The development of the nervous system takes place in two main steps: first an extensive preliminary network is formed and then it is pruned and trimmed to establish the final form. This refinement is achieved by mechanisms that include cell death, selective growth and loss of neurites and the stabilization and elimination of synapses. The focus of this review is on selective neurite retraction during development, with particular emphasis on the role of electrical activity.

In many developing vertebrate and invertebrate neurones, the frequency and duration of ongoing impulse activity determine the final arborizations and the pattern of connections. When impulse traffic is silenced, axons fail to retract branches that had grown to inappropriate destinations in the mammalian visual system, cerebellum and neuromuscular junctions. Similarly, in crustaceans, Drosophila melanogaster and leeches, refinements in axonal morphology during development are influenced by impulse activity. From experiments made in culture, it has been possible to mimic these events and to show a clear link between the density of voltage-activated calcium channels in a neurite and its retraction following stimulation. The distribution of these calcium channels in turn is determined by the substratum with which the neurites are in contact or by the formation of synapses. Several lines of evidence suggest that calcium entry into the growth cone leads to collapse by disruption of actin filaments. One candidate for coupling membrane events to neurite retraction is the microfilament-associated protein gelsolin which, in its calcium-activated state, severs actin filaments. Open questions that remain concern the differential effects of activity on dendrites and axons as well as the mechanisms by which the growth cone integrates information derived from stimuli in the cell and in the extracellular environment.

The first phase of neural development, the laying down of a preliminary network, is characterized by an initial overproduction of processes and synaptic connections. This phase is followed by a second stage of refinement and pruning, during which synapses are eliminated and neurites retract. Many signals are known that promote neurite outgrowth. In addition, evidence is accumulating that signals leading to cessation of growth and retraction of processes play an important role in defining morphology. Growth-inhibitory activity has been associated with components of the extracellular matrix, membrane-associated proteins and diffusible factors (Schwab et al. 1993). Electrical activity of the neurones themselves was one of the earliest factors shown to cause rearrangement of axonal arborizations (Wiesel and Hubel, 1963; LeVay et al. 1978; Shatz and Stryker, 1978; Fields and Nelson, 1992).

Mammalian visual system

Retinal ganglion cells

The role that electrical activity plays in the proper development of axonal arborizations and the specific spatial patterns of synaptic connections has been examined in the mammalian visual system (Shatz, 1990; Katz, 1993). Retinal ganglion cells provide signals to the lateral geniculate nucleus (LGN). The terminal arborizations of the retinal ganglion cells in the adult LGN are segregated into eye-specific layers (Minkowski, 1920; Hayhow, 1958; Hickey and Guillery, 1974; Bowling and Michael, 1980) and are arranged within each layer in a retinotopic order, reflecting the neighbour relationships of the ganglion cells in the retina (Hubel and Wiesel, 1961, 1972). Initial outgrowth of foetal retinal ganglion neurones does not reflect this orderly arrangement: the retinogeniculate terminals from the two eyes overlap and form active synapses on the same postsynaptic cell (Rakic, 1976, 1977; Shatz and Kirkwood, 1984). The eye-specific layers in the LGN emerge as axons from the two eyes gradually remodel by withdrawing branches from inappropriate territory and growing extensive terminal arborizations within appropriate areas (Rakic and Riley, 1983; Sretavan and Shatz, 1986). Remodeling of axonal arborizations requires spontaneous firing in the retina and the LGN. Thus, tetrodotoxin (TTX, which blocks Na+-dependent action potentials) prevents the segregation of retinal terminals into eye-specific layers in the LGN and results in a major increase in the total length of retinogeniculate axons (Shatz and Stryker, 1988; Sretavan et al. 1988). By contrast, the dendrites of the retinal ganglion cells acquire higher-order branches and many spines prenatally which are subsequently lost (Dann et al. 1988; Ramoa et al. 1988) even in the absence of neural activity (Wong et al. 1991).

Lateral geniculate neurones

Geniculocortical afferents terminate in layer IV of the visual cortical area 17 in alternating eye-specific bands, the ocular dominance columns (Hubel and Wiesel, 1972; LeVay and Gilbert, 1976; Shatz et al. 1977; Anderson et al. 1988). The terminal arborizations of mature geniculocortical afferents are arranged in discrete clusters of collaterals 400–600 μm in width, alternating with zones relatively free of innervation from that eye (Rakic, 1976, 1977; Ferster and LeVay, 1978; Humphrey et al. 1985a,b). During the initial phase of development, however, segregation of fibres derived from the two eyes is not yet evident and the terminal arborization is formed by long branches that sparsely innervate a wide region of area 17 (Rakic, 1976, 1977; Antonini and Stryker, 1993). These arborizations are then remodelled: the arborization density increases and the terminal arborizations develop short and confined collaterals and form distinct patches, coinciding with the segregation of left and right eye afferents. This remodelling is in part due to elimination of fibres projecting to inappropriate areas and in part to growth extending to the appropriate region (Antonini and Stryker, 1993). Neuronal activity of the afferent visual pathway plays an important role in this reshaping. Blockade by binocular injections of TTX did not change the complexity and extent of terminal arborizations, but perturbed the refinement of ocular dominance columns (Reiter et al. 1986; Stryker and Harris, 1986; Antonini and Stryker, 1993). As in retinal ganglion cells, growth of dendrites of LGN neurones is independent of neural activity, except for spine density which increases threefold in the presence of TTX (Dalva et al. 1994).

Cortical neurones

Pyramidal cells in layer 2/3 of adult cat striate cortex have long, horizontal axon collaterals in layers 2/3 and 5. These collaterals form periodic ‘clusters’ of finer axon branches that link columns of similar orientation selectivity (Gilbert and Wiesel, 1989). During the first three postnatal weeks, there is a steady increase in the number of long unbranched horizontal axon collaterals extending from the efferent axons of the pyramidal cells and arranged in crude clusters (Callaway and Katz, 1990). During the fourth postnatal week, the number of these collaterals decreases and axonal arborizations take on a more clustered appearance. This refinement is due mainly to the specific elimination of inappropriately projecting collaterals, but selective axon outgrowth may also occur. Cluster refinement depends on visual activity. Thus, binocular deprivation does not inhibit the maturation of horizontal axonal arborizations and development into crude clusters, but the specificity of axon rearrangements by elimination of incorrectly projecting collaterals and perhaps addition of axons to correct columns is prevented (Callaway and Katz, 1991). In strabismic kittens, in which images on the two retinae cannot be brought into register, tangential collaterals preferentially connect ‘orientation clusters’ driven by the same eye column, while in normal animals no such preference is obvious. These results show that the proper construction of horizontal connections requires not only visual activity per se, but correlations in the patterns of cellular activity (Löwel and Singer, 1992).

Cerebellum and corpus callosum

Another instance of axon elimination occurs in the cerebellum. The Purkinje cells of the newborn rat are innervated by several different climbing fibres, all but one of which are lost over the first few weeks of life (Delhaye-Bouchaud et al. 1975; Crepel et al. 1976; Mariani and Changeux, 1981). Mutations or treatments that affect afferent input to Purkinje cells result in persistence of multiple innervation by climbing fibres (Mariani, 1982, and references therein) and dramatic alterations in the morphology of Purkinje cell dendritic branching (Berry and Bradley, 1976; Privat and Drian, 1976; Mariani et al. 1977; Ito, 1984). Recently it has been shown that activation of cerebellar N-methyl-D-aspartate (NMDA) receptors plays a critical role in the elimination of climbing fibres (Rabacchi et al. 1992).

Restructuring of axonal arborizations is not restricted to the fine-tuning of neuronal circuits. In some situations, process elimination is responsible for the removal of entire neural pathways. The elimination of long collateral projections has been observed during the development of the corticospinal projections (Stanfield et al. 1982; Stanfield, 1992) and the corpus callosum (Innocenti, 1981; O’Leary et al. 1981; Ivy and Killackey, 1982). Changes in the level or patterning of visual activity to the developing cortex affect the pruning of callosal axons. Thus, neural activity seems to be necessary for the remodelling of callosal projections from the visual cortex (Lund et al. 1978; Innocenti and Frost, 1979, 1980; Rhoades and Dellacroce, 1980; Cusick and Lund, 1982; Berman and Payne, 1983; Cowan et al. 1984; Innocenti et al. 1985; Frost and Moy, 1989; Frost et al. 1990; O’Leary, 1992; Grigonis and Murphy, 1994).

Neuromuscular junction

The first evidence for local elimination of neuronal processes came from studies of the developing neuromuscular junction (Redfern, 1970; Brown et al. 1976; Van Essen, 1982). In the neonatal rat (and other mammals), skeletal muscle fibres are innervated by as many as five or six separate axons. During early postnatal life, all axons except one are eliminated, so that eventually every muscle fibre is contacted by only one motor axon (Redfern, 1970; Brown et al. 1976; Purves and Lichtman, 1980; Thompson, 1986; Goodman and Shatz, 1993). Axonal retraction (rather than degeneration) is the predominant mechanism by which excess neuromuscular synapses are removed (Riley, 1977, 1981). This elimination of axonal branches is dependent on neural activity (O’Brien et al. 1978; Thompson, 1983). The process of synapse elimination is slowed by decreased neural activity (Benoit and Changeux, 1975; Riley, 1978; Thompson et al. 1979) and hastened by enhanced activity (O’Brien et al. 1978; Thompson, 1983).

