps are 2.6 pA. 1, Current in absence of serotonin. Current record was well fit by binomial distribution assuming that 5 channels are active in the patch and that each channel opens with probability of 0.84. 2, Current recorded 2 min after addition of 30 μM serotonin to bath. 3, Current recorded 1 min after addition of further dose of serotonin, raising total concentration to 60 μM. The 2 traces are a continuous recording. Downward current deflection at end of bottom trace is from intracellular current pulse used to monitor cell resistance. Serotonin produced decrease in average number of open channels but had no effect on unit current. 4, Current recorded ∼5 min after superfusion with serotonin‐free artificial seawater (ASW). Note increase in number of open channels.

A adapted from Klein and Kandel 234; B adapted from Siegelbaum et al. 364.


Figure 6.

Biochemical analyses of presynaptic facilitation in siphon sensory cells. A: injection of cAMP‐dependent protein kinase produces presynaptic facilitation. In base‐line control period, intracellular stimulation of sensory neuron (SN, bottom trace) evoked EPSP in a follower neuron (FN, top trace). Injection of catalytic subunit of cAMP‐dependent protein kinase led to increase in duration of sensory neuron action potential and increase in amplitude of EPSP evoked in follower cell. Experiment conducted in 100 mM tetraethylammonium (TEA). B: injection of kinase inhibitor blocks serotonin‐induced spike broadening. Recordings of action potentials in 2 different sensory neurons before (left) and after (right) application of serotonin. Serotonin produced marked spike broadening in control sensory cell but relatively little spike broadening in sensory neuron that had been injected with Walsh inhibitor, which blocks activation of cAMP‐dependent protein kinase. Experiment conducted in TEA. C: proposed sequence of steps underlying presynaptic facilitation in siphon sensory cells.

A from Castellucci et al. 82; B from Castellucci et al. 83.


Figure 7.

Analysis of habituation of crayfish escape response. A: neural circuit mediating escape response and proposed sites of plasticity. Tactile stimulation of posterior parts of crayfish produces forward somersault that is mediated by pair of neurons called lateral giants (LG). The LG have rectifying electrical synapse onto motor giants (MO G) and electrical synapse onto fast flexor motoneurons (FF MN). The LG receive direct electrical connections from sensory neurons (SN) as well as di‐ or trisynaptic input through heterogeneous group of interneurons (INT). These connections are believed to mediate the α‐ and β‐components, respectively, of complex EPSP evoked in LG by tactile stimuli. Habituation has been shown to involve a decrease in transmitter release at chemical synapses from primary sensory neurons onto interneurons. Synaptic depression can also occur at motor giant neuromuscular junction (↓) and posttetanic potentiation has been observed at neuromuscular junction of fast flexor motoneuron (↑); however, contribution of these forms of plasticity to behavioral learning is not clearly established. B: example of decrement in responses recorded from lateral giant fibers in response to repeated stimulation (1/min) of second root of abdominal nerve cord. Initial stimuli evoke action potential that fails as the β‐component of complex EPSP grows smaller with successive stimuli. Note that α‐component of EPSP remains relatively stable. C: example of decrement of monosynaptic EPSP in tactile interneuron. Stimulation of second root was adjusted to recruit single tactile afferent that elicited monosynaptic EPSP in tactile interneuron. With repeated stimulation, amplitude of EPSP decreases. C1, C2, C3: first, fourth, and seventh responses to stimulation at 0.5 Hz. [A and C adapted from Zucker 450,451; B adapted from Krasne 239.]

H(↓)


Figure 8.

Habituation and sensitization of crayfish defense response to tactile stimuli. A: proposed neural circuit and possible sites of plasticity. During habituation, neural responses elicited by tactile stimulus to back or tail decrease in interneurons (INT) and in excitor motoneuron (E) but do not change in inhibitor motoneuron (I) [H(0)]; these results suggest that habituation is due to changes in excitatory reflex pathway and not to buildup of peripheral inhibition. Sensitization has been proposed to involve increase in responses of interneurons and excitor motoneurons [S(↑)]; posttetanic potentiation (PTP) at neuromuscular junction of excitor motoneurons [S(↑)]; and decrease in responses of inhibitor motoneuron [S(↓)]. SN, sensory neuron; CN, command neuron. B: changes in number of excitor and inhibitor spikes accompanying sensitization. Note increase of excitor responses and decrease of inhibitor responses. C: PTP at excitor neuromuscular junction. Excitatory junction potentials recorded during test trains of excitor stimulation (30 Hz for 0.5 s) before (top trace) and 5 s after (bottom trace) conditioning train of excitor stimulation (10 Hz for 10 s). [B and C from Hawkins and Bruner 189.]

H(↓)


Figure 9.

Neural circuit mediating cockroach escape response and proposed sites of plasticity during habituation and sensitization. Primary sensory neurons (SN) make monosynaptic connections onto giant and nongiant interneurons (INT), which in turn project (possibly monosynaptically) to leg motoneurons (MN). Synaptic depression occurs at sensory neuron‐interneuron synapses as well as interneuron‐motoneuron synapses [H(↓)]. Synaptic facilitation [S(↑)] occurs at interneuron‐motoneuron synapses.



Figure 10.

Pigeon heart‐rate conditioning. A: schematic illustration of major ascending visual pathways mediating pigeon heart‐rate conditioning to visual stimuli. Acquisition of conditioned response can be prevented by lesions of retina, as indicated by L1; by combined lesions of principle optic nucleus, nucleus rotundus, and tectofugal fibers to nucleus dorsolateralis posterior of thalamus, as indicated by L2; and by combined lesions of striate and extrastriate cortical areas, as indicated by L3. During conditioning, output of retinal ganglion cells appears to remain constant, as indicated by (0). Increases and decreases in neuronal responses have been recorded in principal optic nucleus and nucleus rotundus (↑, ↓). B: effects of visual‐system lesions on acquisition of conditioned response. L1, L2, L3, performance of animals with combined lesions as indicated in A. Solid line, performance of control (CONT) animals. Each point, mean heart‐rate change in beats per minute (BPM) between 6‐s conditioned‐stimulus (CS) period and an immediately preceding 6‐s control period. C: training‐induced modifications of “type I” geniculate (principal optic nucleus) neurons as function of their response to training stimuli. I/I, cells that increase their discharge at CS onset and unconditioned‐stimulus (US) onset. I/II, cells that increase their discharge at CS onset but decrease their discharge at US onset. COND, units studied during associative training; SENS, units studied during nonassociative training. Curves, mean percentage change from response to light prior to training for phasic responses at CS onset. Note that type I/II cells showed marked response enhancement during associative training, while other cells showed only response decrement. [A adapted from Cohen 97; B adapted from Cohen 95; C from Cohen 96.]

Data for cells that decrease their discharge at CS onset (type II cells) are not shown.


Figure 11.

Eye‐blink conditioning in cat. A: proposed neural circuit. Unconditioned eye‐blink response (UR) to glabella tap unconditioned stimulus (US) is mediated by facial nucleus (VII), which receives input directly and/or indirectly from trigeminal nucleus (V). Conditioned response (CR) pathway has been proposed to involve coronal pericruciate sensorimotor cortex (SENS‐MOTOR CX). Lesions (L) of sensorimotor cortex block acquisition of CR. Increases in neural excitability occurring in sensorimotor cortex and facial nucleus (↑) have been proposed to be intrinsic to these regions and to contribute to response specificity during learning. Increases in neural responses to conditioned stimulus (CS) have been recorded in auditory association cortex (AUD ASSOC CX) and have been proposed to be due to changes in afferent or presynaptic elements and to contribute to stimulus specificity during conditioning. Lesions of caudal cortical areas, including auditory association cortex, do not prevent acquisition, and essential afferent pathways for click CS are not known. B: effects of bilateral lesions of cortical motor areas on acquisition of conditioned eye blink. Average percent conditioned responses (± SD) of normal cats (N) and cats with lesions of rostral cortex (L) during training of conditioned eye blink, n = 5 and n = 5, respectively. One additional animal (not included) whose lesion was confined to coronal pericruciate region reached 35% performance levels. C: mean ± SD threshold currents required to discharge “both” projective neurons of coronalpericruciate cortex with intracellular stimulation for 6 behavioral groups. Cortical cells were classified as “both” cells if their intracellular or extracellular stimulation evoked electromyogram responses in nose and eye musculature. Stippled bars, groups of cats that received paired CS‐US presentations, including extinction (US‐CS) group for whom order of stimuli was reversed during testing phase. Open bars, groups that received only CS‐alone trials or US‐alone trials. Delay groups were tested 3–28 days (DEL CS‐US) or 25–100 days (DEL US) after training and received no stimuli during testing phase. Asterisks, significant decreases (P < 0.05) in threshold compared with CS‐only group. US‐only group exhibited only moderate decreases in thresholds (P < 0.10). D: average changes in membrane resistance in cells of coronal pericruciate cortex given either iontophoresis of acetylcholine (ACh) plus depolarizing current sufficient to repeatedly discharge cell (ACh + DISCHARGE), iontophoresis of ACh alone (ACh ONLY), or current‐induced discharge only (DISCHARGE ONLY). Iontophoresis of ACh occurred from period of Ach to 0 as indicated on abscissa.

B from Woody et al. 437; C from Brons and Woody 57; D adapted from Woody et al. 435.


Figure 12.

Behavioral and electrophysiological changes during classical conditioning of rabbit nictitating membrane response for animals that received paired training (open circles) or explicitly unpaired training (closed circles). Each point, ∼23 training trials. Top: changes in magnitude of behavioral conditioned nictitating membrane response. Middle: changes in number of hippocampal multiple‐unit discharges occurring between conditioned stimulus onset and unconditioned stimulus onset. Bottom: changes in amplitude of population spike elicited in dentate granule cell layer by stimulation of perforant path. LTP, long‐term potentiation.

From Teyler and Discenna 385.


Figure 13.

Rabbit nictitating membrane conditioning. A: proposed neural circuitry underlying rabbit nictitating membrane and eyelid responses. Unconditioned‐response (UR) pathway is a di‐ or trisynaptic arc involving fifth sensory ganglion (VG), fifth sensory nuclei (VS), and motoneurons mediating response in accessory abducens, abducens, and facial nuclei (VIA, VI, and VIII). Secondary pathway through reticular formation (RF) has also been proposed. Lesions (L) of middle cerebellar peduncle (MCP), dentate‐interpositus nuclei (DENT INT), superior cerebellar peduncle (SCP) before and after point of decussation, and red nucleus (RED N) all abolish conditioned responses (CR) to tone or light but have little or no effect on unconditioned responses to corneal air puff, indicating that these areas are essential components of CR pathway (shaded structures). At present, role of cerebellar cortex is controversial. Conditioned‐stimulus (CS) pathway has been proposed to involve auditory input from cochlear nucleus (VIII) projecting ultimately to dorsolateral pontine nuclei (PONT N) and possibly lateral reticular nucleus, and their mossy fiber projections to cerebellum via middle cerebellar peduncle. Unconditioned‐stimulus (US) pathway has been proposed to involve somatosensory projections to inferior olive, which in turn projects to cerebellum via climbing fibers in inferior cerebellar peduncle (ICP). Training‐dependent increases in neuronal activity evoked by tone CS have been recorded in deep cerebellar nuclei (↑). Similar responses have also been recorded in cerebellar cortex and red nucleus. B: effects of unilateral lesions of dentate‐interpositus cerebellar nuclei on mean peak amplitude of CR and UR nictitating membrane (NM) responses. CR were measured in 250‐ms period between tone CS onset and air‐puff US onset; UR were measured in the 250‐ms period following US onset. Data presented in 4 periods of training trials per session, ∼27 trials per period. Animals received 3 days of training (P1–3) with US delivered to left eye, while movement of left NM was monitored (n = 14). Following lesions of left cerebellar nuclei, animals were trained for 4 more days (L1–4) to test for retention and recovery of CR. CR were almost completely abolished by lesion, but UR amplitude was unaffected. On fifth postlesion day (L5), US was switched to right (nonlesioned) side, and movement of right NM was monitored; training was then returned to left eye (n = 13). Right (nonlesioned) side learned quickly, controlling for nonspecific lesion effects, while conditioned responding on left side showed essentially no recovery. C: example of unusually robust change in neuronal unit activity recorded from medial dentate lateral interpositus nuclei during explicitly unpaired training (C1) and paired training (C2). Top trace, NM extension; bottom trace, peristimulus histogram (9‐ms bins) of multiple‐unit activity recorded from dentateinterpositus (DI) nuclei. Left vertical line, CS tone onset; right vertical line, US corneal air‐puff onset. C1: average of behavioral and neural responses to tone CS on first day of training in which tone CS and air‐puff US were presented in explicitly unpaired fashion. C2: responses to tone CS on second day of paired training. Onset of unit response preceded behavioral NM response within a trial by 36–58 ms. Note that pattern of unit discharge appears to parallel topography of conditioned behavioral response more closely than unconditioned behavioral response evoked by air puff. Stimulation through recording electrode elicited ipsilateral eyelid closure and NM extension. D: comparison of changes in conditioned NM response and neuronal activity recorded from dentate‐interpositus in second half of CS period over course of training (n = 7). Magnitude of conditioned NM response (dashed line) was measured as area under curve described by amplitude time course of NM response in millimeter milliseconds. Standard scores of unit activity were calculated by comparing neuronal response in CS period to pre‐CS background period according to following equation: (CS half‐block – pre‐CS half‐block)/(SD pre‐CS session). Note that neuronal responses of dentate‐interpositus nuclei increase in close relation to size of CR (r = 0.90).

A adapted from Thompson 387; B from Clark et al. 91; C and D adapted from McCormick and Thompson 298.


Figure 14.

Cat leg‐flexion conditioning. A: experimental arrangement and proposed neural circuit. Conditioned stimulus (CS) is train of pulses to cerebral peduncle (CP); unconditioned stimulus (US) is shock to forelimb. Cerebral peduncle is cut caudal to red nucleus to restrict corticofugal output to red nucleus. Red nucleus neurons receive monosynaptic excitatory synaptic input onto distal dendritic regions from cerebral cortex and from nucleus interpositus of cerebellum (IP) onto proximal dendritic and somatic regions. Red nucleus projects via interneurons (INT) to flexor motoneurons innervating biceps brachii muscle controlling leg flexion. Conditioning in this paradigm has been proposed to involve sprouting of new corticorubral synapses onto proximal dendritic and somatic regions of cells in red nucleus (dashed connection, ↑). B: example of acquisition and extinction of conditioned leg flexion in 1 cat. During acquisition, subject received paired CS‐US training for 10 sessions; during extinction, animal received backward US‐CS trails. C: physiological evidence for sprouting of new corticorubral synapses following conditioning. EPSPs elicited in red nucleus neurons by stimulation of cerebral peduncle in conditioned cat (C1) exhibit component with relatively fast rise time, whereas EPSPs elicited by stimulation of cerebral peduncle in control cat that had received no training (C2) exhibit relatively slow rise time. Frequency histograms of time to peak for EPSP elicited by stimulation of cerebral peduncle in conditioned cat (C3) and nontrained control cat (C4). Ordinate, number of cells; abscissa, time to peak for EPSPs.

A from Tsukahara 395; B from Tsukahara et al. 402; C from Tsukahara and Oda 401.


Figure 15.

A: experimental arrangement for leg‐position learning in locust. Animal given paired training (P) receives electric shock from stimulator (STIM) when leg is lowered below a defined position. Since the 2 animals are arranged in series, yoked control (R) receives same shock irrespective of where it holds its leg. B: effects of up‐training on firing rate of AAdC, a motoneuron controlling leg position. Demand level was set at 15 Hz, and 10 shocks were delivered. 1, Before up‐training, firing rate in high Mg2+‐zero Ca2+ solution was 11.4 Hz. 2, In normal saline, firing rate was similar (11.0 Hz). 3, After up‐training, firing rate of AAdC had increased to 28.5 Hz. 4, After second infusion of high Mg2+‐zero Ca2+, firing rate was 22.1 Hz. The fact that firing rate remained elevated in high Mg2+‐zero Ca2+ solution (which inhibits chemical synaptic transmission) suggests that some of the increase was due to intrinsic changes in AAdC motoneuron. C: effect of up‐training on input resistance of AAdC motoneurons. Intracellular constant current pulse (bottom trace) elicits hyperpolarization of AAdC motoneuron (top trace). Note increase in size of voltage step (sharp downward deflection) after up‐training (C2) relative to voltage step in naive preparation (C1). Note also change in action potentials (small up‐ward deflections) after training.

A adapted from Horridge 204; B adapted from Woollacott and Hoyle 439; C adapted from Hoyle 208.


Figure 16.

Partial neural circuit mediating Pleurobranchaea feeding response and sites of altered neural response following aversive conditioning. Chemosensory neurons (SN) in oral veil produce both indirect excitation and inhibition of command neurons (CN) for feeding. Conditioned food stimuli produce less excitation and more inhibition (↓) in command neurons in animals that have had paired training than in animals that have had unpaired training or naive animals. This effect is partially accounted for by greater inhibition (↓) of interneurons (INT) that excite feeding command cells and greater excitation (↑) of interneurons that inhibit feeding command cells. MN, motoneuron. Triangles, excitatory synapses; circles, inhibitory synapses.



Figure 17.

Classical conditioning in Hermissenda. A: acquisition, retention, and reacquisition of changes in response latencies to enter a test light following different training paradigms (RR, random rotation; RL, random light; ULR, unpaired light and rotation; RLR, random light and rotation; NLR, no light or rotation; PLR, paired light and rotation). Median response ratio compared latency during test (A) with pretraining base‐line response latency (B); values below 0.5 indicate increase in response latency. Group receiving paired light and rotation showed significantly greater suppression of phototactic response during both acquisition and retention phases than did control groups, indicating conditioning had occurred. During reacquisition, the 3 groups shown were subject to paired training. B: proposed neural circuit and sites of plasticity mediating conditioning in Hermissenda. Open circles, inhibitory synapses; open triangles, excitatory synapses. Following pairing of a light CS and a rotation US with caudal orientation [which excites caudal hair cells (CH)], there is an increase in excitability of type B photoreceptors (↑) and a decrease in excitability of type A photoreceptors (↓). Electrically coupled S and E cells of optic ganglion are shown as single cell. [A from Crow and Alkon 108, copyright 1978 by the American Association for the Advancement of Science; B adapted from Alkon 8.]

1 — A/(A + B)


Figure 18.

Electrophysiological analyses of changes in type B photoreceptors as function of conditioning in Hermissenda. A: comparison of input resistance of isolated (cut‐nerve), dark‐adapted type B photoreceptors from animals given paired or random presentations of light and rotation. 1, Representative linear current‐voltage relationship from experimental and random control groups. 2, Example of changes in membrane potential (top trace) to hyperpolarizing current pulses (bottom trace) in paired experimental and random control animals. Change in membrane potential to a given current pulse is larger following paired training relative to random control. B: example of voltage‐dependent outward currents in type B cells isolated from paired, random, and naive animals. Command pulses to 0 mV (bottom trace) elicit an early, inactivating K+ current (IA) and a more slowly inactivating Ca2+‐dependent K+ current () (top trace). Records were chosen to illustrate reduction of IA and for paired animals as compared with random and naive animals. C: intracellular voltage recordings of type B photoreceptors during and after presentation of unpaired light and rotation (1), light alone (2), or paired light and rotation (3). Responses to second of 2 succeeding 30‐s light steps are shown. Square pulse (solid line), light step. Ramped pulse (connected circles), rotational stimulus. Dashed lines, cell's initial resting potential preceding first light step; shaded areas, depolarization above resting level after second light step.

A from Crow and Alkon 109, copyright 1980 by the American Association for the Advancement of Science; B from Alkon et al. 18; C from Alkon 10.


Figure 19.

Differential classical conditioning of siphon‐withdrawal response in Aplysia. A: experimental preparation. Either siphon or mantle shelf could serve as conditioned stimulus (CS). Unconditioned stimulus (US) was an electric shock to tail. For illustrative purposes, parapodia and mantle shelf are shown retracted; however, all experiments were conducted using freely moving animals whose parapodia had been surgically removed. B: experimental paradigm. One group (SIPHON+) received siphon CS (CS+) followed by tail‐shock US and specifically unpaired mantle CS (CS). Other group (MANTLE+) received mantle CS (CS+) followed by tail‐shock US and specifically unpaired siphon CS (CS). Intertrial interval was 5 min. C: results from differential conditioning experiment using paradigm shown in B (n = 12). Testing was carried out 30 min after 15 training trials. SIPHON+ animals showed significantly longer siphon‐withdrawal responses to siphon CS than mantle CS (P < 0.05), while MANTLE+ animals showed significantly longer siphon‐withdrawal responses to mantle CS than to siphon CS (P < 0.01). Data in C and D expressed as means ± SE. D: pooled data from C. Test scores from CS and CS+ pathways following training are compared with their respective test scores obtained before training. Responses to stimuli delivered to CS+ pathway were significantly longer than responses to stimuli to CS pathway (P < 0.001), indicating differential conditioning had occurred.

From Carew et al. 74, copyright 1983 by the American Association for the Advancement of Science.


Figure 20.

Electrophysiological analysis of classical conditioning of siphon‐withdrawal reflex in Aplysia. A: experimental arrangement (A1) and training protocol (A2). Intracellular recordings were obtained in vitro from 2 different sensory neurons (SN) and siphon motoneuron (MN) to which they both projected. EPSP elicited in motoneuron by each sensory neuron was tested before and after training. Training consisted of trials in which train of spikes in 1 SN (paired SN) was followed by electric tail‐shock US; train of spikes in other SN (unpaired SN) was explicitly unpaired. Intertrial interval was 5 min. B: examples of EPSPs produced in siphon MN by paired SN and unpaired SN before (PRE) and 1 h after (POST) training protocol shown in A. Note that there is greater facilitation of EPSP elicited by paired SN than by unpaired SN. C: example of differential broadening of action potential in paired SN relative to an unpaired SN in the same preparation. Action potentials recorded before (PRE) and 3 h after (POST) 15 training trials using protocol shown in A. Experiment conducted in 50 mM TEA. D: removal of extracellular Ca2+ diminishes spike broadening produced by serotonin (5‐HT) paired with spike activity. Trains of action potentials in single SN were paired with brief (‐1 s) applications of serotonin alternately in normal seawater (10−2 M Ca2+) and in Ca2+‐free seawater. Action‐potential duration was measured in normal seawater before and after pairing in each solution. Note that after action potentials were paired with serotonin application in normal seawater, increase in action‐potential duration was significantly greater than after pairing in absence of Ca2+ influx (n = 8, P < 0.001).