Poo and his collaborators have shown in an elegant system that the elimination of synapses at the neuromuscular junction is a result of competitive interactions between different motoneurones innervating the same muscle fibre. A single embryonic muscle cell in culture was innervated by two spinal neurones. Stimulation of one neurone resulted in functional suppression of the synapse made by the other unstimulated neurone. When both neurones were stimulated concurrently, such synaptic suppression was not observed (Lo and Poo, 1991). The synaptic suppression was dependent on postsynaptic activity and a rise in postsynaptic cytosolic calcium level. It could be elicited by postsynaptic application of acetylcholine in the absence of synchronous presynaptic stimulation. Postsynaptic depolarization in the absence of acetylcholine elicited a reduced response. These results suggest that the activation of the postsynaptic acetylcholine receptors plays a dominant role in the suppression of a synapse at the neuromuscular junction (Dan and Poo, 1992; Lo and Poo, 1994; Lo et al. 1994). A transient retrograde messenger has been suggested that acts on the presynaptic terminal to weaken the synapse unless the presynaptic terminal itself has elevated calcium levels due to synchronous activity (Dan and Poo, 1992, 1994a, b) (Fig. 1). How synaptic suppression leads to axon retraction remains unclear. Loss of adhesion due to the redistribution of receptors away from the inactive synapse and towards the active synapse has been suggested as a first step in the process of axonal retraction (Poo and Young, 1990; Dan and Poo, 1994a).

Fig. 1.

Scheme for retrograde signalling at the developing neuromuscular junction. In the pathway on the right, calcium influx into the postsynaptic cell leads to the production of membrane-permeant retrograde factors. These factors diffuse from the postsynaptic to the presynaptic cell, where they modulate synaptic activity. Alternatively, according to the scheme on the left, a raised postsynaptic calcium concentration could stimulate the exocytosis of vesicles containing retrograde factors which then bind to retrograde factor receptors localized on the presynaptic membrane. The activation of these receptors eventually results in a change of synaptic activity (reproduced with permission, after Dan and Poo, 1994a). ACh, acetylcholine.

Fig. 1.

Scheme for retrograde signalling at the developing neuromuscular junction. In the pathway on the right, calcium influx into the postsynaptic cell leads to the production of membrane-permeant retrograde factors. These factors diffuse from the postsynaptic to the presynaptic cell, where they modulate synaptic activity. Alternatively, according to the scheme on the left, a raised postsynaptic calcium concentration could stimulate the exocytosis of vesicles containing retrograde factors which then bind to retrograde factor receptors localized on the presynaptic membrane. The activation of these receptors eventually results in a change of synaptic activity (reproduced with permission, after Dan and Poo, 1994a). ACh, acetylcholine.

Invertebrate nervous systems

Crayfish sensory neurones, lobster motoneurones and Drosophila motoneurones

The importance of electrical activity for the proper development of synaptic connections has been analyzed for regenerating mechanoreceptor neurones in the crayfish. After the activity of a group of mechanoreceptors known to innervate an identified interneurone is silenced by mechanically preventing the movement of the transducer hair, the target interneurone accepts innervation from inappropriate mechanoreceptors (Krasne, 1987). Nothing is as yet known about the effects of electrical activity on changes in the axonal morphology that might accompany or result from this synaptic competition.

In embryonic, larval and early juvenile lobsters, an identified motoneurone innervates the muscles of its own segment and the two adjacent segments. In late juvenile and adult animals, the same motoneurone innervates only the muscle of its own segment (Stephens and Govind, 1981). These motoneurones establish their target fields by selective elimination of functional synapses initially distributed in a wide peripheral field.

An elegant genetic approach suggests that electrical activity affects the growth and branching pattern of Drosophila motoneurones (Budnik et al. 1990). Hyperexcitable Drosophila mutants show a phenotype of anaesthesia-induced leg shaking and the different alleles have been linked to potassium channels (Ganetzky and Wu, 1986; Baumann et al. 1987; Kamb et al. 1987; Papazian et al. 1987; Tempel et al. 1987). In these mutants, higher-order axonal branches of motoneurones are more complex and the density of varicosities on the neurites is higher than in wild-type flies (Budnik et al. 1990). Temperature-sensitive mutations that affect sodium channels cause flies to become paralyzed within seconds after exposure to elevated temperatures (Ganetzky and Wu, 1986). In these hypoexcitable mutants, a slight decrease in axonal branching of motoneurones has been observed (Budnik et al. 1990). Interestingly, the morphology of sensory neurones in both types of mutants reveals no obvious abnormality (Burg and Wu, 1986, 1989).

Changes in the branching pattern of motoneurones were also observed in Drosophila mutants that have elevated cyclic AMP concentrations (Zhong et al. 1992). In these mutants, as in the hyperexcitable mutants, the numbers of terminal varicosities and higher-order motoneurone branches are increased. In double mutants that are hyperexcitable and have elevated cyclic AMP concentrations, these changes in axonal morphology are more pronounced than in flies with a single mutation (Zhong et al. 1992). These observations led to the hypothesis that enhanced neural activity in the hyperexcitable mutants increases calcium influx, stimulating the synthesis of cyclic AMP via calcium/calmodulin activation of adenylate cyclase. Cyclic AMP has been shown to affect the growth of neurites in culture (Lohof et al. 1992, and references therein).

Leech neurones

During the development of the leech nervous system, several identified neurones exhibit an initial overproduction, followed by a selective loss, of neurites (Kuwada and Kramer, 1983; Wallace, 1984; Gao and Macagno, 1987a,b; Wolszon and Macagno, 1992). The retraction of the neurites in developing (and regenerating) leech neurones follows one of two patterns. The first is a global retraction, whereby the growth cone that reaches its target seems to send a retrograde message causing the neurone to retract other projections (Jellies et al. 1987; Loer et al. 1987; Baptista and Macagno, 1988). The second type of retraction is localized so that the growth cone that reaches its target collapses and retracts, while the other neurites of the same cell are not affected (Gao and Macagno, 1987a,b). Recent observations on the local growth cone response of a neurone known as the anterior pagoda (AP) suggest that activity might play a role in neurite retraction. As AP neurones develop (there are two in each ganglion), they project two axons to the periphery and one axon each into the anterior and posterior connectives. The axons projecting into the connectives grow until they contact the growth cone of the AP cell from the adjacent ganglion; they then stop growing and retract. If one of the AP cells is ablated, the longitudinal projection of its neighbour is retained (Gao and Macagno, 1987b). Morphological and physiological evidence shows that the two axons are coupled by gap junctions through which 5-hydroxytryptamine (5-HT) might be transported between AP neurones of neighbouring ganglia (Wolszon et al. 1994). Gap junctions have also been implicated in the development of the vertebrate nervous system (Katz, 1993; Peinado et al. 1993).

In all the examples mentioned, neurite retraction is limited to a selected set of processes, while other neurites of the same neurone continue to grow. Why then are not all processes affected by depolarization and what is the nature of the mediator of electrical activity on growth cone motility?

Effect of electrical activity on Purkinje cell development in culture

The Purkinje cell is one of the few neurones that has been analyzed both in vivo and in vitro with respect to the effect of electrical activity on its growth pattern. Electrophysiological, biochemical and morphological changes in cultured Purkinje cells mirror the in vivo developmental programme (Schilling et al. 1991). The emergence of electrical activity and dendritic branching coincide. Blockage of action potentials with TTX results in a dramatic change in dendritic morphology. Whereas control cells have a dense tree of branched dendrites with numerous protuberances (probably spines), silenced Purkinje cells have dendrites that are 3–4 times longer and exhibit little branching and no spines (Schilling et al. 1991). Axon morphology is not affected by changes in electrical activity. As electrical activity and dendritic branching appear, intracellular calcium levels become sensitive to changes in membrane potential. Calcium conductances in Purkinje cells are prominent in the dendritic tree and reduced in the axon (Llinas and Sugimori, 1979, 1980a,b; Tank et al. 1988; Ross et al. 1990). These observations suggest that calcium may be involved by acting as a second messenger that couples membrane events to changes in Purkinje cell dendritic development.

The role of calcium

Calcium currents have been localized on growing tips of regenerating neurites in vivo (Meiri et al. 1981) and growth cones of cultured neurones (Grinvald and Farber, 1981; Anglister et al. 1982); moreover, calcium channels are the first of the voltage-sensitive channels to be detected during the development of a variety of neurones (Spitzer and Baccaglini, 1976; Spitzer, 1979). The role calcium influx may play in activity-induced changes in neurite outgrowth has been investigated in several types of neurones grown in culture.

Depolarization-induced neurite retraction is mediated by a rise in intracellular calcium concentration

A phasic pattern of stimulation of dorsal root ganglion (DRG) neurones causes immediate growth cone collapse with retraction of filopodia and lamellipodia. This response is often followed by retraction of the neurite. Growth cone morphology and growth rate do not change, however, when the action potentials are blocked with TTX (Fields et al. 1990). Depolarization of DRG neurones with either a high extracellular potassium concentration (45 mmol l−1) or bradykinin also inhibits neurite outgrowth (Robson and Burgoyne, 1989). Electrical stimulation or depolarization with a high extracellular potassium concentration results in an elevation of intracellular calcium concentration. This increase in calcium concentration is due to a transmembrane flux of calcium ions through voltage-dependent L-type calcium channels. Blocking the calcium influx abolishes the growth-inhibitory effect of depolarization (Robson and Burgoyne, 1989; Fields et al. 1993). These experiments show that the inhibition of neurite outgrowth by depolarization is mediated by a rise in intracellular calcium concentration. DRG growth cones also collapse after exposure to a myelin-associated protein (NI-35). Interestingly, this response is due to release of calcium from intracellular stores (Bandtlow et al. 1993).

Environment and synapse formation affect calcium channel expression

The effect of environmental influences on the response of neurites to electrical activity was analyzed in identified cultured leech neurones. Neurite morphology is strongly influenced by the substrata on which leech neurones are plated. When grown on laminin, neurites are long and thin and rarely branch, whereas on the plant lectin Concanavalin A, they are broad, curly and have a complex branching pattern (Chiquet and Acklin, 1986; Chiquet and Nicholls, 1987; Grumbacher-Reinert, 1989). Leech neurones growing on laminin respond with growth cone collapse and neurite retraction to electrical stimulation or elevated extracellular potassium, a response that is dependent on calcium influx (Fig. 2). The same cells cultured on Concanavalin A continue to grow when they are stimulated or depolarized (Grumbacher-Reinert and Nicholls, 1992; Neely, 1993). Ross et al. (1988) have shown that voltage-dependent calcium channels are abundant in leech neurites growing on laminin, but scarce in processes of neurones growing on Concanavalin A. The hypothesis that these differences in calcium channels might be responsible for differences in response to depolarization is suggested by the finding that treatment of leech neurones growing on Concanavalin A with the calcium ionophore A23187 results in cessation of outgrowth (Neely and Gesemann, 1994). The environment with which the growth cone is in contact therefore influences the expression of calcium channels and the response of a neurite to neural activity.

Fig. 2.

Growth cones of leech neurones cultured on a laminin-enriched substratum collapse and neurites retract after depolarization. An anterior pagoda (AP) neurone in culture was photographed before (A) and after (B) 30 min of exposure to medium with elevated [K+] (60 mmol l−1) (arrow points to cell body). This depolarization caused growth cone collapse (box E) and neurite retraction (box D). Scanning electron microscopy revealed retraction bulbs (arrowheads) and retraction fibres (arrows) on neurites that had retracted (D). Loss of filopodia and rounding up of lamellipodia are characteristic features of collapsed growth cones (E) (compare E with C, which shows a control growth cone). D and E are scanning electron micrographs of the boxed regions in B. The cell body in A and B has a diameter of 50 μm. (reproduced with permission, after Neely, 1993).

Fig. 2.

Growth cones of leech neurones cultured on a laminin-enriched substratum collapse and neurites retract after depolarization. An anterior pagoda (AP) neurone in culture was photographed before (A) and after (B) 30 min of exposure to medium with elevated [K+] (60 mmol l−1) (arrow points to cell body). This depolarization caused growth cone collapse (box E) and neurite retraction (box D). Scanning electron microscopy revealed retraction bulbs (arrowheads) and retraction fibres (arrows) on neurites that had retracted (D). Loss of filopodia and rounding up of lamellipodia are characteristic features of collapsed growth cones (E) (compare E with C, which shows a control growth cone). D and E are scanning electron micrographs of the boxed regions in B. The cell body in A and B has a diameter of 50 μm. (reproduced with permission, after Neely, 1993).

The establishment of contacts between leech neurites prevents their retraction, while the free neurites on the same cell shorten following electrical stimulation (R. von Bernhardi, unpublished observations). Synapse formation has been shown to change calcium channel distribution in cultured leech neurones. After the formation of chemical synapses between two identified leech neurones, calcium currents at the cell stump (region of outgrowth) in the postsynaptic cell are reduced (Fernandez-de Miguel et al. 1992). This change is accompanied by reduced outgrowth in the postsynaptic cell when compared with single neurones (Cooper et al. 1992). Synapse formation can therefore alter calcium channel activity and neurite outgrowth.

Variety of growth cone responses to calcium

Whereas a rise in intracellular calcium concentration inhibits growth cone motility in leech neurones (Grumbacher-Reinert and Nicholls, 1992; Neely, 1993) and DRG neurones (Robson and Burgoyne, 1989; Fields et al. 1993), stimulation of growth occurs in neuroblastoma cells (Anglister et al. 1982; Bedlack et al. 1992), ciliary ganglion neurones (Nishi and Berg, 1981), neurones from the diencephalon (Connor, 1986), embryonic spinal neurones (Holliday and Spitzer, 1990) and embryonic retinal neurones (Suarez-Isla et al. 1984). Neurites of Helisoma trivolvis neurones stop growing when the intracellular calcium concentration is raised or reduced, suggesting that there is perhaps an optimal range of calcium concentration, above or below which neurite outgrowth is impaired (Cohan and Kater, 1986; Cohan et al. 1987; Mattson and Kater, 1987; Kater et al. 1988; Cohan, 1992).

In cultured sympathetic neurones, electrical stimulation increases the cytosolic calcium concentration, but has no effect on the rate of neurite outgrowth (Garyantes and Regehr, 1992). Voltage-sensitive calcium channels also do not seem to play a critical role for neuronal differentiation of PC12 cells. PC12 cells of a line devoid of such channels grow normal neurites after they are induced to differentiate (Usowicz et al. 1990).

We conclude that electrical activity and the concomitant calcium influx are not universal regulators of neurite outgrowth. Inherent properties of neurones, environmental influences (such as the substratum) and synapse formation, can all affect calcium currents and modify the response of a neurone to electrical activity.

Cytoskeletal organization in the growth cone

Changes in growth cone morphology and motility are produced by changes in the organization of its cytoskeleton, which is composed of microtubules and microfilaments (Yamada et al. 1970, 1971; Luduena and Wessells, 1973; Letourneau et al. 1986; Forscher and Smith, 1988). The microtubules extend from the neurite shaft and splay into the centre part of the growth cone. They have ocasionally been observed to extend to peripheral parts of the growth cone or to enter filopodia. The actin cytoskeleton is composed of two distinct microfilament subpopulations: one forms a meshwork of relatively short randomly oriented filaments in the lamellipodia, the second is composed of parallel bundles of up to a dozen filaments that radiate from the leading edge of the lamellipodia and extend into the filopodia (Yamada et al. 1970, 1971; Luduena and Wessells, 1973; Letourneau and Ressler, 1983; Forscher and Smith, 1988; Lewis and Bridgman, 1992). Disruption of this cytoskeletal organization with cytoskeletal drugs such as cytochalasin, nocodazole and taxol results in alterations of growth cone morphology and motility (Yamada et al. 1970, 1971; Letourneau and Ressler, 1984; Marsh and Letourneau, 1984; Letourneau et al. 1986; Forscher and Smith, 1988; Neely and Gesemann, 1994). In particular, disruption of the microfilaments by cytochalasin results in loss of filopodia, retraction of lamellipodia and cessation of growth cone motility (Yamada et al. 1970, 1971; Marsh and Letourneau, 1984; Bentley and Toroian-Raymond, 1986; Neely and Gesemann, 1994), the same response that is observed after electrical stimulation and/or elevation of intracellular calcium concentration (Cohan and Kater, 1986; Fields et al. 1990; Grumbacher-Reinert and Nicholls, 1992; Neely, 1993).

Cytoskeletal organization in growth cones after calcium influx

Growth cones of leech neurones collapse after an increase in intracellular calcium concentration due to activity, depolarization or treatment with a calcium ionophore (Grumbacher-Reinert and Nicholls, 1992; Neely, 1993; Neely and Gesemann, 1994). These changes are accompanied by a loss of microfilaments from the peripheral areas of growth cones, while the organization of microtubules is unaffected (Fig. 3). Pre-incubation of the neurones with phalloidin, a microfilament stabilizing agent, inhibits calcium-induced growth cone collapse (Neely and Gesemann, 1994). In DRG growth cones, elevated intracellular calcium concentrations cause disruption of microfilaments and destabilization of microtubules (Lankford and Letourneau, 1989). Interestingly, the loss of microfilaments and collapse of DRG growth cones caused by a membrane-associated collapsing factor is not associated with an increase in cytosolic calcium concentration (Fan et al. 1993).

Fig. 3.

Raised intracellular calcium concentration leads to disruption of microfilaments in leech neurones. The cytoskeleton of leech neurones growing on the plant lectin Concanavalin A was analyzed by immunofluorescence. In control cells, microfilaments (stained with Rhodamine phalloidin) are abundant in the peripheral areas of the neurites (A), whereas the microtubules (stained with an anti-tubulin antibody) are concentrated in the central parts of the growth cone, with some microtubules extending to more peripheral regions (B). Exposure of leech neurones to the calcium ionophore A23187 (which increases the intracellular calcium concentration) leads to a dramatic loss of microfilaments in the neurites (C). The distribution of the microtubules, however, is not changed by this treatment (D). Scale bars: A, C, 50 μm; B, D, 20 μm (reproduced with permission, after Neely and Gesemann, 1994).

Fig. 3.