A, B, and C from Hawkins et al. 188, copyright 1983 by the American Association for the Advancement of Science; D from Abrams 1.
References
 1. Abrams, T. W. Activity‐dependent presynaptic facilitation: an associative mechanism in Aplysia. Cell Mol. Neurobiol. 5: 123–145, 1985.
 2. Abrams, T. W., T. J. Carew, R. D. Hawkins, and E. R. Kandel. Aspects of the cellular mechanism of temporal specificity in conditioning in Aplysia: preliminary evidence for Ca2+ influx as a signal of activity. Soc. Neurosci. Abstr. 9: 168, 1983.
 3. Abrams, T. W., V. F. Castellucci, J. S. Camardo, E. R. Kandel, and P. E. Lloyd. Two endogenous neuropeptides modulate the gill and siphon withdrawal reflex in Aplysia by presynaptic facilitation involving cAMP‐dependent closure of a serotonin‐sensitive potassium channel. Proc. Natl. Acad. Sci. USA 81: 7956–7960, 1984.
 4. Abrams, T. W., L. Eliot, Y. Dudai, and E. R. Kandel. Activation of adenylate cyclase in Aplysia neural tissue by Ca2+/calmodulin, a candidate for an associative mechanism during conditioning. Soc. Neurosci. Abstr. 11: 797, 1985.
 5. Aceves‐Piña, E. O., R. Booker, J. S. Duerr, M. S. Livingstone, W. G. Quinn, R. F. Smith, P. P. Sziber, B. L. Tempel, and T. P. Tully. Learning and memory in Drosophila, studied with mutants. Cold Spring Harbor Symp. Quant. Biol. 48: 831–840, 1983.
 6. Adey, W. R. Electrophysiological patterns and cerebral impedance characteristics in orienting and discriminative behavior. Proc. Int. Congr. Physiol. Sci., 23rd, Tokyo, 1965, p. 324–339.
 7. Alkon, D. L. Associative training of Hermissenda. J. Gen. Physiol. 64: 70–84, 1974.
 8. Alkon, D. L. Voltage‐dependent calcium and potassium ion conductances: a contingency mechanism for an associative learning model. Science Wash. DC 205: 810–816, 1979.
 9. Alkon, D. L. Cellular analysis of a gastropod (Hermissenda crassicornis) model of associative learning. Biol. Bull. Woods Hole 159: 505–560, 1980.
 10. Alkon, D. L. Membrane depolarization accumulates during acquisition of an associative behavioral change. Science Wash. DC 210: 1375–1376, 1980.
 11. Alkon, D. L. Calcium‐mediated reduction of ionic currents: a biophysical memory trace. Science Wash. DC 226: 1037–1045, 1984.
 12. Alkon, D. L., J. Acosta‐Urquidi, J. Olds, G. Kuzma, and J. T. Neary. Protein kinase injection reduces voltage‐dependent potassium currents. Science Wash. DC 219: 303–306, 1983.
 13. Alkon, D. L., and J. Farley (editors). Primary Neural Substrates of Learning and Behavioral Change. Cambridge, UK: Cambridge Univ. Press, 1984.
 14. Alkon, D. L., J. Farley, M. Sakakibara, and B. Hay. Voltage‐dependent calcium and calcium‐activated potassium currents of a molluscan photoreceptor. Biophys. J. 46: 605–614, 1984.
 15. Alkon, D. L., and Y. Grossman. Long‐lasting depolarization and hyperpolarization in eye of Hermissenda. J. Neurophysiol. 41: 1328–1342, 1978.
 16. Alkon, D. L., I. Lederhendler, and J. J. Shoukimas. Primary changes of membrane currents during retention of associative learning. Science Wash. DC 215: 693–695, 1982.
 17. Alkon, D. L., and M. Sakakibara. Calcium activates and inactivates a photoreceptor soma potassium current. Biophys. J. 48: 983–995, 1985.
 18. Alkon, D. L., M. Sakakibara, R. Forman, J. Harrigan, J. Lederhendler, and J. Farley. Reduction of two voltage‐dependent K+ currents mediates retention of a learned association. Behav. Neural Biol. 44: 278–300, 1985.
 19. Alkon, D. L., J. J. Shoukimas, and E. Heldman. Calcium‐mediated decrease of a voltage‐dependent potassium current. Biophys. J. 40: 245–250, 1982.
 20. Aréchiga, H., B. Barrera‐Mera, and B. Fuentes‐Pardo. Habituation of mechanoreceptive interneurons in the crayfish. J. Neurobiol. 6: 131–144, 1975.
 21. Audesirk, T. E., J. E. Alexander, Jr., G. J. Audesirk, and C. M. Moyer. Rapid, nonaversive conditioning in a freshwater gastropod. I. Effects of age and motivation. Behav. Neural Biol. 36: 379–390, 1982.
 22. Auerbach, S., L. Grover, and J. Farley. HPLC and immunocytochemical analyses of serotonin in Hermissenda central nervous systems. Soc. Neurosci. Abstr. 11: 481, 1985.
 23. Bailey, C. H., and M. Chen. Morphological basis of longterm habituation and sensitization in Aplysia. Science Wash. DC 220: 91–93, 1983.
 24. Bailey, C. H., and M. Chen. Further studies on the morphological basis of long‐term habituation in Aplysia. Soc. Neurosci. Abstr. 10: 131, 1984.
 25. Bailey, C. H., and M. Chen. Morphological basis of shortterm habituation in Aplysia. Soc. Neurosci. Abstr. 11: 1110, 1985.
 26. Bailey, C. H., and M. Chen. Long‐term sensitization in Aplysia increases the total number of varicosities of single identified sensory neurons. Soc. Neurosci. Abstr. 12: 860, 1986.
 27. Bailey, C. H., R. D. Hawkins, and M. Chen. Uptake of [3H] serotohin in the abdominal ganglion of Aplysia californica: further studies on the morphological and biochemical basis of presynaptic facilitation. Brain Res 272: 71–81, 1983.
 28. Bailey, C. H., R. D. Hawkins, M. C. Chen, and E. R. Kandel. Interneurons involved in mediation and modulation of gill‐withdrawal reflex in Aplysia. IV. Morphological basis of presynaptic facilitation. J. Neurophysiol. 45: 340–360, 1981.
 29. Baraban, J. M., S. H. Snyder, and B. E. Alger. Protein kinase C regulates ionic conductance in hippocampal pyramidal neurons: electrophysiological effects of phorbol esters. Proc. Natl. Acad. Sci. USA 82: 2538–2542, 1985.
 30. Baranyi, A., and O. Fehér. Intracellular studies on cortical synaptic plasticity: conditioning effect of antidromic activation on test‐EPSPs. Exp. Brain Res. 41: 124–134, 1981.
 31. Baranyi, A., and O. Fehér. Long‐term facilitation of excitatory synaptic transmission in single motor cortical neurones of the cat produced by repetitive pairing of synaptic potentials and action potentials following intracellular stimulation. Neurosci. Lett. 23: 303–308, 1981.
 32. Baranyi, A., and O. Fehér. Synaptic facilitation requires paired activation of convergent pathways in the neocortex. Nature Lond. 290: 413–415, 1981.
 33. Barnes, C. A., and B. L. McNaughton. Spatial memory and hippocampal synaptic plasticity in senescent and middle‐aged rats. In: Psychobiology of Aging: Problems and Perspectives, edited by O. Stein. Amsterdam: Elsevier, 1980, p. 253–272.
 34. Barrionuevo, G., and T. H. Brown. Associative long‐term potentiation in hippocampal slices. Proc. Natl. Acad. Sci. USA 80: 7347–7351, 1983.
 35. Belardetti, F., C. Biondi, M. Brunelli, M. Fabri, and A. Trevisani. Heterosynaptic facilitation and behavioral sensitization are inhibited by lowering endogenous cAMP in Aplysia. Brain Res. 288: 95–104, 1983.
 36. Belardetti, F., S. Schacher, E. R. Kandel, and S. A. Siegelbaum. The growth cones of Aplysia sensory neurons: modulation by serotonin of action potential duration and single potassium channel currents. Proc. Natl. Acad. Sci. USA 83: 7094–7098, 1986.
 37. Berger, T. W. Long‐term potentiation of hippocampal synaptic transmission affects rate of behavioral learning. Science Wash. DC 224: 627–630, 1984.
 38. Berger, T. W., B. E. Alger, and R. F. Thompson. Neuronal substrates of classical conditioning in the hippocampus. Science Wash. DC 192: 483–485, 1976.
 39. Berger, T. W., G. A. Clark, and R. F. Thompson. Learning‐dependent neuronal responses recorded from limbic system brain structures during classical conditioning. Physiol. Psychol. 8: 155–167, 1980.
 40. Berger, T. W., R. I. Laham, and R. F. Thompson. Hippocampal unit‐behavior correlations during classical conditioning. Brain Res. 193: 229–248, 1980.
 41. Berger, T. W., and W. B. Orr. Hippocampectomy selectively disrupts discrimination reversal learning of the rabbit nictitating membrane response. Behav. Brain Res. 8: 49–68, 1983.
 42. Berger, T. W., P. C. Rinaldi, D. J. Weisz, and R. F. Thompson. Single‐unit analysis of different hippocampal cell types during classical conditioning of rabbit nictitating membrane response. J. Neurophysiol. 50: 1197–1219, 1983.
 43. Berger, T. W., and R. F. Thompson. Limbic system interrelations: functional division among hippocampal‐septal connections. Science Wash. DC 197: 587–589, 1977.
 44. Berger, T. W., and R. F. Thompson. Identification of pyramidal cells as the critical elements in hippocampal neuronal plasticity during learning. Proc. Natl. Acad. Sci. USA 75: 1572–1576, 1978.
 45. Berger, T. W., and R. F. Thompson. Neuronal plasticity in the limbic system during classical conditioning of the rabbit nictitating membrane response. I. The hippocampus. Brain Res. 145: 323–346, 1978.
 46. Berger, T. W., and R. F. Thompson. Neuronal plasticity in the limbic system during classical conditioning of the rabbit nictitating membrane response. II. Septum and mamillary bodies. Brain Res. 156: 293–314, 1978.
 47. Bernier, L., V. F. Castellucci, E. R. Kandel, and J. H. Schwartz. Facilitatory transmitter causes a selective and prolonged increase in adenosine 3′:5′‐monophosphate in sensory neurons mediating the gill and siphon withdrawal reflex in Aplysia. J. Neurosci. 2: 1682–1691, 1982.
 48. Berry, S. D., and R. F. Thompson. Medial septal lesions retard classical conditioning of the nictitating membrane response in rabbits. Science Wash. DC 205: 209–211, 1979.
 49. Berthier, N. E., and J. W. Moore. The nictitating membrane response: an electrophysiological study of the abducens nerve and nucleus and the accessory abducens nucleus in rabbit. Brain Res. 258: 201–210, 1983.
 50. Birt, D., and M. E. Olds. Auditory response enhancement during differential conditioning in behaving rats. In: Conditioning: Representation of Involved Neural Functions, edited by C. D. Woody. New York: Plenum, 1982, p. 483–501.
 51. Black‐Cleworth, P., C. D. Woody, and J. Niemann. A conditioned eye blink obtained by using electrical stimulation of the facial nerve as the unconditioned stimulus. Brain Res. 90: 45–56, 1975.
 52. Bliss, T. V. P., and A. R. Gardner‐Medwin. Long‐lasting potentiation of synaptic transmission in the dentate area of the unanaesthetized rabbit following stimulation of the per‐forant path. J. Physiol. Lond. 232: 357–374, 1973.
 53. Bliss, T. V. P., G. V. Goddard, and M. Riives. Reduction of long‐term potentiation in the dentate gyrus of the rat following selective depletion of monoamines. J. Physiol. Lond. 334: 475–491, 1983.
 54. Bliss, T. V. P., and T. Lomo. Long‐lasting potentiation of synaptic transmission in the dentate area of the anaesthetized rabbit following stimulation of the perforant path. J. Physiol. Lond. 232: 331–356, 1973.
 55. Boyle, M. B., M. Klein, S. J. Smith, and E. R. Kandel. Serotonin increases intracellular Ca2+ transients in voltage‐clamped sensory neurons of Aplysia californica. Proc. Natl. Acad. Sci. USA 81: 7642–7646, 1984.
 56. Breen, C. A., and H. L. Atwood. Octopamine—a neurohormone with presynaptic activity‐dependent effects at crayfish neuromuscular junctions. Nature Lond. 303: 716–718, 1983.
 57. Brons, J. F., and C. D. Woody. Long‐term changes in excitability of cortical neurons after Pavlovian conditioning and extinction. J. Neurophysiol. 44: 605–615, 1980.
 58. Brons, J. F., C. D. Woody, and N. Allon. Changes in excitability to weak‐intensity extracellular electrical stimulation of units of pericruciate cortex in cats. J. Neurophysiol. 47: 377–388, 1982.
 59. Brown, L. T. Corticorubral projections in the rat. J. Comp. Neurol. 154: 149–168, 1974.
 60. Broyles, J. L., and D. H. Cohen. An input from locus coeruleus is necessary for discharge modification of avian lateral geniculate neurons during visual learning. Soc. Neurosci. Abstr. 11: 1109, 1985.
 61. Brunelli, M., V. Castellucci, and E. R. Kandel. Synaptic facilitation and behavioral sensitization in Aplysia: possible role of serotonin and cAMP. Science Wash. DC 194: 1178–1181, 1976.
 62. Bruner, J., and D. Kennedy. Habituation: occurrence at a neuromuscular junction. Science Wash. DC 169: 92–94, 1970.
 63. Bryan, J. S., and F. B. Krasne. Protection from habituation of the crayfish lateral giant fibre escape response. J. Physiol. Lond. 271: 351–368, 1977.
 64. Bryan, J. S., and F. B. Krasne. Presynaptic inhibition: the mechanism of protection from habituation of the crayfish lateral giant fibre escape response. J. Physiol. Lond. 271: 369–390, 1977.
 65. Burke, W. Neuronal models for conditioned reflexes. Nature Lond. 210: 269–271, 1966.
 66. Byers, D., R. L. Davis, and J. A. Kiger. Defect in cyclic AMP phosphodiesterase due to the dunce mutation of learning in Drosophila melanogaster. Nature Lond. 289: 79–81, 1981.
 67. Byrne, J. H. Comparative aspects of neural circuits for inking behavior and gill withdrawal in Aplysia californica. J. Neurophysiol. 45: 98–106, 1981.
 68. Byrne, J. H., V. F. Castellucci, T. J. Carew, and E. R. Kandel. Stimulus‐response relations and stability of mechan‐oreceptor and motor neurons mediating defensive gill‐withdrawal reflex in Aplysia. J. Neurophysiol. 41: 402–417, 1978.
 69. Byrne, J., V. Castellucci, and E. R. Kandel. Receptive fields and response properties of mechanoreceptor neurons innervating siphon skin and mantle shelf in Aplysia. J. Neurophysiol. 37: 1041–1064, 1974.
 70. Callec, J. J., J. C. Guillet, Y. Pichon, and J. Boistel. Further studies on synaptic transmission in insects. II. Relations between sensory information and its synaptic integration at the level of a single giant axon in the cockroach. J. Exp. Biol. 55: 123–149, 1971.
 71. Carew, T. J., V. F. Castellucci, and E. R. Kandel. An analysis of dishabituation and sensitization of the gill‐withdrawal reflex in Aplysia. Int. J. Neurosci. 2: 79–98, 1971.
 72. Carew, T. J., V. F. Castellucci, and E. R. Kandel. Sensitization in Aplysia: restoration of transmission in synapses inactivated by long‐term habituation. Science Wash. DC 205: 417–419, 1979.
 73. Carew, T. J., R. D. Hawkins, T. W. Abrams, and E. R. Kandel. A test of Hebb's postulate at identified synapses which mediate classical conditioning in Aplysia. J. Neurosci. 4: 1217–1224, 1984.
 74. Carew, T. J., R. D. Hawkins, and E. R. Kandel. Differential classical conditioning of a defensive withdrawal reflex in Aplysia californica. Science Wash. DC 219: 397–400, 1983.