Raised intracellular calcium concentration leads to disruption of microfilaments in leech neurones. The cytoskeleton of leech neurones growing on the plant lectin Concanavalin A was analyzed by immunofluorescence. In control cells, microfilaments (stained with Rhodamine phalloidin) are abundant in the peripheral areas of the neurites (A), whereas the microtubules (stained with an anti-tubulin antibody) are concentrated in the central parts of the growth cone, with some microtubules extending to more peripheral regions (B). Exposure of leech neurones to the calcium ionophore A23187 (which increases the intracellular calcium concentration) leads to a dramatic loss of microfilaments in the neurites (C). The distribution of the microtubules, however, is not changed by this treatment (D). Scale bars: A, C, 50 μm; B, D, 20 μm (reproduced with permission, after Neely and Gesemann, 1994).

A hypothesis for the mechanism of calcium-induced microfilament disruption

Intracellular calcium concentrations after depolarization

How could calcium cause the disruption of microfilaments? Intracellular calcium concentrations in growth cones are in the range 100–250 nmol l−1, reaching 200–500 nmol l−1 after depolarization (Connor, 1986; Cohan et al. 1987; Lipscombe et al. 1988; Ishida et al. 1991; Garyantes and Regehr, 1992), and are therefore lower than the micromolar concentration required to activate proteins controlling cell functions (Baker and Knight, 1981; Manalan and Klee, 1984). Some studies using calcium-sensitive dyes, however, have shown higher intracellular calcium concentrations in spatially restricted areas (‘calcium hotspots’) of depolarized growth cones (Bolsover and Spector, 1986; Kater et al. 1988; Smith and Augustine, 1988; Tank et al. 1988; Regehr et al. 1989). Patch-clamp, toxin-binding and freeze-fracture studies demonstrate that calcium channels are grouped in clusters on growth cones (Pumplin et al. 1981; Fox et al. 1987; Lipscombe et al. 1988; Jones et al. 1989). Silver et al. (1990) have shown that calcium channel clustering can cause such ‘calcium hotspots’ with an average diameter of 7 μm and a calcium concentration of up to 1 μmol l−1. Growth cones contain between 1–5 such ‘hotspots’ which are often located at the base of filopodia and coincide with the site of greatest outgrowth of neuroblastoma growth cone margins. These studies therefore suggest that local intracellular calcium concentrations may be high enough to activate many calcium-dependent proteins.

The structure of the microfilaments is regulated by the microfilament-associated proteins. Calcium could affect the organization of the microfilaments through modification of these proteins by calcium-dependent second-messenger cascades (Nairn et al. 1985; Aderem, 1992; Schulman et al. 1992). Alternatively, calcium might act directly on microfilament-associated proteins whose activity is calcium-dependent (Stossel et al. 1985; Vandekerckhove, 1990; Weeds and Maciver, 1993). Here we focus on the role of calcium-binding microfilament-associated proteins.

Is gelsolin a mediator of calcium-induced microfilament disruption?

Two types of proteins could be involved in the calcium-induced loss of peripheral microfilaments and growth cone collapse: actin filament associated motors (such as myosin) or filament severing proteins (such as gelsolin) (Sobue, 1993; Weeds and Maciver, 1993; Cramer et al. 1994). The disappearance of the microfilaments in the peripheral areas of the growth cone is not due to their contraction, but to an actual loss (Neely and Gesemann, 1994). This suggests the involvement of a microfilament-severing protein rather than a microfilament-associated motor protein.

Gelsolin is a protein that severs actin filaments when it is activated by calcium. Biochemical studies have shown that gelsolin remains tightly bound to the fragments after cleaving, preventing repolymerization (Yin et al. 1980, 1981b). Subsequent binding of gelsolin to polyphosphoinositides (phosphatidylinositol 4-monophosphate and phosphatidylinositol 4,5-bisphosphate) results in dissociation of the fragments from gelsolin, and the freed filament fragments then act as nuclei for rapid actin reassembly (Janmey and Stossel, 1987, 1989; Janmey et al. 1987).

Ultrastructural analysis of the distribution of gelsolin has shown that it is mostly in the cytoplasm in resting macrophages and platelets. Upon activation of these cells, gelsolin is redistributed to the plasma membrane associated with the ends of short actin filaments (Hartwig et al. 1989). Presumably binding of gelsolin to phosphatidylinositol phosphates in the membrane causes the release of the filament fragments and a burst of actin polymerization at the membrane–cytoskeleton interface.

The role of gelsolin in cell motility has been analyzed in living fibroblasts and macrophages. Increasing the content of gelsolin in cultured fibroblasts by gene transfection enhances their motility (Cunningham et al. 1991). Injection of a constitutively severing form of gelsolin into fibroblasts or macrophages induces loss of stress fibres, inhibits membrane ruffling and causes these cells to become rounded (Cooper et al. 1987).

Gelsolin is expressed in many different tissues including the nervous system (Yin et al. 1981a). In the rat central and peripheral nervous systems, it is enriched in oligodendrocytes and Schwann cells (Tanaka and Sobue, 1994). In the rat retina, gelsolin occurs in glial cells and neurones, the photoreceptors and ganglion cells (Legrand et al. 1991). In culture, gelsolin occurs in sympathetic neurones and co-localizes with the microfilaments of the peripheral areas of the growth cone and the filopodia in differentiated PC12 cells, DRG neurones (Tanaka et al. 1993) and leech neurones (M. D. Neely and E. Macaluso, unpublished observation). The high concentration of gelsolin in the filopodia of leech neurones is interesting in correlation with a recent publication reporting that isolated filopodia respond to membrane depolarization with an increase in intracellular calcium concentration and a decrease in length (Davenport et al. 1993). Gelsolin’s severing activity, intracellular distribution, effect on cell motility and abundance in neuronal growth cones suggest it as a possible effector of the calcium-induced disruption of microfilaments.

In this review, we have compared the effect of neural activity on neurones in vivo and in culture. In neurones in culture, changes in growth cone motility after electrical stimulation are accompanied by an influx of calcium through voltage-sensitive calcium channels. The effects of electrical activity and increases in intracellular calcium concentration on growth cone morphology are not the same for all neurones. Some growth cones collapse, some show greater motility and others do not respond at all, depending on the type of neurone, the type of neurite (axon versus dendrite) and environmental factors. These findings suggest that there are several different signals and signal transduction mechanisms that ultimately result in alterations of the cytoskeletal structure and growth cone motility. The nature of these signals and signal transduction pathways, and the mechanism by which the growth cone integrates all the information with which it is presented, remain to be elucidated. Gelsolin is one candidate that could couple altered levels of electrical activity to reorganizations of the cytoskeleton and therefore to changes in growth cone motility.