 75. Carew, T. J., H. M. Pinsker, and E. R. Kandel. Long‐term habituation of a defensive withdrawal reflex in Aplysia. Science Wash. DC 175: 451–454, 1972.
 76. Carew, T. J., E. T. Walters, and E. R. Kandel. Associative learning in Aplysia: cellular correlates supporting a conditioned fear hypothesis. Science Wash. DC 211: 501–504, 1981.
 77. Carew, T. J., E. T. Walters, and E. R. Kandel. Classical conditioning in a simple withdrawal reflex in Aplysia californica. J. Neurosci. 1: 1426–1437, 1981.
 78. Castellucci, V. F., L. Bernier, J. H. Schwartz, and E. R. Kandel. Persistent activation of adenylate cyclase underlies the time course of short‐term sensitization in Aplysia. Soc. Neurosci. Abstr. 9: 169, 1983.
 79. Castellucci, V. F., T. J. Carew, and E. R. Kandel. Cellular analysis of long‐term habituation of the gill‐withdrawal reflex of Aplysia californica. Science Wash. DC 202: 1306–1308, 1978.
 80. Castellucci, V. F., and E. R. Kandel. A quantal analysis of the synaptic depression underlying habituation of the gill‐withdrawal reflex in Aplysia. Proc. Natl. Acad. Sci. USA 71: 5004–5008, 1974.
 81. Castellucci, V., and E. R. Kandel. Presynaptic facilitation as a mechanism for behavioral sensitization in Aplysia. Science Wash. DC 194: 1176–1178, 1976.
 82. Castellucci, V. F., E. R. Kandel, J. H. Schwartz, F. D. Wilson, A. C. Nairn, and P. Greengard. Intracellular injection of the catalytic subunit of cyclic AMP‐dependent protein kinase simulates facilitation of transmitter release underlying behavioral sensitization in Aplysia. Proc. Natl. Acad. Sci. USA 77: 7492–7496, 1980.
 83. Castellucci, V. F., A. Nairn, P. Greengard, J. H. Schwartz, and E. R. Kandel. Inhibitor of adenosine 3′:5′‐ monophosphate‐dependent protein kinase blocks presynaptic facilitation in Aplysia. J. Neurosci. 2: 1673–1681, 1982.
 84. Castellucci, V., H. Pinsker, I. Kupfermann, and E. R. Kandel. Neuronal mechanisms of habituation and dishabituation of the gill‐withdrawal reflex in Aplysia. Science Wash. DC 167: 1745–1748, 1970.
 85. Cedar, H., E. R. Kandel, and J. H. Schwartz. Cyclic adenosine monophosphate in the nervous system of Aplysia californica. I. Increased synthesis in response to synaptic stimulation. J. Gen. Physiol. 60: 558–569, 1972.
 86. Cedar, H., and J. H. Schwartz. Cyclic adenosine monophosphate in the nervous system of Aplysia californica. II. Effect of serotonin and dopamine. J. Gen. Physiol. 60: 570–587, 1972.
 87. Cegavske, C. F., M. M. Patterson, and R. F. Thompson. Neuronal unit activity in the abducens nucleus during classical conditioning of the nictitating membrane response in the rabbit (Oryctolagus cuniculus). J. Comp. Physiol. Psychol. 93: 595–609, 1979.
 88. Clark, G. A. A cellular mechanism for the temporal specificity of classical conditioning of the siphon‐withdrawal response in Aplysia. Soc. Neurosci. Abstr. 10: 268, 1984.
 89. Clark, G. A. A system for the study of synapse specificity in long‐term memory in Aplysia. Soc. Neurosci. Abstr. 12: 1338, 1986.
 90. Clark, G. A., and E. R. Kandel. Branch‐specific heterosynaptic facilitation in Aplysia siphon sensory cells. Proc. Natl. Acad. Sci. USA 81: 2577–2581, 1984.
 91. Clark, G. A., D. A. McCormick, D. G. Lavond, and R. F. Thompson. Effects of lesions of cerebellar nuclei on conditioned behavioral and hippocampal neuronal responses. Brain Res. 291: 125–136, 1984.
 92. Coates, S. R., and R. F. Thompson. Comparing neural plasticity in the hippocampus during classical conditioning of the rabbit nictitating membrane response to light and tone. Soc. Neurosci. Abstr. 4: 806, 1978.
 93. Cohen, D. H. Development of a vertebrate experimental model for cellular neurophysiological studies of learning. Cond. Reflex 4: 61–80, 1969.
 94. Cohen, D. H. The neural pathways and informational flow mediating a conditioned autonomic response. In: Limbic and Autonomic Nervous System Research, edited by L. V. DiCara. New York: Plenum, 1974, p. 223–275.
 95. Cohen, D. H. The functional neuroanatomy of a conditioned response. In: Neural Mechanisms of Goal‐Directed Behavior and Learning, edited by R. F. Thompson, L. H. Hicks, and V. B. Shvyrkov. New York: Academic, 1980, p. 283–302.
 96. Cohen, D. H. Identification of vertebrate neurons modified during learning: analysis of sensory pathways. In: Primary Neural Substrates of Learning and Behavioral Change, edited by D. L. Alkon and J. Farley. Cambridge, UK: Cambridge Univ. Press, 1984, p. 129–154.
 97. Cohen, D. H. Some organizational principles of a vertebrate conditioning pathway: is memory a distributed property? In: Memory Systems of the Brain, edited by N. M. Weinberger, J. L. McGaugh, and G. Lynch. New York: Guilford, 1985, p. 27–48.
 98. Cohen, D. H., and R. G. Durkovic. Cardiac and respiratory conditioning, differentiation, and extinction in the pigeon. J. Exp. Anal. Behav. 9: 681–688, 1966.
 99. Cohen, D. H., C. M. Gibbs, J. Siegelman, P. Gamlin, and J. Broyles. Is locus coeruleus involved in plasticity of lateral geniculate neurons during learning? Soc. Neurosci. Abstr. 8: 666, 1982.
 100. Cohen, D. H., and R. L. MacDonald. Some variables affecting orienting and conditioned heart‐rate responses in the pigeon. J. Comp. Physiol. Psychol. 74: 123–133, 1971.
 101. Colwill, R. M. Context conditioning in Aplysia californica. Soc. Neurosci. Abstr. 11: 796, 1985.
 102. Connor, J., and D. L. Alkon. Light‐ and voltage‐dependent increases of calcium ion concentration in molluscan photoreceptors. J. Neurophysiol. 51: 745–752, 1984.
 103. Cook, D. G., and T. J. Carew. Operant conditioning of head waving in Aplysia. Proc. Natl. Acad. Sci. USA 83: 1120–1124, 1986.
 104. Crick, F. Memory and molecular turnover (Abstract). Nature Lond. 312: 101, 1984.
 105. Crow, T. Conditioned modification of locomotion in Hermissenda crassicornis: analysis of time‐dependent associative and nonassociative components. J. Neurosci. 3: 2621–2628, 1983.
 106. Crow, T. Cellular mechanisms of associative learning in Hermissenda: contribution of light‐activated conductances. Soc. Neurosci. Abstr. 11: 794, 1985.
 107. Crow, T. Conditioned modification of phototactic behavior in Hermissenda. II. Differential adaptation of B‐photoreceptors. J. Neurosci. 5: 215–223, 1985.
 108. Crow, T. J., and D. L. Alkon. Retention of an associative behavioral change in Hermissenda. Science Wash. DC 201: 1239–1241, 1978.
 109. Crow, T. J., and D. L. Alkon. Associative behavioral modification in Hermissenda: cellular correlates. Science Wash. DC 209: 412–414, 1980.
 110. Crow, T., and N. Offenbach. Modification of the initiation of locomotion in Hermissenda: behavioral analysis. Brain Res. 271: 301–310, 1983.
 111. Dale, N., E. R. Kandel, and S. Schacher. 5‐HT induces long‐lasting excitability changes in cultured Aplysia neurons. Soc. Neurosci. Abstr. 12: 1339, 1986.
 112. Davis, H. P., and L. R. Squire. Protein synthesis and memory: a review. Psychol. Bull. 96: 518–559, 1984.
 113. Davis, W. J., and R. Gillette. Neural correlate of behavioral plasticity in command neurons of Pleurobranchaea. Science Wash. DC 199: 801–804, 1978.
 114. Davis, W. J., R. Gillette, M. P. Kovac, R. P. Croll, and E. M. Matera. Organization of synaptic inputs to paracerebral feeding command interneurons of Pleurobranchaea californica. III. Modifications induced by experience. J. Neurophysiol. 49: 1557–1572, 1983.
 115. Davis, W. J., J. Villet, D. Lee, M. Rigler, R. Gillette, and E. Prince. Selective and differential avoidance learning in the feeding and withdrawal behavior of Pleurobranchaea californica. J. Comp. Physiol. 138: 157–165, 1980.
 116. Del Castillo, J., and B. Katz. Statistical factors involved in neuromuscular facilitation and depression. J. Physiol. Lond. 124: 574–585, 1954.
 117. Desmond, J. E., N. E. Berthier, and J. W. Moore. Brainstem elements essential for the classically conditioned nictitating membrane response of rabbit. Soc. Neurosci. Abstr. 7: 650, 1981.
 118. Desmond, J. E., and J. W. Moore. A brainstem region essential for the classically conditioned but not unconditioned nictitating membrane response. Physiol. Behav. 28: 1029–1033, 1982.
 119. Deutsch, J. A. (editor). The Physiological Basis of Memory. New York: Academic, 1983.
 120. Diamond, M., and N. M. Weinberger. Physiological plasticity of single neurons in auditory cortex of the cat during acquisition of the pupillary conditioned response. II. Secondary field (AII). Behav. Neurosci. 98: 189–210, 1984.
 121. Diamond, D. M., and N. M. Weinberger. Classical conditioning rapidly induces specific changes in frequency receptive fields of single neurons in secondary and ventral ectosylvian auditory cortical fields. Brain Res. 372: 357–360, 1986.
 122. Disterhoft, J. F., D. A. Coulter, and D. L. Alkon. Conditioning‐specific membrane changes of rabbit hippocampal neurons measured in vitro. Proc. Natl. Acad. Sci. USA 83: 2733–2737, 1986.
 123. Disterhoft, J. F., H. H. Kwan, and W. O. Lo. Nictitating membrane conditioning to tone in the immobilized albino rabbit. Brain Res. 137: 127–143, 1977.
 124. Disterhoft, J. F., and J. Olds. Differential development of conditioned unit changes in thalamus and cortex of rat. J. Neurophysiol. 35: 665–679, 1972.
 125. Disterhoft, J. F., and D. K. Stuart. Trial sequence of changed unit activity in auditory system of alert rat during conditioned response acquisition and extinction. J. Neurophysiol. 39: 266–281, 1976.
 126. Disterhoft, J. F., and D. K. Stuart. Differentiated short latency response increases after conditioning in inferior collic‐ulus neurons of alert rat. Brain Res. 130: 315–333, 1977.
 127. Donegan, N. H., R. W. Lowery, and R. F. Thompson. Effects of lesioning cerebellar nuclei on conditioned leg‐flexion responses. Soc. Neurosci. Abstr. 9: 331, 1983.
 128. Downey, P., and B. Jahan‐Parwar. Cooling as a reinforcing stimulus in Aplysia. Am. Zool. 12: 507–512, 1972.
 129. Dudai, Y., Y.‐N. Jan, D. Byers, W. G. Quinn, and S. Benzer. Dunce, a mutant of Drosophila deficient in learning. Proc. Natl. Acad. Sci. USA 73: 1684–1688, 1976.
 130. Dudai, Y., A. Uzzan, and S. Zvi. Abnormal activity of adenylate cyclase in the Drosophila memory mutant rutabaga. Neurosci. Lett. 42: 207–212, 1983.
 131. Dudai, Y., and S. Zvi. Adenylate cyclase in the Drosophila memory mutant rutabaga displays an altered Ca2+ sensitivity. Neurosci. Lett. 47: 119–124, 1984.
 132. Duerr, J. S., and W. G. Quinn. Three Drosophila mutants that block associative learning also affect habituation and sensitization. Proc. Natl. Acad. Sci. USA 79: 3646–3650, 1982.
 133. Dufossé, M., M. Ito, P. J. Jastreboff, and Y. Miyashita. A neuronal correlate in rabbit's cerebellum to adaptive modification of the vestibulo‐ocular reflex. Brain Res. 150: 611–616, 1978.
 134. Eberly, L. B., and H. M. Pinsker. Neuroethological studies of reflex plasticity in intact Aplysia. Behav. Neurosci. 98: 609–630, 1984.
 135. Eisenstein, E. M., and M. J. Cohen. Learning in an isolated insect ganglion. Anim. Behav. 13: 104–108, 1965.
 136. Eliot, L. S., Y. Dudai, E. R. Kandel, and T. W. Abrams. Activation of adenylate cyclase in Aplysia by Ca2+/calmodulin: a possible molecular site of stimulus convergence in associative conditioning. Soc. Neurosci. Abstr. 12: 400, 1986.
 137. Engel, J., Jr., and C. D. Woody. Effects of character and significance of stimulus on unit activity at coronal‐pericruciate cortex of cat during performance of conditioned motor response. J. Neurophysiol. 35: 220–229, 1972.
 138. Engel, J., Jr., and C. D. Woody. Changes in unit activity and thresholds to electrical micro‐stimulation at coronal‐pericruciate cortex of cat with classical conditioning of different facial movements. J. Neurophysiol. 35: 230–241, 1972.
 139. Farel, P. B., D. L. Glanzman, and R. F. Thompson. Habituation of a monosynaptic response in vertebrate central nervous system: lateral column‐motoneuron pathway in isolated frog spinal cord. J. Neurophysiol. 36: 1117–1130, 1973.
 140. Farel, P. B., and R. F. Thompson. Habituation of a monosynaptic response in frog spinal cord: evidence for a presynaptic mechanism. J. Neurophysiol. 39: 661–666, 1976.
 141. Farley, J., and D. L. Alkon. Associative neural and behavioral change in Hermissenda: consequences of nervous system orientation for light and pairing specificity. J. Neurophysiol. 48: 785–807, 1982.
 142. Farley, J., and S. Auerbach. Protein kinase C activation induces conductance changes in Hermissenda photoreceptors like those seen in associative learning. Nature Lond. 319: 220–223, 1986.
 143. Farley, J., W. G. Richards, L. J. Ling, E. Liman, and D. L. Alkon. Membrane changes in a single photoreceptor cause associative learning in Hermissenda. Science Wash. DC 221: 1201–1203, 1983.
 144. Forman, R. R. Leg position learning by an insect. I. A heat avoidance learning paradigm. J. Neurobiol. 15: 127–140, 1984.
 145. Foy, M. R., J. E. Steinmetz, and R. F. Thompson. Single unit analysis of cerebellum during classically conditioned eyelid response. Soc. Neurosci. Abstr. 10: 122, 1984.
 146. Frazier, W. T., E. R. Kandel, I. Kupfermann, R. Waziri, and R. E. Coggeshall. Morphological and functional properties of identified neurons in the abdominal ganglion of Aplysia californica. J. Neurophysiol. 30: 1288–1351, 1967.
 147. Frost, W. N., V. F. Castellucci, R. D. Hawkins, and E. R. Kandel. Monosynaptic connections made by the sensory neurons of the gill‐ and siphon‐withdrawal reflex in Aplysia participate in the storage of long‐term memory for sensitization. Proc. Natl. Acad. Sci. USA 82: 8266–8269, 1985.
 148. Frost, W. N., G. A. Clark, and E. R. Kandel. Changes in cellular excitability in a new class of siphon motor neurons during sensitization in Aplysia. Soc. Neurosci. Abstr. 11: 643, 1985.
 149. Frost, W. N., and E. R. Kandel. Sensitizing stimuli reduce the effectiveness of the L30 inhibitory interneurons in the siphon withdrawal reflex circuit of Aplysia. Soc. Neurosci. Abstr. 10: 510, 1984.
 150. Fujito, Y., N. Tsukahara, Y. Oda, and M. Yoshida. Formation of functional synapses in the adult feline red nucleus from the cerebrum following cross‐innervation of forelimb flexor and extensor nerves. II. Analysis of newly‐appeared postsynaptic potentials. Exp. Brain Res. 45: 13–18, 1982.
 151. Gabriel, M. Short‐latency discriminative unit response: engram or bias? Physiol. Psychol. 4: 275–280, 1976.
 152. Gabriel, M., J. D. Miller, and S. E. Saltwick. Unit activity in cingulate cortex and anteroventral thalamus of the rabbit during differential conditioning and reversal. J. Comp. Physiol. Psychol. 91: 423–433, 1977.
 153. Gabriel, M., E. Orona, K. Foster, and R. W. Lambert. Mechanism and generality of stimulus significance coding in a mammalian model system. In: Conditioning: Representation of Involved Neural Functions, edited by C. D. Woody. New York: Plenum, 1982, p. 535–565.
 154. Gabriel, M., S. E. Saltwick, and J. D. Miller. Conditioning and reversal of short‐latency multiple‐unit responses in the rabbit medial geniculate nucleus. Science Wash. DC 189: 1108–1109, 1975.
 155. Garcia, J., B. K. McGowan, and K. F. Green. Biological constraints on conditioning. In: Classical Conditioning II: Current Research and Theory, edited by A. H. Black and W. F. Prokasy. New York: Appleton‐Century‐Crofts, 1972, p. 3–27.
 156. Gelperin, A. Rapid food‐aversion learning by a terrestrial mollusk. Science Wash. DC 189: 567–570, 1975.
 157. Gibbs, C. M., J. L. Broyles, and D. H. Cohen. Further studies of the involvement of locus coeruleus in plasticity of avian lateral geniculate neurons during learning. Soc. Neurosci. Abstr. 9: 641, 1983.
 158. Gibbs, C. M., D. H. Cohen, and J. L. Broyles. Modification of the discharge of lateral geniculate neurons during visual learning. J. Neurosci. 6: 627–636, 1986.
 159. Gillette, R., M. P. Kovac, and W. J. Davis. Command neurons in Pleurobranchaea receive synaptic feedback from the motor network they excite. Science Wash. DC 199: 798–801, 1978.
 160. Gingrich, K. J., and J. H. Byrne. Simulation of synaptic depression, posttetanic potentiation, and presynaptic facilitation of synaptic potentials from sensory neurons mediating gill‐withdrawal reflex in Aplysia. J. Neurophysiol. 53: 652–669, 1985.
 161. Glantz, R. M. The visually evoked defense reflex of the crayfish: habituation, facilitation, and the influence of picrotoxin. J. Neurobiol. 5: 263–280, 1974.
 162. Glantz, R. M. Habituation of the motion detectors of the crayfish optic nerve: their relationship to the visually evoked defense reflex. J. Neurobiol. 5: 489–501, 1974.
 163. Glantz, R. M. Visual input and motor output of command interneurons in the crayfish defense reflex pathway. In: Identified Neurons and Behavior of Arthropods, edited by G. Hoyle. New York: Plenum, 1977, p. 259–274.
 164. Glanzman, D. L., S. Mackey, and E. R. Kandel. Depletion of serotonin in Aplysia interferes with facilitation produced by sensitizing stimuli. Soc. Neurosci. Abstr. 12: 1339, 1986.
 165. Glanzman, D. L., and R. F. Thompson. Evidence against conduction failure as the mechanism underlying monosynaptic habituation in frog spinal cord. Brain Res. 174: 329–332, 1979.
 166. Glanzman, D. L., and R. F. Thompson. Alterations in spontaneous miniature potential activity during habituation of a vertebrate monosynaptic pathway. Brain Res. 189: 377–390, 1980.
 167. Gluck, M. A., and R. F. Thompson. Modeling the neural substrates of associative learning and memory: a computational approach. Psychol. Rev. In press.
 168. Goelet, P., V. F. Castellucci, S. Schacher, and E. R. Kandel. The long and the short of long‐term memory: a molecular framework. Nature Lond. 322: 419–422, 1986.
 169. Goh, Y., and D. L. Alkon. Sensory, interneuronal, and motor interactions within Hermissenda visual pathway. J. Neurophysiol. 52: 156–169, 1984.
 170. Goh, Y., I. Lederhendler, and D. L. Alkon. Input and output changes of an identified neural pathway are correlated with association learning in Hermissenda. J. Neurosci. 5: 536–543, 1985.
 171. Gold, M. R., and D. H. Cohen. Modification of the discharge of vagal cardiac neurons during learned heart rate change. Science Wash. DC 214: 345–347, 1981.
 172. Gold, M. R., and D. H. Cohen. The discharge characteristics of vagal cardiac neurons during classically conditioned heart rate change. J. Neurosci. 4: 2963–2971, 1984.
 173. Goldberg, J. I., and K. Lukowiak. Transfer of habituation in Aplysia: contributions of heterosynaptic pathways in habituation of the gill‐withdrawal reflex. J. Neurobiol. 15: 395–411, 1984.
 174. Goldstein, R., H. B. Kistler, H. W. M. Steinbusch, and J. H. Schwartz. Distribution of serotonin‐immunoreactivity in juvenile Aplysia. Neuroscience 11: 535–547, 1984.
 175. Gormezano, I. Investigations of defense and reward conditioning in the rabbit. In: Classical Conditioning II: Current Research and Theory, edited by A. H. Black and W. F. Prokasy. New York: Appleton‐Century‐Crofts, 1972, p. 151–181.
 176. Gormezano, I., W. F. Prokasy, and R. F. Thompson (editors). Classical Conditioning III: Behavioral, Neurophysiological and Neurochemical Studies in the Rabbit. Hillsdale, NJ: Erlbaum, in press.
 177. Gormezano, I., N. Schneiderman, E. B. Deaux, and I. Fuentes. Nictitating membrane: classical conditioning and extinction in the albino rabbit. Science Wash. DC 138: 33–34, 1962.
 178. Greenberg, S. M., H. Bayley, V. F. Castellucci, and J. H. Schwartz. Selective loss of the regulatory subunit of the cAMP‐dependent protein kinase in Aplysia after sensitizing treatments. Soc. Neurosci. Abstr. 12: 1339, 1986.
 179. Grether, W. F. Pseudo‐conditioning without paired stimulation encountered in attempted backward conditioning. J. Comp. Psychol. 25: 91–96, 1938.
 180. Grover, L., and J. Farley. Temporal order sensitivity of associative learning in Hermissenda. Soc. Neurosci. Abstr. 9: 915, 1983.
 181. Groves, P. M., and R. F. Thompson. Habituation: a dualprocess theory. Psychol. Rev. 77: 419–450, 1970.
 182. Gustafsson, B., and H. Wigstrom. Hippocampal longlasting potentiation produced by pairing single volleys and brief conditioning tetani evoked by separate afferents. J. Neurosci. 6: 1575–1582, 1986.
 183. Haley, D. A., D. G. Lavond, and R. F. Thompson. Effects of contralateral red nuclear lesions in retention of the classically conditioned nictitating membrane/eyelid response. Soc. Neurosci. Abstr. 9: 643, 1983.
 184. Hawkins, R. D. Identified facilitating neurons are excited by cutaneous stimuli used in sensitization and classical conditioning of Aplysia. Soc. Neurosci. Abstr. 7: 354, 1981.
 185. Hawkins, R. D. Interneurons involved in mediation and modulation of gill‐withdrawal reflex in Aplysia. III. Identified facilitating neurons increase Ca2+ current in sensory neurons. J. Neurophysiol. 45: 327–339, 1981.
 186. Hawkins, R. D. Localization of potential serotonergic facilitator neurons by glyoxylic acid histofluorescence and retrograde fluorescent labeling in Aplysia. Soc. Neurosci. Abstr. 12: 1339, 1986.
 187. Hawkins, R. D., and T. W. Abrams. Evidence that activity‐dependent facilitation underlying classical conditioning in Aplysia involves modulation of the same ionic current as normal presynaptic facilitation. Soc. Neurosci. Abstr. 10: 268, 1984.
 188. Hawkins, R. D., T. W. Abrams, T. J. Carew, and E. R. Kandel. A cellular mechanism of classical conditioning in Aplysia: activity‐dependent amplification of presynaptic facilitation. Science Wash. DC 219: 400–405, 1983.
 189. Hawkins, R. D., and J. Bruner. Activity of excitor and inhibitor claw motor neurones during habituation and dishabituation of the crayfish defence response. J. Exp. Biol. 91: 145–164, 1981.
 190. Hawkins, R. D., T. J. Carew, and E. R. Kandel. Effects of interstimulus interval and contingency on classical conditioning of the Aplysia siphon withdrawal reflex. J. Neurosci. 6: 1695–1701, 1986.
 191. Hawkins, R. D., V. F. Castellucci, and E. R. Kandel. Interneurons involved in mediation and modulation of gillwithdrawal reflex in Aplysia. I. Identification and characterization. J. Neurophysiol. 45: 304–314, 1981.
 192. Hawkins, R. D., V. F. Castellucci, and E. R. Kandel. Interneurons involved in mediation and modulation of gill‐withdrawal reflex in Aplysia. II. Identified neurons produce heterosynaptic facilitation contributing to behavioral sensitization. J. Neurophysiol. 45: 315–326, 1981.
 193. Hawkins, R. D., G. A. Clark, and E. R. Kandel. Operant conditioning and differential classical conditioning of gill withdrawal in Aplysia. Soc. Neurosci. Abstr. 11: 796, 1985.
 194. Hawkins, R. D., and E. R. Kandel. Is there a cell biological alphabet for simple forms of learning? Psychol. Rev. 91: 375–391, 1984.
 195. Hebb, D. O. The Organization of Behavior: A Neuropsychological Theory. New York: Wiley, 1949.
 196. Hilgard, E. R., and G. H. Bower. Theories of Learning. Englewood Cliffs, NJ: Prentice‐Hall, 1975.
 197. Hochner, B., O. Braha, M. Klein, and E. R. Kandel. Distinct processes in presynaptic facilitation contribute to sensitization and dishabituation in Aplysia: possible involvement of C kinase in dishabituation. Soc. Neurosci. Abstr. 12: 1340, 1986.
 198. Hochner, B., M. Klein, S. Schacher, and E. R. Kandel. A novel component in the cellular mechanism of presynaptic facilitation contributes to behavorial dishabituation in Aplysia. Proc. Natl. Acad. Sci. USA. 83: 8794–8798, 1986.
 199. Hochner, B., S. Schacher, M. Klein, and E. R. Kandel. Presynaptic facilitation in Aplysia sensory neurons: a process independent of K+ current modulation becomes important when transmitter release is depressed. Soc. Neurosci. Abstr. 11: 29, 1985.
 200. Hoehler, F. K., and R. F. Thompson. Effect of the interstimulus (CS‐UCS) interval on hippocampal unit activity during classical conditioning of the nictitating membrane response of the rabbit (Oryctolagus cuniculus). J. Comp. Physiol. Psychol. 94: 201–215, 1980.
 201. Hoffmann, P. Über die doppelte Innervation der Krebsmuskeln. Zugleich ein Beitrag zur Kenntnis nervoser Hemmungen. Z. Biol. 63: 411–442, 1914.
 202. Hopkins, W. F., and D. Johnston. Frequency‐dependent noradrenergic modulation of long‐term potentiation in the hippocampus. Science Wash. DC 226: 350–352, 1984.
 203. Horn, R., and J. J. Miller. A prolonged, voltage‐dependent calcium permeability revealed by tetraethylammonium in the soma and axon of Aplysia giant neuron. J. Neurobiol. 8: 399–415, 1977.
 204. Horridge, G. A. Learning of leg position by headless insects. Nature Lond. 193: 697–698, 1962.
 205. Horridge, G. A. Learning of leg position by the ventral nerve cord in headless insects. Proc. R. Soc. Lond. B Biol. Sci. 157: 33–52, 1962.
 206. Hoyle, G. Neurophysiological studies on “learning” in headless insects. In: The Physiology of the Insect Central Nervous System, edited by J. E. Treherne and J. W. L. Beament. New York: Academic, 1965, p. 203–232.
 207. Hoyle, G. Instrumental conditioning of the leg lift in the locust. Neurosci. Res. Program Bull. 17: 577–586, 1979.
 208. Hoyle, G. Learning, using natural reinforcements, in insect preparations that permit cellular neuronal analysis. J. Neurobiol. 11: 323–354, 1980.
 209. Hoyle, G. Cellular basis of operant‐conditioning of leg position. In: Conditioning: Representation of Involved Neural Functions, edited by C. D. Woody. New York: Plenum, 1982, p. 197–211.
 210. Humphrey, G. The Nature of Learning in its Relation to the Living System. New York: Harcourt Brace, 1933.
 211. Ingram, D. A., and E. T. Walters. Differential classical conditioning of tail and siphon withdrawal in Aplysia. Soc. Neurosci. Abstr. 10: 270, 1984.
 212. Ito, M. Neural design of the cerebellar motor control system. Brain Res. 40: 81–84, 1972.
 213. Ito, M. Cerebellar control of the vestibular‐ocular reflex— around the flocculus hypothesis. Annu. Rev. Neurosci. 5: 275–296, 1982.
 214. Ito, M., M. Sakurai, and P. Tongroach. Climbing fibre induced depression of both mossy fibre responsiveness and glutamate sensitivity of cerebellar Purkinje cells. J. Physiol. Lond. 324: 113–134, 1982.
 215. Jacklet, J. W., and J. Rine. Facilitation at neuromuscular junctions: contribution to habituation and dishabituation of the Aplysia gill withdrawal reflex. Proc. Natl. Acad. Sci. USA 74: 1267–1271, 1977.
 216. James, W. The Principles of Psychology. New York: Holt, 1890.
 217. Jerne, N. K. Antibodies and learning: selection versus instruction. In: The Neurosciences: A Study Program, edited by G. C. Quarton, T. Melnechuk, and F. O. Schmitt. New York: Rockefeller Univ. Press, 1967, p. 200–205.
 218. John, E. R. Mechanisms of Memory. New York: Academic, 1967.
 219. John, E. R., Y. Tang, A. B. Brill, R. Young, and K. Ono. Double‐labeled metabolic maps of memory. Science Wash. DC 233: 1167–1175, 1986.
 220. Kandel, E. R. A Cell‐Biological Approach to Learning. Bethesda, MD: Soc. Neurosci., 1978.
 221. Kandel, E. R., T. Abrams, L. Bernier, T. J. Carew, R. D. Hawkins, and J. H. Schwartz. Classical conditioning and sensitization share aspects of the same molecular cascade in Aplysia. Cold Spring Harbor Symp. Quant. Biol. 48: 821–830, 1983.
 222. Kandel, E. R., and J. H. Schwartz. Molecular biology of learning: modulation of transmitter release. Science Wash. DC 218: 433–443, 1982.
 223. Kandel, E. R., and W. A. Spencer. Cellular neurophysiological approaches in the study of learning. Physiol. Rev. 48: 65–134, 1968.
 224. Katz, B., and R. Miledi. Tetrodotoxin‐resistant electrical activity in presynaptic terminals. J. Physiol. Lond. 203: 459–487, 1969.
 225. Kelso, S. R., and T. H. Brown. Differential conditioning of associative synaptic enhancement in hippocampal brain slices. Science Wash. DC 232: 85–87, 1986.
 226. Kelso, S. R., A. H. Ganong, and T. H. Brown. Hebbian synapses in hippocampus. Proc. Natl. Acad. Sci. USA 83: 5326–5330, 1986.
 227. Kennedy, D., R. L. Calabrese, and J. J. Wine. Presynaptic inhibition: primary afferent depolarization in crayfish neurons. Science Wash. DC 186: 451–454, 1974.
 228. Kettner, R. E., and R. F. Thompson. Auditory signal detection and decision processes in the nervous system. J. Comp. Physiol. Psychol. 96: 328–331, 1982.
 229. Kim, E. H.‐J., C. D. Woody, and N. E. Berthier. Rapid acquisition of conditioned eye blink responses in cats following pairing of an auditory CS with glabella tap US and hypothalamic stimulation. J. Neurophysiol. 49: 767–779, 1983.