Aderem
,
A.
(
1992
).
Signal transduction and the actin cytoskeleton: The roles of MARCKS and profilin
.
Trends biochem. Sci.
17
,
438
443
.
Anderson
,
P. A.
,
Olavarria
,
J.
and
Van Sluyters
,
R. C.
(
1988
).
The overall pattern of ocular dominance bands in cat visual cortex
.
J. Neurosci.
8
,
2183
2200
.
Anglister
,
L.
,
Farber
,
I. C.
,
Shahar
,
A.
and
Grinvald
,
A.
(
1982
).
Localization of voltage-sensitive calcium channels along developing neurites: Their possible role in regulating neurite elongation
.
Devl Biol.
94
,
351
365
.
Antonini
,
A.
and
Stryker
,
M. P.
(
1993
).
Development of individual geniculocortical arbors in cat striate cortex and effects of binocular impulse blockade
.
J. Neurosci.
13
,
3549
3573
.
Baker
,
P. F
and
Knight
,
D. E.
(
1981
).
Calcium control of exocytosis and endocytosis in bovine adrenal medullary cells
.
Phil. Trans. R. Soc. Lond. B
296
,
83
103
.
Bandtlow
,
C. E.
,
Schmidt
,
M. F.
,
Hassinger
,
T. D.
,
Schwab
,
M. E.
and
Kater
,
S. B.
(
1993
).
Role of intracellular calcium in NI-35-evoked collapse of neuronal growth cones
.
Science
259
,
80
83
.
Baptista
,
C. A.
and
Macagno
,
E. R.
(
1988
).
Modulation of the pattern of axonal projections of a leech motor neuron by ablation or transplantation of its target
.
Neuron
1
,
949
962
.
Baumann
,
A.
,
Krah-Jentgens
,
I.
,
Müller
,
R.
,
Müller-Holtkamp
,
F.
,
Seidel
,
R.
,
Kecskemethy
,
N.
,
Casal
,
J.
,
Ferrus
,
A.
and
Pongs
,
O.
(
1987
).
Molecular organization of the maternal effect region of the Shaker complex of Drosophila: Characterization of an IA channel transcript with homology to vertebrate Na+ channel
.
EMBO J.
6
,
3419
3429
.
Bedlack
,
R. S.
, Jr
,
Wei
,
M.-D.
and
Loew
,
L. M.
(
1992
).
Localized membrane depolarizations and localized calcium influx during electric field-guided neurite growth
.
Neuron
9
,
393
403
.
Benoit
,
P.
and
Changeux
,
J.-P.
(
1975
).
Consequences of tenotomy on the evolution of multiinnervation in developing rat soleus muscle
.
Brain Res.
99
,
354
358
.
Bentley
,
D.
and
Toroian-Raymond
,
A.
(
1986
).
Disoriented pathfinding by pioneer neurone growth cones deprived of filopodia by cytochalasin treatment
.
Nature
323
,
712
715
.
Berman
,
N. E.
and
Payne
,
B. R.
(
1983
).
Alterations in connections of the corpus callosum following convergent and divergent strabismus
.
Brain Res.
274
,
201
212
.
Berry
,
M.
and
Bradley
,
P.
(
1976
).
The growth of the dendritic trees of Purkinje cells in irradiated agranular cerebellar cortex
.
Brain Res.
116
,
361
387
.
Bolsover
,
S. R.
and
Spector
,
I.
(
1986
).
Measurements of calcium transients in the soma, neurite and growth cone of single cultured neurons
.
J. Neurosci.
6
,
1934
1940
.
Bowling
,
D. B.
and
Michael
,
C. R.
(
1980
).
Projection patterns of single physiologically characterized optic tract fibers in cat
.
Nature
286
,
899
902
.
Brown
,
M. C.
,
Jansen
,
J. K. S.
and
Van Essen
,
D.
(
1976
).
Polyneuronal innervation of skeletal muscle in new-born rats and its elimination during maturation
.
J. Physiol., Lond.
261
,
387
422
.
Budnik
,
V.
,
Zhong
,
Y.
and
Wu
,
C.-F.
(
1990
).
Morphological plasticity of motor axons in Drosophila mutants with altered excitability
.
J. Neurosci.
10
,
3754
3768
.
Burg
,
M. G.
and
Wu
,
C.-F.
(
1986
).
Differentiation and central projections of peripheral sensory cells with action-potential block in Drosophila mosaics
.
J. Neurosci.
6
,
2968
2976
.
Burg
,
M. G.
and
Wu
,
C.-F.
(
1989
).
Central projections of peripheral mechanosensory cells with increased excitability in Drosophila mosaics
.
Devl Biol.
131
,
505
514
.
Callaway
,
E. M.
and
Katz
,
L. C.
(
1990
).
Emergence and refinement of clustered horizontal connections in cat striate cortex
.
J. Neurosci.
10
,
1134
1153
.
Callaway
,
E. M.
and
Katz
,
L. C.
(
1991
).
Effects of binocular deprivation on the development of clustered horizontal connections in cat striate cortex
.
Proc. natn. Acad. Sci. U.S.A.
88
,
745
749
.
Chiquet
,
M.
and
Acklin
,
S. E.
(
1986
).
Attachment of Con A or extracellular matrix initiates rapid sprouting by cultured leech neurons
.
Proc. natn. Acad. Sci. U.S.A.
83
,
6188
6192
.
Chiquet
,
M.
and
Nicholls
,
J. G.
(
1987
).
Neurite outgrowth and synapse formation by identified leech neurones in culture
.
J. exp. Biol.
132
,
191
206
.
Cohan
,
C. S.
(
1992
).
Depolarization-induced changes in neurite elongation and intracellular Ca2+ in isolated Helisoma neurons
.
J. Neurobiol.
23
,
983
996
.
Cohan
,
C. S.
,
Connor
,
J. A.
and
Kater
,
S. B.
(
1987
).
Electrically and chemically mediated increases in intracellular calcium in neuronal growth cones
.
J. Neurosci.
7
,
3588
3599
.
Cohan
,
C. S.
and
Kater
,
S. B.
(
1986
).
Suppression of neurite elongation and growth cone motility by electrical activity
.
Science
232
,
1638
1640
.
Connor
,
J. A.
(
1986
).
Digital imaging of free calcium changes and of spatial gradients in growing processes in single, mammalian central nervous system cells
.
Proc. natn. Acad. Sci. U.S.A.
83
,
6179
6183
.
Cooper
,
J. A.
,
Bryan
,
J.
,
Schwab Iii
,
B.
,
Frieden
,
C.
and
Loftus
,
D. J.
(
1987
).
Microinjection of gelsolin into living cells
.
J. Cell Biol.
104
,
491
501
.
Cooper
,
R. L.
,
Fernandez-De-Miguel
,
F.
and
Adams
,
W. B.
(
1992
).
Anterograde and retrograde effects of synapse formation on calcium currents and neurite outgrowth in cultured leech neurons
.
Proc. R. Soc. Lond. B
249
,
217
222
.
Cowan
,
W. M.
,
Fawcett
,
J. W.
,
O’leary
,
D. D. M.
and
Stanfield
,
B. B.
(
1984
).
Regressive events in neurogenesis
.
Science
255
,
1258
1265
.
Cramer
,
L. P.
,
Mitchison
,
T. J.
and
Theriot
,
J. A.
(
1994
).
Actin-dependent motile forces and cell motility
.
Curr. Opin. Cell Biol.
6
,
82
86
.
Crepel
,
F.
,
Mariani
,
J.
and
Delhaye-Bouchaud
,
N.
(
1976
).
Evidence for a multiple innervation of Purkinje cells by climbing fibers in the immature rat cerebellum
.
J. Neurobiol.
7
,
567
578
.
Cunningham
,
C. C.
,
Stossel
,
T. P.
and
Kwiatkowski
,
D. J.
(
1991
).
Enhanced motility in NIH 3T3 fibroblasts that overexpress gelsolin
.
Science
251
,
1233
1236
.
Cusick
,
C. G.
and
Lund
,
R. D.
(
1982
).
Modification of visual callosal projections in rats
.
J. comp. Neurol.
212
,
385
398
.
Dalva
,
M. B.
,
Gosh
,
A.
and
Shatz
,
C. J.
(
1994
).
Independent control of dendritic and axonal form in the developing lateral geniculate nucleus
.
J. Neurosci.
14
,
3588
3602
.
Dan
,
Y.
and
Poo
,
M.-M.
(
1992
).
Hebbian depression of isolated neuromuscular synapses in vitro
.
Science
256
,
1570
1573
.
Dan
,
Y.
and
Poo
,
M.-M.
(
1994a
).
Retrograde interactions during formation and elimination of neuromuscular synapses
.
Curr. Opin. Neurobiol.
4
,
95
100
.
Dan
,
Y.
and
Poo
,
M.-M.
(
1994b
).
Calcium-dependent postsynaptic exocytosis: A possible mechanism for activity-dependent synaptic modulation
.
J. Neurobiol.
25
,
336
341
.
Dann
,
J. F.
,
Buhl
,
E. H.
and
Peichl
,
L.
(
1988
).
Postnatal dendritic maturation of alpha and beta ganglion cells in cat retina
.
J. Neurosci.
8
,
1485
1499
.
Davenport
,
R. W.
,
Dou
,
P.
,
Rehder
,
V.
and
Kater
,
S. B.
(
1993
).
A sensory role for neuronal growth cone filopodia
.
Nature
361
,
721
724
.
Delhaye-Bouchaud
,
N.
,
Crepel
,
F.
and
Mariani
,
J.
(
1975
).
Mise en evidence d’une multi-innervation temporaire des cellules Purkinje du cervelet par les fibres grimpantes au cours du development chez le rat
.
C.R. hébd. Séanc. Acad. Sci. Paris
281
,
909
912
.
Fan
,
J.
,
Mansfield
,
S. G.
,
Redmond
,
T.
,
Gordon-Weeks
,
P. R.
and
Raper
,
J. A.
(
1993
).
The organization of F-actin and microtubules in growth cones exposed to a brain-derived collapsing factor
.
J. Cell Biol.
121
,
867
878
.
Fernandez-De-Miguel
,
F.
,
Cooper
,
R. L.
and
Adams
,
W. B.
(
1992
).
Synaptogenesis and calcium current distribution in cultured leech neurons
.
Proc. R. Soc. Lond. B
247
,
215
221
.
Ferster
,
D.
and
Levay
,
S.
(
1978
).
The axonal arborizations of lateral geniculate neurons in the striate cortex of the cat
.
J. comp. Neurol.
182
,
923
944
.
Fields
,
R. D.
,
Guthrie
,
P. B.
,
Russell
,
J. T.
,
Kater
,
S. B.
,
Malhotra
,
B. S.
and
Nelson
,
P. G.
(
1993
).
Accommodation of mouse DRG growth cones to electrically induced collapse: Kinetic analysis of calcium transients and set-point theory
.
J. Neurobiol.
24
,
1080
1098
.
Fields
,
R. D.
,
Neale
,
E. A.
and
Nelson
,
P. G.
(
1990
).
Effects of patterned electrical activity on neurite outgrowth from mouse sensory neurons
.
J. Neurosci.
10
,
2950
2964
.
Fields
,
R. D.
and
Nelson
,
P. G.
(
1992
).
Activity-dependent development of the vertebrate nervous system
.
Int. Rev. Neurobiol.
34
,
133
214
.
Forscher
,
P.
and
Smith
,
S. J.
(
1988
).
Actions of cytochalasins on the organization of actin filaments and microtubules in a neuronal growth cone
.
J. Cell Biol.
107
,
1505
1516
.
Fox
,
A. P.
,
Nowycky
,
M. C.
and
Tsien
,
R. W.
(
1987
).
Single-channel recordings of three types of calcium channels in chick sensory neurones
.
J. Physiol., Lond.
394
,
173
200
.
Frost
,
D. O.
and
Moy
,
Y. P.
(
1989
).
Effects of dark rearing on the development of visual callosal connections
.
Expl Brain Res.
78
,
203
213
.
Frost
,
D. O.
,
Moy
,
Y. P.
and
Smith
,
D. C.
(
1990
).
Effects of alternating monocular occlusion on the development of the visual callosal connections
.
Expl Brain Res.
83
,
200
209
.
Ganetzky
,
B.
and
Wu
,
C.-F.
(
1986
).
Neurogenetics of membrane excitability in Drosophila
.
A. Rev. Genet.
20
,
13
44
.
Gao
,
W.-Q.
and
Macagno
,
E. R.
(
1987a
).
Extension and retraction of axonal projections by some developing neurons in the leech depends upon the existence of neighboring homologues. I. The HA cells
.
J. Neurobiol.
18
,
43
59
.
Gao
,
W.-Q.
and
Macagno
,
E. R.
(
1987b
).
Extension and retraction of axonal projections by some developing neurons in the leech depends upon the existence of neighboring homologues. II. The AP and AE neurons
.
J. Neurobiol.
18
,
295
313
.
Garyantes
,
T. K.
and
Regehr
,
W. G.
(
1992
).
Electrical activity increases growth cone calcium but fails to inhibit neurite outgrowth from rat sympathetic neurons
.
J. Neurosci.
12
,
96
103
.
Gilbert
,
C. D.
and
Wiesel
,
T. N.
(
1989
).
Columnar specificity of intrinsic horizontal and corticocortical connections in cat visual cortex
.
J. Neurosci.
9
,
2432
2442
.
Goodman
,
C. S.
and
Shatz
,
C. J.
(
1993
).
Developmental mechanisms that generate precise patterns of neuronal connectivity
.
Cell 72/Neuron
10
(
Suppl
.),
77
98
.
Grigonis
,
A. M.
and
Murphy
,
E. H.
(
1994
).
The effects of epileptic cortical activity on the development of callosal projections
.
Devl Brain Res.
77
,
251
255
.
Grinvald
,
A.
and
Farber
,
I. C.
(
1981
).
Optical recording of calcium action potentials from growth cones of cultured neurons with a laser microbeam
.
Science
212
,
1164
1167
.
Grumbacher-Reinert
,
S.
(
1989
).
Local influence of substrate molecules in determining distinctive growth patterns of identified neurons in culture
.
Proc. natn. Acad. Sci. U.S.A.
86
,
7270
7274
.
Grumbacher-Reinert
,
S.
and
Nicholls
,
J.
(
1992
).
Influence of substrate on retraction of neurites following electrical activity of leech Retzius cells in culture
.
J. exp. Biol.
167
,
1
14
.
Hartwig
,
J. H.
,
Chambers
,
K. A.
and
Stossel
,
T. P.
(
1989
).
Association of gelsolin with actin filaments and cell membranes of macrophages and platelets
.
J. Cell Biol.
108
,
467
479
.
Hayhow
,
W. R.
(
1958
).
The cytoarchitecture of the lateral geniculate body in the cat in relation to the distribution of crossed and uncrossed optic fibers
.
J. comp. Neurol.
110
,
1
64
.
Hickey
,
T. L.
and
Guillery
,
R. W.
(
1974
).
An autoradiographic study of retinogeniculate pathways in the cat and the fox
.
J. comp. Neurol.
156
,
239
254
.
Holliday
,
J.
and
Spitzer
,
N. C.
(
1990
).
Spontaneous calcium influx and its roles in differentiation of spinal neurons in culture
.
Devl Biol.
141
,
13
23
.
Hubel
,
D. H.
and
Wiesel
,
T. N.
(
1961
).
Integrative action in the cat’s lateral geniculate body
.
J. Physiol., Lond.
155
,
385
398
.
Hubel
,
D. H.
and
Wiesel
,
T. N.
(
1972
).
Laminar and columnar distribution of geniculo-cortical fibers in the macaque monkey
.
J. comp. Neurol.
146
,
421
450
.
Humphrey
,
A. L.
,
Sur
,
M.
,
Uhlrich
,
D. J.
and
Sherman
,
S. M.
(
1985a
).
Projection patterns of individual X- and Y-cell axons from the lateral geniculate nucleus to cortical area 17 in the cat
.
J. comp. Neurol.
233
,
159
189
.
Humphrey
,
A. L.
,
Sur
,
M.
,
Ulrich
,
D. J.
and
Sherman
,
S. M.
(
1985b
).
Termination patterns of individual X- and Y-cell axons in the visual cortex of the cat: Projections to area 18, to the 17/18 border region and to both areas 17 and 18
.
J. comp. Neurol.
233
,
190
212
.
Innocenti
,
G. M.
(
1981
).
Growth and reshaping of axons in the establishment of visual callosal connections
.
Science
212
,
824
827
.
Innocenti
,
G. M.
and
Frost
,
D. O.
(
1979
).
Effects of visual experience on the maturation of the efferent system to the corpus callosum
.
Nature
280
,
231
234
.
Innocenti
,
G. M.
and
Frost
,
D. O.
(
1980
).
The postnatal development of visual callosal connections in the absence of visual experience or of the eyes
.
Expl Brain Res.
39
,
365
375
.
Innocenti
,
G. M.
,
Frost
,
D. O.
and
Illes
,
J.
(
1985
).
Maturation of visual callosal connections in visually deprived kittens: A challenging critical period
.
J. Neurosci.
5
,
255
267
.
Ishida
,
A. T.
,
Bindokas
,
V. P.
and
Nuccitelli
,
R.
(
1991
).
Calcium ion levels in resting and depolarized goldfish retinal ganglion cell somata and growth cones
.
J. Neurophysiol.
65
,
968
979
.
Ito
,
M.
(
1984
).
Purkinje cells: Abnormal morphology
. In
The Cerebellum and Neural Control
(ed.
M.
Ito
), pp.
40
50
.
New York
:
Raven Press
.
Ivy
,
G. O.
and
Killackey
,
H. P.
(
1982
).
Ontogenetic changes in the projections of neocortical neurons
.
J. Neurosci.
2
,
735
743
.
Janmey
,
P. A.
,
Iida
,
K.
,
Yin
,
H. L.
and
Stossel
,
T. P.
(
1987
).
Polyphosphoinositide micelles and polyphosphoinositide-containing vesicles dissociate endogenous gelsolin–actin complexes and promote actin assembly from the fast-growing end of actin filaments blocked by gelsolin
.
J. biol. Chem.
262
,
12228
12236
.
Janmey
,
P. A.
and
Stossel
,
T. P.
(
1987
).
Modulation of gelsolin function by phosphatidylinositol 4,5-bisphosphate
.
Nature
325
,
362
364
.
Janmey
,
P. A.
and
Stossel
,
T. P.
(
1989
).
Gelsolin–polyphosphoinositide interaction
.
J. biol. Chem.
264
,
4825
4831
.
Jellies
,
J.
,
Loer
,
C. M.
and
Kristan
,
W. B.
Jr
,
(
1987
).
Morphological changes in leech Retzius neurons after target contact during embryogenesis
.
J. Neurosci.
7
,
2618
2629
.
Jones
,
O. T.
,
Kunze
,
D. L.
and
Angelides
,
K. J.
(
1989
).
Localization and mobility of v-conotoxin-sensitive Ca2+ channels in hippocampal CA1 neurons
.
Science
244
,
1189
1193
.
Kamb
,
A.
,
Iverson
,
L. E.
and
Tanouye
,
M. A.
(
1987
).
Molecular characterization of Shaker, a Drosophila gene that encodes a potassium channel
.
Cell
50
,
405
413
.
Kater
,
S. B.
,
Mattson
,
M. P.
,
Cohan
,
C.
and
Connor
,
J.
(
1988
).
Calcium regulation of the neuronal growth cone
.
Trends Neurosci.
11
,
315
321
.
Katz
,
L. C.
(
1993
).