 230. Kimble, G. A. Hilgard and Marquis' Conditioning and Learning. New York: Appleton‐Century‐Crofts, 1961.
 231. King, J. S., R. M. Dom, J. B. Conner, and G. F. Martin. An experimental light and electron microscopic study of cerebellorubral projections in the oppossum, Didelphis marsupialis virginiana. Brain. Res. 52: 61–78, 1973.
 232. King, J. S., G. F. Martin, and J. B. Conner. A light and electron microscopic study of corticorubral projections in the oppossum, Didelphis marsupialis virginiana. Brain Res. 38: 251–265, 1972.
 233. Kirk, M. D. Presynaptic inhibition in the crayfish CNS: pathways and synaptic mechanisms. J. Neurophysiol. 54: 1305–1325, 1985.
 234. Kistler, H. B., Jr., R. D. Hawkins, J. Koester, H. W. M. Steinbusch, E. R. Kandel, and J. H. Schwartz. Distribution of serotonin‐immunoreactive cell bodies and processes in the abdominal ganglion of mature Aplysia. J. Neurosci. 5: 72–80, 1985.
 235. Klein, M., J. Camardo, and E. R. Kandel. Serotonin modulates a specific potassium current in the sensory neurons that show presynaptic facilitation in Aplysia. Proc. Natl. Acad. Sci. USA 79: 5713–5717, 1982.
 236. Klein, M., and E. R. Kandel. Presynaptic modulation of voltage‐dependent Ca2+ current: mechanism for behavioral sensitization. Proc. Natl. Acad. Sci. USA 75: 3512–3516, 1978.
 237. Klein, M., and E. R. Kandel. Mechanism of calcium current modulation underlying presynaptic facilitation and behavioral sensitization in Aplysia. Proc. Natl. Acad. Sci. USA 77: 6912–6916, 1980.
 238. Klein, M., E. Shapiro, and E. R. Kandel. Synaptic plasticity and the modulation of the Ca++ current. J. Exp. Biol. 89: 117–157, 1980.
 239. Koester, J., and E. R. Kandel. Further identification of neurons in the abdominal ganglion of Aplysia using behavioral criteria. Brain Res. 121: 1–20, 1977.
 240. Konorski, J. Integrative Activity of the Brain: An Interdisciplinary Approach. Chicago, IL: Univ. of Chicago Press, 1967.
 241. Kovac, M. P., W. J. Davis, E. M. Matera, A. Morielli, and R. P. Croll. Learning: neural analysis in the isolated brain of a previously trained mollusc, Pleurobranchaea californica. Brain Res. 331: 275–284, 1985.
 242. Krasne, F. B. Excitation and habituation of the crayfish escape reflex: the depolarization response in lateral giant fibers of the isolated abdomen. J. Exp. Biol. 50: 29–46, 1969.
 243. Krasne, F. B., and J. S. Bryan. Habituation: regulation through presynaptic inhibition. Science Wash. DC 182: 590–592, 1973.
 244. Krasne, F. B., and D. L. Glanzman. Sensitization of the crayfish lateral giant escape reaction. J. Neurosci. 6: 1013–1020, 1986.
 245. Krasne, F. B., and A. Roberts. Habituation of the crayfish escape response during release from inhibition induced by picrotoxin. Nature Lond. 215: 769–770, 1967.
 246. Krasne, F. B., and K. S. Woodsmall. Waning of the crayfish escape response as a result of repeated stimulation. Anim. Behav. 17: 416–424, 1969.
 247. Krasusky, V. K. General nature of changes of food conditioned reflexes in dogs following a surgical lesion of the cerebellum. Zh. Vyssh. Nervn. Deyat. IM IP Pavlova 7: 733–740, 1957.
 248. Kraus, N., and J. F. Disterhoft. Response plasticity of single neurons in rabbit auditory association cortex during tone‐signalled learning. Brain Res. 246: 205–215, 1982.
 249. Kuo, J. F., and P. Greengard. Cyclic nucleotide‐dependent protein kinases. IV. Widespread occurrence of adenosine 3′, 5′ ‐monophosphate‐dependent protein kinase in various tissues and phyla of the animal kingdom. Proc. Natl. Acad. Sci. USA 64: 1349–1355, 1969.
 250. Kupfermann, I., T. J. Carew, and E. R. Kandel. Local, reflex and central commands controlling gill and siphon movements in Aplysia. J. Neurophysiol. 37: 996–1019, 1974.
 251. Kupfermann, I., V. Castellucci, H. Pinsker, and E. R. Kandel. Neuronal correlates of habituation and dishabituation of the gill‐withdrawal reflex in Aplysia. Science Wash. DC 167: 1743–1745, 1970.
 252. Kupfermann, I., and E. R. Kandel. Neuronal controls of a behavioral response mediated by the abdominal ganglion of Aplysia. Science Wash. DC 164: 847–850, 1969.
 253. Kupfermann, I., and H. Pinsker. Plasticity in Aplysia neurons and some simple neuronal models of learning. In: Reinforcement and Behavior, edited by J. Tapp. New York: Academic, 1969, p. 356–386.
 254. Land, P. W., and T. Crow. Serotonin immunoreactivity in the circumesophageal nervous system of Hermissenda crassicornis. Neurosci. Lett. 62: 199–205, 1985.
 255. Lashley, K. S. Brain Mechanisms and Intelligence: A Quantitative Study of Injuries to the Brain. Chicago, IL: Univ. of Chicago Press, 1929.
 256. Lavond, D. G., T. L. Hembree, and R. F. Thompson. Effect of kainic acid lesions of the cerebellar interpositus nucleus on eyelid conditioning in the rabbit. Brain Res. 326: 179–182, 1985.
 257. Lavond, D. G., J. S. Lincoln, D. A. McCormick, and R. F. Thompson. Effect of bilateral lesions of the dentate and interpositus cerebellar nuclei on conditioning of heart‐rate and nictitating membrane/eyelid responses in the rabbit. Brain Res. 305: 323–330, 1984.
 258. Lavond, D. G., D. A. McCormick, G. A. Clark, D. T. Holmes, and R. F. Thompson. Effects of ipsilateral rostral pontine reticular lesions on retention of classically conditioned nictitating membrane and eyelid responses. Physiol. Psychol. 9: 335–339, 1981.
 259. Lavond, D. G., D. A. McCormick, and R. F. Thompson. A nonrecoverable learning deficit. Physiol. Psychol. 12: 103–110, 1984.
 260. Lederhendler, I., S. Gart, and D. L. Alkon. Classical conditioning of Hermissenda: origin of a new response. J. Neurosci. 6: 1325–1331, 1986.
 261. Lee, R. M. Aplysia behavior: effect of contingent water‐level variation. Commun. Behav. Biol. Part A Orig. Artic. 4: 157–164, 1969.
 262. Lee, R. M. Aplysia behavior: operant‐response differentiation. Proc. Annu. Conv. Am. Psychol. Assoc. 78: 249–250, 1970.
 263. Lee, R. M. Conditioning of Pleurobranchaea. Science Wash. DC 193: 72–73, 1976.
 264. Levitan, I. B., and S. H. Barondes. Octopamine‐ and serotonin‐stimulated phosphorylation of specific protein in the abdominal ganglion of Aplysia californica. Proc. Natl. Acad. Sci. USA 71: 1145–1148, 1974.
 265. Levy, W. B., and O. Steward. Synapses as associative memory elements in the hippocampal formation. Brain Res. 175: 233–245, 1979.
 266. Levy, W. B., and O. Steward. Temporal contiguity requirements for long‐term associative potentiation/depression in the hippocampus. Neuroscience 8: 791–797, 1983.
 267. Lickey, M. Learned behavior in Aplysia vaccaria. J. Comp. Physiol. Psychol. 66: 712–718, 1968.
 268. Lifshitz, N. N. Influence of cerebellar ablation on conditioned reflexes of dogs. Tr. Fiziol. Inst. Pavlova Acad. Sci. Moscow 2: 11–50, 1947.
 269. Lincoln, J. S., D. A. McCormick, and R. F. Thompson. Ipsilateral cerebellar lesions prevent learning of the classically conditioned nictitating membrane/eyelid response. Brain Res. 242: 190–193, 1982.
 270. Lisman, J. E. A mechanism for memory storage insensitive to molecular turnover: a bistable autophosphorylating enzyme. Proc. Natl. Acad. Sci. USA 82: 3055–3057, 1985.
 271. Livingstone, M. S., P. P. Sziber, and W. G. Quinn. Loss of calcium/calmodulin responsiveness in adenylate cyclase of rutabaga, a Drosophila learning mutant. Cell 37: 205–215, 1984.
 272. Livingstone, M. S., and B. L. Tempel. Genetic dissection of monoamine neurotransmitter synthesis in Drosophila. Nature Lond. 303: 67–70, 1983.
 273. London, J. A., and R. Gillette. Mechanism for food avoidance learning in the central pattern generator of feeding behavior of Pleurobranchaea californica. Proc. Natl. Acad. Sci. USA 83: 4058–4062, 1986.
 274. Longley, R. D., and A. J. Longley. Serotonin immunoreactivity of neurons in the gastropod Aplysia californica. J. Neurobiol. 17: 339–358, 1986.
 275. Lukowiak, K. In vitro classical conditioning of a gill withdrawal reflex in Aplysia: neural correlates and possible neural mechanisms. J. Neurobiol. 17: 83–101, 1986.
 276. Lukowiak, K., and C. Sahley. The in vitro classical conditioning of the gill withdrawal reflex of Aplysia californica. Science Wash. DC 212: 1516–1518, 1981.
 277. Lynch, G., and M. Baudry. The biochemistry of memory: a new and specific hypothesis. Science Wash. DC 224: 1057–1063, 1984.
 278. Lynch, G., S. Halpain, and M. Baudry. Effects of high‐frequency synaptic stimulation on glutamate receptor binding studied with a modified in vitro hippocampal slice preparation. Brain Res. 244: 101–111, 1982.
 279. Lynch, G., J. L. McGaugh, and N. M. Weinberger (editors). Neurobiology of Learning and Memory. New York: Guil ford, 1984.
 280. MacDonald, J. F., and J. A. Pearson. Some observations on habituation of the flexor reflex in the rat: the influence of strychnine, bicuculline, spinal transection, and decerebration. J. Neurobiol. 10: 67–78, 1979.
 281. MacDonald, J. F., and J. A. Pearson. Inhibition of spinal interneuronal activity by repeated cutaneous stimulation: a possible substrate of flexor reflex habituation. J. Neurobiol. 10: 79–92, 1979.
 282. Mackey, S. L., R. D. Hawkins, and E. R. Kandel. Neurons in 5‐HT containing region of cerebral ganglia produce facilitation of LE cells in Aplysia. Soc. Neurosci. Abstr. 12: 1340, 1986.
 283. Mackintosh, N. J. The Psychology of Animal Learning. London: Academic, 1974.
 284. Madden, J., IV, D. A. Haley, J. D. Barchas, and R. F. Thompson. Microinfusion of picrotoxin into the caudal red nucleus selectively abolishes the classically conditioned nictitating membrane/eyelid response in the rabbit. Soc. Neurosci. Abstr. 9: 830, 1983.
 285. Madison, D. V., and R. A. Nicoll. Noradrenaline blocks accommodation of pyramidal cell discharge in the hippocampus. Nature Lond. 299: 636–638, 1982.
 286. Malinow, R., and J. P. Miller. Postsynaptic hyperpolarization during conditioning reversibly blocks induction of longterm potentiation. Nature Lond. 320: 529–530, 1986.
 287. Mamounas, L. A., J. Madden IV, J. D. Barchas, and R. F. Thompson. Microinfusion of GABA antagonists into the cerebellar deep nuclei selectively abolishes the classically conditioned eyelid response in the rabbit. Soc. Neurosci. Abstr. 9: 830, 1983.
 288. Mamounas, L. A., R. F. Thompson, G. Lynch, and M. Baudry. Classical conditioning of the rabbit eyelid response increases glutamate receptor binding in hippocampal synaptic membranes. Proc. Natl. Acad. Sci. USA 81: 2548–2552, 1984.
 289. Marr, D. A theory of cerebellar cortex. J. Physiol. Lond. 202: 437–470, 1969.
 290. Martin, G. K., T. Land, and R. F. Thompson. Classical conditioning of the rabbit (Oryctolagus cuniculus) nictitating membrane response, with electrical brain stimulation as the unconditioned stimulus. J. Comp. Physiol. Psychol. 94: 216–226, 1980.
 291. Matsumura, M., and C. D. Woody. Excitability changes of facial motoneurons of cats related to conditioned and unconditioned facial motor responses. In: Conditioning: Representation of Involved Neural Functions, edited by C. D. Woody. New York: Plenum, 1982, p. 451–457.
 292. Mauk, M. D., J. E. Steinmetz, and R. F. Thompson. Classical conditioning using stimulation of the inferior olive as the unconditioned stimulus. Proc. Natl. Acad. Sci. USA 83: 5349–5353, 1986.
 293. Mauk, M. D., J. T. Warren, and R. F. Thompson. Selective, naloxone‐reversible morphine depression of learned behavioral and hippocampal responses. Science Wash. DC 216: 434–436, 1982.
 294. Maximova, O. A., and P. M. Balaban. Neuronal correlates of aversive learning in command neurons for avoidance behavior of Helix lucorum L. Brain Res. 292: 139–149, 1984.
 295. McCormick, D. A., G. A. Clark, D. G. Lavond, and R. F. Thompson. Initial localization of the memory trace for a basic form of learning. Proc. Natl Acad. Sci. USA 79: 2731–2735, 1982.
 296. McCormick, D. A., P. E. Guyer, and R. F. Thompson. Superior cerebellar peduncle lesions selectively abolish the ipsilaterally conditioned nictitating membrane/eyelid response of the rabbit. Brain Res. 245: 347–350, 1982.
 297. McCormick, D. A., D. G. Lavond, G. A. Clark, R. E. Kettner, C. E. Rising, and R. F. Thompson. The engram found? Role of cerebellum in classical conditioning of nictitating membrane and eyelid responses. Bull. Psychon. Soc. 18: 103–105, 1981.
 298. McCormick, D. A., D. G. Lavond, and R. F. Thompson. Neuronal responses of the rabbit brainstem during performance of the classically conditioned nictitating membrane (NM)/eyelid response. Brain Res. 271: 73–88, 1983.
 299. McCormick, D. A., J. E. Steinmetz, and R. F. Thompson. Lesions of the inferior olivary complex cause extinction of the classically conditioned eyeblink response. Brain Res. 359: 120–130, 1985.
 300. McCormick, D. A., and R. F. Thompson. Cerebellum: essential involvement in the classically conditioned eyelid response. Science Wash. DC 223: 296–299, 1984.
 301. McCormick, D. A., and R. F. Thompson. Neuronal responses of the rabbit cerebellum during acquisition and performance of a classically conditioned nictitating membrane‐eyelid response. J. Neurosci. 4: 2811–2822, 1984.
 302. McElearney, A., and J. Farley. Persistent changes in Hermissenda B photoreceptor membrane properties with associative training: a role for pharmacological modulation. Soc. Neurosci. Abstr. 9: 915, 1983.
 303. Menzel, R., and J. Erber. Learning and memory in bees. Sci. Am. 239: 102–110, 1978.
 304. Miles, F. A., D. J. Braitman, and B. M. Dow. Long‐term adaptive changes in primate vestibuloocular reflex. IV. Electrophysiological observations in flocculus of adapted monkeys. J. Neurophysiol. 43: 1477–1493, 1980.
 305. Miller, N. E. Certain facts of learning relevant to the search for its physical basis. In: The Neurosciences: A Study Program, edited by G. C. Quarton, T. Melnechuk, and F. O. Schmitt. New York: Rockefeller Univ. Press, 1967, p. 643–652.
 306. Miller, S. G., and M. B. Kennedy. Regulation of brain type II Ca2+/calmodulin‐dependent protein kinase by autophos‐phorylation. A Ca2+‐triggered molecular switch. Cell 44: 861–870, 1986.