Coordinate activity in retinal and cortical development
.
Curr. Opin. Neurobiol.
3
,
93
99
.
Krasne
,
F. B.
(
1987
).
Silencing normal input permits regenerating foreign afferents to innervate an identified crayfish sensory interneuron
.
J. Neurobiol.
18
,
61
73
.
Kuwada
,
J. Y.
and
Kramer
,
A. P.
(
1983
).
Embryonic development of the leech nervous system: Primary axon outgrowth of identified neurons
.
J. Neurosci.
3
,
2098
2111
.
Lankford
,
K. L.
and
Letourneau
,
P. C.
(
1989
).
Evidence that calcium may control neurite outgrowth by regulating the stability of actin filaments
.
J. Cell Biol.
109
,
1229
1243
.
Legrand
,
C.
,
Ferraz
,
C.
,
Clavel
,
M.-C.
and
Rabie
,
A.
(
1991
).
Distribution of gelsolin in the retina of the developing rabbit
.
Cell Tissue Res.
264
,
335
338
.
Letourneau
,
P. C.
and
Ressler
,
A. H.
(
1983
).
Differences in the organization of actin in the growth cones compared with the neurites of cultured neurons from chick embryos
.
J. Cell Biol.
97
,
963
973
.
Letourneau
,
P. C.
and
Ressler
,
A. H.
(
1984
).
Inhibition of neurite initiation and growth by taxol
.
J. Cell Biol.
98
,
1355
1362
.
Letourneau
,
P. C.
,
Shattuck
,
T. A.
and
Ressler
,
A. H.
(
1986
).
Branching of sensory and sympathetic neurites in vitro is inhibited by treatment with taxol
.
J. Neurosci.
6
,
1912
1917
.
Levay
,
S.
and
Gilbert
,
C. D.
(
1976
).
Laminar patterns of geniculocortical projection in the cat
.
Brain Res.
113
,
1
19
.
Levay
,
S.
,
Stryker
,
M. P.
and
Shatz
,
C. J.
(
1978
).
Ocular dominance columns and their development in layer IV of the cat’s visual cortex: A quantitative study
.
J. comp. Neurol.
179
,
223
244
.
Lewis
,
A. K.
and
Bridgman
,
P. C.
(
1992
).
Nerve growth cone lamellipodia contain two populations of actin filaments that differ in organization and polarity
.
J. Cell Biol.
119
,
1219
1243
.
Lipscombe
,
D.
,
Madison
,
D. V.
,
Poenie
,
M.
,
Reuter
,
H.
,
Tsien
,
R. Y.
and
Tsien
,
R. W.
(
1988
).
Spatial distribution of calcium channels and cytosolic calcium transients in growth cones and cell bodies of sympathetic neurons
.
Proc. natn. Acad. Sci. U.S.A.
85
,
2398
2402
.
Llinas
,
R.
and
Sugimori
,
M.
(
1979
).
Calcium conductances in Purkinje cell dendrites: their role in development and integration
.
Prog. Brain Res.
51
,
323
334
.
Llinas
,
R.
and
Sugimori
,
M.
(
1980a
).
Electrophysiological properties of in vitro Purkinje cell somata in mammalian cerebellar slices
.
J. Physiol., Lond.
305
,
171
195
.
Llinas
,
R.
and
Sugimori
,
M.
(
1980b
).
Electrophysiological properties of in vitro Purkinje cell dendrites in mammalian cerebellar slices
.
J. Physiol., Lond.
305
,
197
213
.
Lo
,
Y.-J.
,
Lin
,
Y.-C.
,
Sanes
,
D. H.
and
Poo
,
M.-M.
(
1994
).
Depression of developing neuromuscular synapses induced by repetitive postsynaptic depolarizations
.
J. Neurosci.
14
,
4694
4704
.
Lo
,
Y.-J.
and
Poo
,
M.-M.
(
1991
).
Activity-dependent synaptic competition in vitro: Heterosynaptic suppression of developing synapses
.
Science
254
,
1019
1022
.
Lo
,
Y.-J.
and
Poo
,
M.-M.
(
1994
).
Heterosynaptic suppression of developing neuromuscular synapses in culture
.
J. Neurosci.
14
,
4684
4693
.
Loer
,
C. M.
,
Jellies
,
J.
and
Kristan
,
W. B.
Jr
,
(
1987
).
Segment-specific morphogenesis of leech Retzius neurons requires particular peripheral targets
.
J. Neurosci.
7
,
2630
2638
.
Lohof
,
A. M.
,
Quillan
,
M.
,
Dan
,
Y.
and
Poo
,
M.-M.
(
1992
).
Asymmetric modulation of cytosolic cyclic AMP activity induces growth cone turning
.
J. Neurosci.
12
,
1253
1261
.
Löwel
,
S.
and
Singer
,
W.
(
1992
).
Selection of intrinsic horizontal connections in the visual cortex by correlated neuronal activity
.
Science
255
,
209
212
.
Luduena
,
M. A.
and
Wessells
,
N. K.
(
1973
).
Cell locomotion, nerve elongation and microfilaments
.
Devl Biol.
30
,
427
440
.
Lund
,
R. D.
,
Mitchell
,
D. E.
and
Henry
,
G. H.
(
1978
).
Squint-induced modification of callosal connections in cats
.
Brain Res.
144
,
169
172
.
Manalan
,
A. S.
and
Klee
,
C. B.
(
1984
).
Calmodulin
.
Adv. cyclic Nucleotide Res.
18
,
227
278
.
Mariani
,
J.
(
1982
).
Extent of multiple innervation of Purkinje cells by climbing fibers in the olivocerebellar system of weaver, reeler and staggerer mutant mice
.
J. Neurobiol.
13
,
119
126
.
Mariani
,
J.
and
Changeux
,
J.-P.
(
1981
).
Ontogenesis of olivocerebellar relationships. I. Studies by intracellular recordings of the multiple innervation of Purkinje cells by climbing fibers in the developing rat cerebellum
.
J. Neurosci.
1
,
696
702
.
Mariani
,
J.
,
Crepel
,
F.
,
Mikoshiba
,
K.
,
Changeux
,
J.-P.
and
Sotelo
,
C.
(
1977
).
Anatomical, physiological and biochemical studies of the cerebellum from reeler mutant mouse
.
Phil. Trans. R. Soc. Lond. B
281
,
1
28
.
Marsh
,
L.
and
Letourneau
,
P. C.
(
1984
).
Growth of neurites without filopodial or lamellipodial activity in the presence of cytochalasin B
.
J. Cell Biol.
99
,
2041
2047
.
Mattson
,
M. P.
and
Kater
,
S. B.
(
1987
).
Calcium regulation of neurite elongation and growth cone motility
.
J. Neurosci.
7
,
4034
4043
.
Meiri
,
H.
,
Spira
,
M. E.
and
Parnas
,
I.
(
1981
).
Membrane conductance and action potential of a regenerating axonal tip
.
Science
211
,
709
712
.
Minkowski
,
M.
(
1920
).
Über den Verlauf, die Endigung und die zentrale Repräsentation von gekreuzten und ungekreuzten Sehnervenfasern bei einigen Säugetieren und beim Menschen
.
Schweiz Arch. Neurol. Psychiatr.
6
,
201
252
.
Nairn
,
A. C.
,
Hemmings
,
H. C.
and
Greengard
,
P.
(
1985
).
Protein kinases in the brain
.
A. Rev. Biochem.
54
,
931
976
.
Neely
,
M. D.
(
1993
).
Role of substrate and calcium in neurite retraction of leech neurons following depolarization
.
J. Neurosci.
13
,
1292
1301
.
Neely
,
M. D.
and
Gesemann
,
M.
(
1994
).
Disruption of microfilaments in growth cones following depolarization and calcium influx
.
J. Neurosci.
14
,
7511
7520
.
Nishi
,
R.
and
Berg
,
D. K.
(
1981
).
Effects of high K+concentrations on the growth and development of ciliary ganglion neurons in cell culture
.
Devl Biol.
87
,
301
307
.
O’brien
,
R. A. D.
,
Ostberg
,
A. J. C.
and
Vrbova
,
G.
(
1978
).
Observations on the elimination of polyneuronal innervation in developing mammalian skeletal muscle
.
J. Physiol., Lond.
282
,
571
582
.
O’leary
,
D. D. M.
(
1992
).
Development of connectional diversity and specificity in the mammalian brain by the pruning of collateral projections
.
Curr. Opin. Neurobiol.
2
,
70
77
.
O’leary
,
D. D. M.
,
Stanfield
,
B. B.
and
Cowan
,
W. M.
(
1981
).
Evidence that early postnatal restriction of the cells of origin of the callosal projection is due to the elimination of axonal collaterals rather than to the death of neurons
.
Devl Brain Res.
1
,
607
617
.
Papazian
,
D. M.
,
Schwartz
,
T. L.
,
Tempel
,
B. L.
,
Jan
,
Y. N.
and
Jan
,
L. Y.
(
1987
).
Cloning of genomic and complementary DNA from Shaker, a putative potassium channel gene from Drosophila
.
Science
237
,
749
753
.
Peinado
,
A.
,
Yuste
,
R.
and
Katz
,
L. C.
(
1993
).
Extensive dye coupling between rat neocortical neurons during the period of circuit formation
.
Neuron
10
,
103
114
.
Poo
,
M.-M.
and
Young
,
S. H.
(
1990
).
Diffusional and electrokinetic redistribution at the synapse: A physicochemical basis of synaptic competition
.
J. Neurobiol.
21
,
157
168
.
Privat
,
A.
and
Drian
,
M. J.
(
1976
).
Postnatal maturation of rat Purkinje cells cultivated in the absence of two afferent systems: An ultrastructural study
.
J. comp. Neurol.
166
,
201
244
.
Pumplin
,
D. W.
,
Reese
,
T. S.
and
Llinas
,
R.
(
1981
).