 307. Mis, F. W., I. Gormezano, and J. A. Harvey. Stimulation of abducens nucleus supports classical conditioning of the rabbit nictitating membrane response. Science Wash. DC 206: 473–475, 1979.
 308. Montarolo, P. G., V. F. Castellucci, P. Goelet, E. R. Kandel, and S. Schacher. Long‐term facilitation of the monosynaptic connections between sensory neurons and motor neurons of the gill‐withdrawal reflex in Aplysia in dissociated cell culture. Soc. Neurosci. Abstr. 11: 795, 1985.
 309. Montarolo, P. G., P. Goelet, V. F. Castellucci, J. Morgan, E. R. Kandel, and S. Schacher. A critical peroid for macrocmolecular synthesis in long‐term heterosynaptic facilitation in Aplysia. Science Wash. DC. 234: 1249–1254, 1986.
 310. Moore, J. W., J. E. Desmond, and N. E. Berthier. The metencephalic basis of the conditioned nictitating membrane response. In: Conditioning: Representation of Involved Neural Functions, edited by C. D. Woody. New York: Plenum, 1982, p. 459–482.
 311. Morielli, A. D., E. M. Matera, M. P. Kovac, R. G. Shrum, K. J. McCormack, and W. J. Davis. Cholinergic suppression: a postsynaptic mechanism of long‐term associative training. Proc. Natl. Acad. Sci. USA 83: 4556–4560, 1986.
 312. Mpitsos, G. J., and S. D. Collins. Learning: rapid aversive conditioning in the gastropod mollusk‐Pleurobranchaea. Science Wash. DC 188: 954–957, 1975.
 313. Mpitsos, G. J., S. D. Collins, and A. D. McClellan. Learning: a model system for physiological studies. Science Wash. DC 199: 497–506, 1978.
 314. Mpitsos, G. J., and W. J. Davis. Learning: classical and avoidance conditioning in the mollusk Pleurobranchaea. Science Wash. DC 180: 317–320, 1973.
 315. Murakami, F., H. Katsumar, K. Saito, and N. Tsukahara. A quantitative study of synaptic reorganization in red nucleus neurons after lesion of the nucleus interpositus of the cat: an electron microscopic study involving intracellular injection of horseradish peroxidase. Brain Res. 242: 41–53, 1982.
 316. Nakamura, Y., and N. Mizuno. An electron microscopic study of the interpositorubral connections of the cat and the rabbit. Brain Res. 35: 283–286, 1971.
 317. Nakamura, Y., N. Mizuno, A. Konishi, and M. Sato. Synaptic organization of the red nucleus after chronic deafferentation from cerebellorubral fibers: an electron microscopic study in the cat. Brain Res. 82: 298–301, 1974.
 318. Neary, J. T., and D. L. Alkon. Protein phosphorylation/ dephosphorylation and the transient, voltage‐dependent potassium conductance in Hermissenda crassicornis. J. Biol. Chem. 258: 8979–8983, 1983.
 319. Neary, J. T., T. Crow, and D. L. Alkon. Change in a specific phosphoprotein following associative learning in Hermissenda. Nature Lond. 293: 658–660, 1981.
 320. Neary, J. T., S. A. DeRiemer, L. K. Kaczmarek, and D. L. Alkon. Ca2+ and cAMP regulation of protein phosphorylation in the Hermissenda nervous system. Soc. Neurosci. Abstr. 10: 805, 1984.
 321. O'Brien, J. H., M. B. Wilder, and C. D. Stevens. Conditioning of cortical neurons in cats with antidromic activation as the unconditioned stimulus. J. Comp. Physiol. Psychol. 91: 918–929, 1977.
 322. Ocorr, K. A., and J. H. Byrne. Membrane responses and changes in cAMP levels in Aplysia sensory neurons produced by serotonin, tryptamine, FMRFamide, and small cardioactive peptide B (SCPB). Neurosci. Lett. 55: 113–118, 1985.
 323. Ocorr, K. A., M. Tabata, and J. H. Byrne. Stimuli that produce sensitization lead to elevation of cyclic AMP levels in tail sensory neurons of Aplysia. Brain Res. 371: 190–192, 1986.
 324. Ocorr, K. A., E. T. Walters, and J. H. Byrne. Associative conditioning analog selectively increases cAMP levels of tail sensory neurons in Aplysia. Proc. Natl. Acad. Sci. USA 82: 2548–2552, 1985.
 325. Oda, Y., K. Kuwa, S. Miyasaka, and N. Tsukahara. Modification of rubral activities during classical conditioning in the cat. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci. 57: 402–405, 1981.
 326. Olds, J., J. F. Disterhoft, M. Segal, C. L. Kornblith, and R. Hirsh. Learning centers of rat brain mapped by measuring latencies of conditioned unit responses. J. Neuro physiol. 35: 202–219, 1972.
 327. Olds, J., R. Nienhuis, and M. E. Olds. Pattern of conditioned unit responses in the auditory system of rat. Exp. Neurol. 59: 209–228, 1978.
 328. Oleson, T. D., J. H. Ashe, and N. M. Weinberger. Modification of auditory and somatosensory system activity during pupillary conditioning in the paralyzed cat. J. Neurophysiol. 38: 1114–1139, 1975.
 329. Olton, D. S., J. T. Becker, and G. E. Handelmann. Hippocampus, space and memory. Behav. Brain Sci. 2: 313–365, 1979.
 330. Ono, J., and R. E. McCaman. Immunocytochemical localization and direct assays of serotonin‐containing neurons in Aplysia. Neuroscience 11: 549–560, 1984.
 331. Patterson, M. M. Effects of forward and backward classical conditioning procedures on a spinal cat hind‐limb flexor nerve response. Physiol. Psychol. 3: 86–91, 1975.
 332. Patterson, M. M., T. W. Berger, and R. F. Thompson. Neuronal plasticity recorded from cat hippocampus during classical conditioning. Brain Res. 163: 339–343, 1979.
 333. Patterson, M. M., C. F. Cegavske, and R. F. Thompson. Effects of a classical conditioning paradigm on hind‐limb flexor nerve response in immobilized spinal cats. J. Comp. Physiol. Psychol. 84: 88–97, 1973.
 334. Pavlov, I. P. Conditioned Reflexes: An Investigation of the Physiological Activity of the Cerebral Cortex. London: Oxford Univ. Press, 1927. [Transl. by G. V. Anrep.]
 335. Perlman, A. J. Central and peripheral control of siphon withdrawal reflex in Aplysia californica. J. Neurophysiol. 42: 510–529, 1979.
 336. Pinsker, H. M., W. A. Hening, T. J. Carew, and E. R. Kandel. Long‐term sensitization of a defensive withdrawal reflex in Aplysia. Science Wash. DC 182: 1039–1042, 1973.
 337. Pinsker, H., I. Kupfermann, V. Castellucci, and E. R. Kandel. Habituation and dishabituation of the gill‐withdrawal reflex in Aplysia. Science Wash. DC 167: 1740–1742, 1970.
 338. Polenchar, B. E., M. M. Patterson, D. G. Lavond, and R. F. Thompson. Cerebellar lesions abolish an avoidance response in rabbit. Behav. Neural Biol. 44: 221–227, 1985.
 339. Pollock, J. D., L. Bernier, and J. Camardo. Serotonin and cyclic adenosine 3′:5′‐monophosphate modulate the potassium current in tail sensory neurons in the pleural ganglion of Aplysia. J. Neurosci. 5: 1862–1871, 1985.
 340. Port, R. L., and M. M. Patterson. Fimbrial lesions and sensory preconditioning. Behav. Neurosci. 98: 584–589, 1984.
 341. Pribram, K. H. Languages of the Brain: Experimental Paradoxes and Principles in Neuropsychology. Englewood Cliffs, NJ: Prentice‐Hall, 1971.
 342. Prosser, C. L., and W. S. Hunter. The extinction of startle responses and spinal reflexes in the white rat. Am. J. Physiol. 117: 609–618, 1936.
 343. Quinn, W. G., W. A. Harris, and S. Benzer. Conditioned behavior in Drosophila melanogaster. Proc. Natl. Acad. Sci. USA 71: 708–712, 1974.
 344. Quinn, W. G., P. P. Sziber, and R. Booker. The Drosophila memory mutant amnesiac. Nature Lond. 277: 212–214, 1979.
 345. Rankin, C. H., and T. J. Carew. Dishabituation and sensitization emerge as separate processes during development in Aplysia. Soc. Neurosci. Abstr. 12: 398, 1986.
 346. Rayport, S. G., and S. Schacher. Synaptic plasticity in vitro: cell culture of identified Aplysia neurons mediating short‐term habituation and sensitization. J. Neurosci. 6: 759–763, 1986.
 347. Rescorla, R. A. Probability of shock in the presence and absence of CS in fear conditioning. J. Comp. Physiol. Psychol. 66: 1–5, 1968.
 348. Rescorla, R. A., and A. R. Wagner. A theory of Pavlovian conditioning: variations in the effectiveness of reinforcement and non‐reinforcement. In: Classical Conditioning II: Current Research and Theory, edited by A. H. Black and W. F. Prokasy. New York: Appleton‐Century‐Crofts, 1972, p. 64–99.
 349. Robinson, D. A. Adaptive gain control of vestibuloocular reflex by the cerebellum. J. Neurophysiol. 39: 954–969, 1976.
 350. Roeder, K. D. Organization of the ascending giant fiber system in the cockroach (Periplaneta americana). J. Exp. Zool. 108: 243–261, 1948.
 351. Rosenfield, M. E., and J. W. Moore. Red nucleus lesions disrupt the classically conditioned nictitating membrane response in rabbits. Behav. Brain Res. 10: 393–398, 1983.
 352. Ruthrich, H., H. Matthies, and T. Ott. Long‐term changes in synaptic excitability of hippocampal cell populations as a result of training. In: Neuronal Plasticity and Memory Formation, edited by C. Ajmone Marsan and H. Matthies. New York: Raven, 1982, p. 589–594.
 353. Sacktor, T. C., C. A. O'Brian, J. B. Weinstein, and J. H. Schwartz. Translocation from cytosol to membrane of protein kinase C after stimulation of Aplysia neurons with serotonin. Soc. Neurosci. Abstr. 12: 1340, 1986.
 354. Sahley, C., J. W. Rudy, and A. Gelperin. An analysis of associative learning in a terrestrial mollusc. I. Higher‐order conditioning, blocking, and a transient US pre‐exposure effect. J. Comp. Physiol. 144: 1–8, 1981.
 355. Saitoh, T., and J. H. Schwartz. Phosphorylation‐dependent subcellular translocation of a Ca2+/calmodulin‐dependent protein kinase produces an autonomous enzyme in Aplysia neurons. J. Cell Biol. 100: 835–842, 1985.
 356. Sakakibara, M., D. L. Alkon, J. T. Neary, R. DeLorenzo, R. Gould, and E. Heldman. Ca2+‐mediated reduction of K+ currents is enhanced by injection of IP3 or neuronal Ca2+/ calmodulin kinase type II. Soc. Neurosci. Abstr. 11: 956, 1985.
 357. Sastry, B. R., J. W. Goh, and A. Auyeung. Associative induction of posttetanic and long‐term potentiation in CA1 neurons of rat hippocampus. Science Wash. DC 232: 988–990, 1986.
 358. Schacher, S., P. G. Montarolo, V. F. Castellucci, E. R. Kandel, and P. Goelet. A critical period of macromolecular synthesis is necessary for the expression of long‐term facilitation of sensorimotor connections in Aplysia. Soc. Neurosci. Abstr. 12: 1338, 1986.
 359. Schmaltz, L. W., and J. Theios. Acquisition and extinction of a classically conditioned response in hippocampectomized rabbits (Oryctolagus cuniculus). J. Comp. Physiol. Psychol. 79: 328–333, 1972.
 360. Schone, H. Complex behavior. In: The Physiology of Crustacea, edited by T. H. Waterman. New York: Academic, 1961, p. 465–520.
 361. Schwartz, J. H., V. F. Castellucci, and E. R. Kandel. Functioning of identified neurons and synapses in abdominal ganglion of Aplysia in absence of protein synthesis. J. Neurophysiol. 34: 939–953, 1971.
 362. Schwartzkroin, P. A., and K. Wester. Long‐lasting facilitation of a synaptic potential following tetanization in the in vitro hippocampal slice. Brain Res. 89: 107–119, 1975.
 363. Segal, M., J. F. Disterhoft, and J. Olds. Hippocampal unit activity during classical aversive and appetitive conditioning. Science Wash. DC 175: 792–794, 1972.
 364. Segal, M., and J. Olds. Behavior of units in hippocampal circuit of the rat during learning. J. Neurophysiol. 35: 680–690, 1972.
 365. Segal, M., and J. Olds. Activity of units in the hippocampal circuit of the rat during differential classical conditioning. J. Comp. Physiol. Psychol. 82: 195–204, 1973.
 366. Sherrington, C. S. The Integrative Action of the Nervous System. New Haven, CT: Yale Univ. Press, 1906.
 367. Shuster, M. J., J. S. Camardo, S. A. Siegelbaum, and E. R. Kandel. Cyclic AMP‐dependent protein kinase closes the serotonin‐sensitive K+ channels of Aplysia sensory neurons in cell‐free membrane patches. Nature Lond. 313: 392–395, 1985.
 368. Siegelbaum, S. A., J. S. Camardo, and E. R. Kandel. Serotonin and cyclic AMP close single K+ channels in Aplysia sensory neurones. Nature Lond. 299: 413–417, 1982.
 369. Skelton, R. W., N. H. Donegan, and R. F. Thompson. Superior colliculus lesions disrupt classical conditioning to visual but not auditory stimuli. Soc. Neurosci. Abstr. 10: 132, 1984.
 370. Smith, A. M. The effects of rubral lesions and stimulation on conditioned forelimb flexion responses in the cat. Physiol. Behav. 5: 1121–1126, 1970.
 371. Smith, M. C., S. R. Coleman, and I. Gormezano. Classical conditioning of the rabbit's nictitating membrane response at backward, simultaneous and forward CS‐US intervals. J. Comp. Physiol. Psychol. 69: 226–231, 1969.
 372. Sokolov, E. N. Higher nervous functions: the orienting reflex. Annu. Rev. Physiol. 25: 545–580, 1963.
 373. Solomon, P. R. The role of dorsal hippocampus in blocking and conditioned inhibition of the rabbit's nictitating membrane response. J. Comp. Physiol. Psychol. 91: 407–417, 1977.
 374. Solomon, P. R., J. L. Lewis, J. J. Loturco, J. E. Steinmetz, and R. F. Thompson. The role of the middle cerebellar peduncle in acquisition and retention of the rabbit's classically conditioned nictitating membrane response. Bull. Psychon. Soc. 24: 75–78, 1986.
 375. Solomon, P. R., and J. W. Moore. Latent inhibition and stimulus generalization of the classically conditioned nictitating membrane response in rabbits (Oryctolagus cuniculus) following dorsal hippocampal ablations. J. Comp. Physiol. Psychol. 89: 1192–1203, 1975.
 376. Solomon, P. R., E. R. Vander Schaaf, A. C. Nobre, D. J. Weisz, and R. F. Thompson. Hippocampus and trace conditioning of the rabbit's nictitating membrane response. Soc. Neurosci. Abstr. 9: 645, 1983.
 377. Spencer, W. A., R. F. Thompson, and D. R. Nielson, Jr. Response decrement of the flexion reflex in the acute spinal cat and transient restoration by strong stimuli. J. Neurophysiol. 29: 221–239, 1966.
 378. Spencer, W. A., R. F. Thompson, and D. R. Nielson, Jr. Alterations in responsiveness of ascending and reflex pathways activated by iterated cutaneous afferent volleys. J. Neurophysiol. 29: 240–252, 1966.
 379. Spencer, W. A., R. F. Thompson, and D. R. Nielson, Jr. Decrement of ventral root electrotonus and intracellularly recorded PSPs produced by iterated cutaneous afferent volleys. J. Neurophysiol. 29: 253–273, 1966.
 380. Squire, L. R. The neuropsychology of human memory. Annu. Rev. Neurosci. 5: 241–273, 1982.
 381. Stanton, P. K., and J. M. Sarvey. Blockade of norepinephrine‐induced long‐lasting potentiation in the hippocampal dentate gyrus by an inhibitor of protein synthesis. Brain Res. 361: 276–283, 1985.
 382. Stanton, P. K., and J. M. Sarvey. Depletion of norepinephrine, but not serotonin, reduces long‐term potentiation in the dentate gyrus of rat hippocampal slices. J. Neurosci. 5: 2169–2176, 1985.
 383. Stanton, P. K., and J. M. Sarvey. The effect of highfrequency electrical stimulation and norepinephrine on cyclic AMP levels in normal versus norepinephrine‐depleted rat hippocampal slices. Brain Res. 358: 343–348, 1985.
 384. Steinmetz, J. E., D. G. Lavond, and R. F. Thompson. Classical conditioning of the rabbit‐eyelid response with mossy fiber stimulation as the conditioned stimulus. Bull. Psychon. Soc. 23: 245–248, 1985.
 385. Stent, G. S. A physiological mechanism for Hebb's postulate of learning. Proc. Natl. Acad. Sci. USA 70: 997–1001, 1973.
 386. Susswein, A. J., M. Schwarz, and E. Feldman. Learned changes of feeding behavior in Aplysia in response to edible and inedible foods. J. Neurosci. 6: 1513–1527, 1986.
 387. Swanson, L. W., T. J. Teyler, and R. F. Thompson. Hippocampal long‐term potentiation: mechanisms and implications for memory. Neurosci. Res. Prog. Bull. 20: 613–769, 1982.
 388. Tempel, B. L., M. S. Livingstone, and W. G. Quinn. Mutations in the dopa decarboxylase gene affect learning in Drosophila. Proc. Natl. Acad. Sci. USA 81: 3577–3581, 1984.
 389. Teyler, T. J., and P. Discenna. Long‐term potentiation as a candidate mnemonic device. Brain Res. Rev. 7: 15–28, 1984.
 390. Thompson, R. F. Foundations of Physiological Psychology. New York: Harper & Row, 1967.
 391. Thompson, R. F. The neurobiology of learning and memory. Science Wash. DC 233: 941–947, 1986.
 392. Thompson, R. F., T. W. Berger, S. D. Berry, F. K. Hoehler, R. E. Kettner, and D. J. Weisz. Hippocampal substrate of classical conditioning. Physiol. Psychol. 8: 262–279, 1980.
 393. Thompson, R. F., T. W. Berger, C. F. Cegavske, M. M. Patterson, R. A. Roemer, T. J. Teyler, and R. A. Young. The search for the engram. Am. Psychol. 31: 209–227, 1976.
 394. Thompson, R. F., G. A. Clark, N. H. Donegan, D. G. Lavond, J. S. Lincoln, J. Madden IV, L. A. Mamounas, M. D. Mauk, D. A. McCormick, and J. K. Thompson. Neuronal substrates of learning and memory: a “multiple trace” view. In: Neurobiology of Learning and Memory, edited by G. Lynch, J. L. McGaugh, and N. M. Weinberger. New York: Guilford, 1984, p. 137–164.
 395. Thompson, R. F., D. A. McCormick, D. G. Lavond, G. A. Clark, R. E. Kettner, and M. D. Mauk. The engram found? Initial localization of the memory trace for a basic form of associative learning. Prog. Psychobiol. Physiol. Psychol. 10: 167–196, 1983.
 396. Thompson, R. F., and W. A. Spencer. Habituation: a model phenomenon for the study of neuronal substrates of behavior. Psychol. Rev. 173: 16–43, 1966.
 397. Tosney, T., and G. Hoyle. Computer‐controlled learning in a simple system. Proc. R. Soc. Lond. B Biol. Sci. 195: 365–393, 1977.
 398. Tritt, S. H., I. P. Lowe, and J. H. Byrne. A modification of the glyoxylic acid induced histofluorescence technique for demonstration of catecholamines and serotonin in tissues of Aplysia californica. Brain Res. 259: 159–162, 1983.
 399. Tsukahara, N. Classical conditioning mediated by the red nucleus: an approach beginning at the cellular level. In: Neurobiology of Learning and Memory, edited by G. Lynch, J. L. McGaugh, and N. M. Weinberger. New York: Guilford, 1984, p. 165–180.
 400. Tsukahara, N., and Y. Fujito. Physiological evidence of formation of new synapses from cerebrum in the red nucleus following cross‐union of forelimb nerves. Brain Res. 106: 184–188, 1976.
 401. Tsukahara, N., Y. Fujito, Y. Oda, and J. Maeda. Formation of functional synapses in adult red nucleus from the cerebrum following cross‐innervation of forelimb flexor and extensor nerves. I. Appearance of new synaptic potentials. Exp. Brain Res. 45: 1–12, 1982.
 402. Tsukahara, N., H. Hultborn, F. Murakami, and Y. Fujito. Electrophysiological study of formation of new synapses and collateral sprouting in red nucleus neurons after partial denervation. J. Neurophysiol. 38: 1359–1372, 1975.
 403. Tsukahara, N., and K. Kosaka. The mode of cerebral excitation of red nucleus neurons. Exp. Brain Res. 5: 102–117, 1968.
 404. Tsukahara, N., F. Murakami, and H. Hultborn. Electrical constants of neurons of the red nucleus. Exp. Brain Res. 23: 49–64, 1975.
 405. Tsukahara, N., and Y. Oda. Appearance of new synaptic potentials at cortico‐rubral synapses after the establishment of classical conditioning. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci. 57: 398–401, 1981.
 406. Tsukahara, N., Y. Oda, and T. Notsu. Classical conditioning mediated by the red nucleus in the cat. J. Neurosci. 1: 72–79, 1981.
 407. Tully, T., and W. G. Quinn. Classical conditioning and retention in normal and mutant Drosophila melanogaster. J. Comp. Physiol. A Sens. Neural Behav. Physiol. 157: 263–277, 1985.
 408. Turker, K. S., and J. S. Miles. Climbing fiber lesions disrupt conditioning of the nictitating membrane response in the rabbit. Brain Res. 363: 376–378, 1986.
 409. Van Harreveld, A., and C. A. G. Wiersma. The triple innervation of crayfish muscle and its function in contraction and inhibition. J. Exp. Biol. 14: 448–461, 1937.
 410. Wall, J. T., C. M. Gibbs, J. L. Broyles, and D. H. Cohen. Modification of neuronal discharge along the ascending tectofugal pathway during visual conditioning. Brain Res. 342: 67–76, 1985.
 411. Wall, P. D. Habituation and post‐tetanic potentiation in the spinal cord. In: Short‐term Changes in Neural Activity and Behavior, edited by G. Horn and R. A. Hinde. Cambridge, UK: Cambridge Univ. Press, 1970, p. 181–210.
 412. Walsh, J. P., and J. H. Byrne. Forskolin mimics and blocks a serotonin‐sensitive decreased K+ conductance in tail sensory neurons of Aplysia. Neurosci. Lett. 52: 7–11, 1984.
 413. Walters, E. T., and J. H. Byrne. Associative conditioning of single sensory neurons suggests a cellular mechanism for learning. Science Wash. DC 219: 405–408, 1983.
 414. Walters, E. T., and J. H. Byrne. Long‐term enhancement produced by activity‐dependent modulation of Aplysia sensory neurons. J. Neurosci. 5: 662–672, 1985.
 415. Walters, E. T., J. H. Byrne, T. J. Carew, and E. R. Kandel. Mechanoafferent neurons innervating tail of Aplysia. II. Modulation by sensitizing stimuli. J. Neurophysiol. 50: 1543–1559, 1983.
 416. Walters, E. T., T. J. Carew, and E. R. Kandel. Classical conditioning in Aplysia californica. Proc. Natl. Acad. Sci. USA 76: 6675–6679, 1979.
 417. Walters, E. T., T. J. Carew, and E. R. Kandel. Associative learning in Aplysia: evidence for conditioned fear in an invertebrate. Science Wash. DC 211: 504–506, 1981.
 418. Weinberger, N. M. Effects of conditioned arousal on the auditory system. In: The Neural Basis of Behavior, edited by A. L. Bechman. New York: Spectrum, 1982, p. 63–91.
 419. Weinberger, N. M. Sensory plasticity and learning: the magnocellular medial geniculate nucleus of the auditory system. In: Conditioning: Representation of Involved Neural Functions, edited by C. D. Woody. New York: Plenum, 1982, p. 697–710.
 420. Weinberger, N. M., W. Hopkins, and D. M. Diamond. Physiological plasticity of single neurons in auditory cortex of the cat during acquisition of the pupillary conditioned response. I. Primary field (AI). Behav. Neurosci. 98: 171–188, 1984.
 421. Weiskrantz, L., and E. G. Warrington. Conditioning in amnesic patients. Neuropsychologia 17: 187–194, 1979.
 422. Weisz, D. J., G. A. Clark, and R. F. Thompson. Increased responsivity of dentate granule cells during nictitating membrane response conditioning in rabbit. Behav. Brain Res. 12: 145–154, 1984.
 423. Weisz, D. J., G. A. Clark, B. Yang, R. F. Thompson, and P. R. Solomon. Activity of dentate gyrus during NM conditioning in rabbit. In: Conditioning: Representation of Involved Neural Functions, edited by C. D. Woody. New York: Plenum, 1982, p. 131–145.
 424. Weisz, D. J., P. R. Solomon, and R. F. Thompson. The hippocampus appears necessary for trace conditioning. Bull. Psychon. Soc. Abstr. 193: 244, 1980.
 425. West, A., E. Barnes, and D. L. Alkon. Primary changes of voltage responses during retention of associative learning. J. Neurophysiol. 48: 1243–1255, 1982.
 426. Wickelgren, B. G. Habituation of spinal motorneurons. J. Neurophysiol. 30: 1404–1423, 1967.
 427. Wickelgren, B. G. Habituation of spinal interneurons. J. Neurophysiol. 30: 1424–1438, 1967.
 428. Wild, J. M., and D. H. Cohen. Invariance of retinal output during visual learning. Brain Res. 331: 127–135, 1985.
 429. Wine, J. J., and F. B. Krasne. The organization of escape behavior in the crayfish. J. Exp. Biol. 56: 1–18, 1972.
 430. Wine, J. J., F. B. Krasne, and L. Chen. Habituation and inhibition of the crayfish lateral giant fibre escape response. J. Exp. Biol. 62: 771–782, 1975.
 431. Woodruff‐Pak, D. S., D. G. Lavond, and R. F. Thompson. Trace conditioning: abolished by cerebellar nuclear lesions but not lateral cerebellar cortex aspirations. Brain Res. 348: 249–260, 1985.
 432. Woody, C. D. Conditioned eye blink: gross potential activity at coronal‐precruciate cortex of the cat. J. Neurophysiol. 33: 838–850, 1970.
 433. Woody, C. D. (editor). Conditioning: Representation of Involved Neural Functions. New York: Plenum, 1982.
 434. Woody, C. D., and P. Black‐Cleworth. Differences in excitability of cortical neurons as a function of motor projection in conditioned cats. J. Neurophysiol. 36: 1104–1116, 1973.
 435. Woody, C. D., and G. Brozek. Changes in evoked responses from facial nucleus of cat with conditioning and extinction of an eye blink. J. Neurophysiol. 32: 717–726, 1969.
 436. Woody, C. D., E. H.‐J. Kim, and N. E. Berthier. Effects of hypothalamic stimulation on unit responses recorded from neurons of sensorimotor cortex of awake cats during conditioning. J. Neurophysiol. 49: 780–791, 1983.
 437. Woody, C. D., J. D. Knispel, T. J. Crow, and P. A. Black‐Cleworth. Activity and excitability to electrical current of cortical auditory receptive neurons of awake cats as affected by stimulus association. J. Neurophysiol. 39: 1045–1061, 1976.
 438. Woody, C. D., Y. Oomura, E. Gruen, J. Miyake, and A. Nerov. Attempts to rapidly condition increased activity to click in single cortical neurons of awake cats using glabella tap and iontophoretically applied glutamate. Soc. Neurosci. Abstr. 10: 25, 1984.
 439. Woody, C. D., B. E. Swartz, and E. Gruen. Effects of acetylcholine and cyclic GMP on input resistance of cortical neurons in awake cats. Brain Res. 158: 373–395, 1978.
 440. Woody, C. D., N. N. Vassilevsky, and J. Engel, Jr. Conditioned eye blink: unit activity at coronal‐precruciate cortex of the cat. J. Neurophysiol. 33: 851–864, 1970.
 441. Woody, C., P. Yarowsky, J. Owens, P. Black‐Cleworth, and T. Crow. Effect of lesions of cortical motor areas on acquisition of conditioned eye blink in the cat. J. Neurophysiol. 37: 385–394, 1974.
 442. Woollacott, M. H., and G. Hoyle. Membrane resistance changes associated with single identified neuron learning. Soc. Neurosci. Abstr. 2: 339, 1976.
 443. Woollacott, M., and G. Hoyle. Neural events underlying learning in insects: changes in pacemaker. Proc. R. Soc. Lond. B Biol. Sci. 195: 395–415, 1977.
 444. Yeo, C. H., M. J. Hardiman, and M. Glickstein. Discrete lesions of the cerebellar cortex abolish the classically conditioned nictitating membrane response of the rabbit. Behav. Brain Res. 13: 261–266, 1984.
 445. Yeo, C. H., M. J. Hardiman, and M. Glickstein. Classical conditioning of the nictitating membrane response of the rabbit. I. Lesions of the cerebellar nuclei. Exp. Brain Res. 60: 87–98, 1985.
 446. Yeo, C. H., M. J. Hardiman, and M. Glickstein. Classical conditioning of the nictitating membrane response. II. Lesions of the cerebellar cortex. Exp. Brain Res. 60: 99–113, 1985.
 447. Yeo, C. H., M. J. Hardiman, and M. Glickstein. Classical conditioning of the nictitating membrane response. III. Connections of cerebral lobule HVI. Exp. Brain Res. 60: 114–126, 1985.
 448. Yeo, C. H., M. J. Hardiman, and M. Glickstein. Classical conditioning of the nictitating membrane response. IV. Lesions of the inferior olive. Exp. Brain Res. 63: 81–92, 1986.
 449. Young, R. A., C. F. Cegavske, and R. F. Thompson. Toneinduced changes in excitability of abducens motoneurons and of the reflex path of nictitating membrane response in rabbit (Oryctolagus cuniculus). J. Comp. Physiol. Psychol. 90: 424–434, 1976.
 450. Yovell, Y., Y. Dudai, and T. W. Abrams. Quantitative analysis of Ca2+/calmodulin activated adenylate cyclase from Aplysia, Drosophila and rat brain: possible relevance for associative learning. Soc. Neurosci. Abstr. 12: 400, 1986.
 451. Zilber‐Gachelin, N. F. Experiences de sensibilisation chez la Blatte. J. Physiol. Paris 58: 276–277, 1966.
 452. Zilber‐Gachelin, N. F., and M. P. Chartier. Modification of the motor reflex responses due to repetition of the peripheral stimulus in the cockroach. I. Habituation at the level of an isolated abdominal ganglion. J. Exp. Biol. 59: 359–381, 1973.
 453. Zilber‐Gachelin, N. F., and M. P. Chartier. Modification of the motor reflex responses due to repetition of the peripheral stimulus in the cockroach. II. Conditions of activation of the motoneurones. J. Exp. Biol. 59: 383–403, 1973.
 454. Zucker, R. S. Crayfish escape behavior and central synapses. I. Neural circuit exciting lateral giant fiber. J. Neurophysiol. 35: 599–620, 1972.
 455. Zucker, R. S. Crayfish escape behavior and central synapses. II. Physiological mechanisms underlying behavioral habituation. J. Neurophysiol. 35: 621–637, 1972.
 456. Zucker, R. S., D. Kennedy, and A. I. Selverston. Neuronal circuit mediating escape responses in crayfish. Science Wash. DC 173: 645–650, 1971.

Contact Editor

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

* Required Field

How to Cite

Robert D. Hawkins, Gregory A. Clark, Eric R. Kandel. Cell Biological Studies of Learning in Simple Vertebrate and Invertebrate Systems. Compr Physiol 2011, Supplement 5: Handbook of Physiology, The Nervous System, Higher Functions of the Brain: 25-83. First published in print 1987. doi: 10.1002/cphy.cp010502