Are the presynaptic membrane particles the calcium channels?
Proc. natn. Acad. Sci. U.S.A.
78
,
7210
7213
.
Purves
,
D.
and
Lichtman
,
J. W.
(
1980
).
Elimination of synapses in the developing nervous system
.
Science
256
,
1823
1825
.
Rabacchi
,
S.
,
Bailly
,
Y.
,
Delhaye-Bouchaud
,
N.
and
Mariani
,
J.
(
1992
).
Involvement of the N-methyl D-aspartate (NMDA) receptor in synapse elimination during cerebellar development
.
Science
256
,
1823
1825
.
Rakic
,
P.
(
1976
).
Prenatal genesis of connections subserving ocular dominance in the rhesus monkey
.
Nature
261
,
467
471
.
Rakic
,
P.
(
1977
).
Prenatal development of the visual system in rhesus monkey
.
Phil. Trans. R. Soc. Lond. B
278
,
245
260
.
Rakic
,
P.
and
Riley
,
K. P.
(
1983
).
Overproduction and elimination of retinal axons in the fetal rhesus monkey
.
Science
219
,
1441
1444
.
Ramoa
,
A. S.
,
Campbell
,
G.
and
Shatz
,
C. J.
(
1988
).
Dendritic growth and remodeling of cat retinal ganglion cells during fetal and postnatal development
.
J. Neurosci.
8
,
4239
4261
.
Redfern
,
P. A.
(
1970
).
Neuromuscular transmission in new-born rats
.
J. Physiol., Lond.
209
,
701
709
.
Regehr
,
W. G.
,
Connor
,
J. A.
and
Tank
,
D. W.
(
1989
).
Optical imaging of calcium accumulation in hippocampal pyramidal cells during synaptic activation
.
Nature
341
,
533
536
.
Reiter
,
H. O.
,
Waitzman
,
D. M.
and
Stryker
,
M. P.
(
1986
).
Cortical activity blockade prevents ocular dominance plasticity in the kitten visual cortex
.
Expl Brain Res.
65
,
182
188
.
Rhoades
,
R. W.
and
Dellacroce
,
D. D.
(
1980
).
Neonatal enucleation induces an asymmetric pattern of visual callosal connections in hamsters
.
Brain Res.
202
,
189
195
.
Riley
,
D. A.
(
1977
).
Spontaneous elimination of nerve terminals from the endplates of developing skeletal myofibers
.
Brain Res.
134
,
279
285
.
Riley
,
D. A.
(
1978
).
Tenotomy delays the postnatal development of the motor innervation of the rat soleus
.
Brain Res.
143
,
162
167
.
Riley
,
D. A.
(
1981
).
Ultrastructural evidence for axon retraction during the spontaneous elimination of polyneuronal innervation of the rat soleus muscle
.
J. Neurocytol.
10
,
425
440
.
Robson
,
S. J.
and
Burgoyne
,
R. D.
(
1989
).
L-type calcium channels in the regulation of neurite outgrowth from rat dorsal root ganglion neurons in culture
.
Neurosci. Lett.
104
,
110
114
.
Ross
,
W. N.
,
Aréchiga
,
H.
and
Nicholls
,
J. G.
(
1988
).
Influence of substrate on the distribution of calcium channels in identified leech neurons in culture
.
Proc. natn. Acad. Sci. U.S.A.
85
,
4075
4078
.
Ross
,
W. N.
,
Lasser-Ross
,
N.
and
Werman
,
R.
(
1990
).
Spatial and temporal analysis of calcium-dependent electrical activity in guinea pig Purkinje cell dendrites
.
Proc. R. Soc. Lond. B
240
,
173
185
.
Schilling
,
K.
,
Dickinson
,
M. H.
,
Connor
,
J. A.
and
Morgan
,
J. I.
(
1991
).
Electrical activity in cerebellar cultures determines Purkinje cell dendritic growth patterns
.
Neuron
7
,
891
902
.
Schulman
,
H.
,
Hanson
,
P. I.
and
Meyer
,
T.
(
1992
).
Decoding calcium signals by multifunctional CaM kinase
.
Cell Calcium
13
,
401
411
.
Schwab
,
M. E.
,
Kapfhammer
,
J. P.
and
Bandtlow
,
C. E.
(
1993
).
Inhibitors of neurite growth
.
A. Rev. Neurosci.
16
,
565
595
.
Shatz
,
C. J.
(
1990
).
Impulse activity and the patterning of connections during CNS development
.
Neuron
5
,
745
756
.
Shatz
,
C. J.
and
Kirkwood
,
P. A.
(
1984
).
Prenatal development of functional connections in the cat’s retinogeniculate pathway
.
J. Neurosci.
4
,
1378
1397
.
Shatz
,
C. J.
,
Lindström
,
S.
and
Wiesel
,
T. N.
(
1977
).
The distribution of afferents representing the right and left eyes in the cat’s visual cortex
.
Brain Res.
131
,
103
116
.
Shatz
,
C. J.
and
Stryker
,
M. P.
(
1978
).
Ocular dominance in layer IV of the cat’s visual cortex and the effects of monocular deprivation
.
J. Physiol., Lond.
281
,
267
283
.
Shatz
,
C. J.
and
Stryker
,
M. P.
(
1988
).
Prenatal tetrodotoxin infusion blocks segregation of retinogeniculate afferents
.
Science
242
,
87
89
.
Silver
,
R. A.
,
Lamb
,
A. G.
and
Bolsover
,
S. R.
(
1990
).
Calcium hotspots caused by L-channel clustering promote morphological changes in neuronal growth cones
.
Nature
343
,
751
754
.
Smith
,
S. J.
and
Augustine
,
G. J.
(
1988
).
Calcium ions, active zones and synaptic transmitter release
.
Trends Neurosci.
11
,
458
464
.
Sobue
,
K.
(
1993
).
Actin-based cytoskeleton in growth cone activity
.
Neurosci. Res.
18
,
91
102
.
Spitzer
,
N. C.
(
1979
).
Ion channels in development
.
A. Rev. Neurosci.
2
,
363
397
.
Spitzer
,
N. C.
and
Baccaglini
,
P. I.
(
1976
).
Development of the action potential in embryo amphibian neurons in vivo
.
Brain Res.
107
,
610
616
.
Sretavan
,
D. W.
and
Shatz
,
C. J.
(
1986
).
Prenatal development of retinal ganglion cell axons: Segregation into eye-specific layers within the cat’s lateral geniculate nucleus
.
J. Neurosci.
6
,
234
251
.
Sretavan
,
D. W.
,
Shatz
,
C. J.
and
Stryker
,
M. P.
(
1988
).
Modification of retinal ganglion cell axon morphology by prenatal infusion of tetrodotoxin
.
Nature
336
,
468
471
.
Stanfield
,
B. B.
(
1992
).
The development of the corticospinal projection
.
Prog. Neurobiol.
38
,
169
202
.
Stanfield
,
B. B.
,
O’leary
,
D. D. M.
and
Fricks
,
C.
(
1982
).
Selective collateral elimination in early postnatal development restricts cortical distribution of rat pyramidal tract neurones
.
Nature
298
,
371
373
.
Stephens
,
P. J.
and
Govind
,
C. K.
(
1981
).
Peripheral innervation fields of single lobster motoneurons defined by synapse elimination during development
.
Brain Res.
212
,
476
480
.
Stossel
,
T. P.
,
Chaponnier
,
C.
,
Ezzell
,
R. M.
,
Hartwig
,
J. H.
,
Janmey
,
P. A.
,
Kwiatkowski
,
D. J.
,
Lind
,
S. E.
,
Smith
,
D. B.
,
Southwick
,
F. S.
,
Yin
,
H. L.
and
Zaner
,
K. S.
(
1985
)
Nonmuscle actin-binding proteins
.
A. Rev. Cell Biol.
1
,
353
402
.
Stryker
,
M. P.
and
Harris
,
W. A.
(
1986
).
Binocular impulse blockade prevents the formation of ocular dominance columns in cat visual cortex
.
J. Neurosci.
6
,
2117
2133
.
Suarez-Isla
,
B. A.
,
Pelto
,
D. J.
,
Thompson
,
J. M.
and
Rapoport
,
S. I.
(
1984
).
Blockers of calcium permeability inhibit neurite extension and formation of neuromuscular synapses in cell culture
.
Devl Brain Res.
14
,
263
270
.
Tanaka
,
J.
,
Kira
,
M.
and
Sobue
,
K.
(
1993
).
Gelsolin is localized in neuronal growth cones
.
Devl Brain Res.
76
,
268
271
.
Tanaka
,
J.
and
Sobue
,
K.
(
1994
).
Localization and characterization of gelsolin in nervous tissues: Gelsolin is specifically enriched in myelin-forming cells
.
J. Neurosci.
14
,
1038
1052
.
Tank
,
D. W.
,
Sugimori
,
M.
,
Connor
,
J. A.
and
Llinas
,
R. R.
(
1988
).
Spatially resolved calcium dynamics of mammalian Purkinje cells in cerebellar slice
.
Science
242
,
773
777
.
Tempel
,
B. L.
,
Papazian
,
D. M.
,
Schwartz
,
T. L.
,
Yan
,
Y. N.
and
Jan
,
L. Y.
(
1987
).
Sequence of a probable potassium channel component encoded at Shaker locus of Drosophila
.
Science
237
,
770
775
.
Thompson
,
W.
(
1983
).
Synapse elimination in neonatal rat muscle is sensitive to pattern of muscle use
.
Nature
302
,
614
616
.
Thompson
,
W. J.
(
1986
).
Changes in the innervation of mammalian skeletal muscle fibers during postnatal development
.
Trends Neurosci.
9
,
25
28
.
Thompson
,
W.
,
Kuffler
,
D. P.
and
Jansen
,
J. K. S.
(
1979
).
The effect of prolonged, reversible block of nerve impulses on the elimination of polyneuronal innervation of new-born rat skeletal muscle fibers
.
Neuroscience
4
,
271
281
.
Usowicz
,
M. M.
,
Porzig
,
H.
,
Becker
,
C.
and
Reuter
,
H.
(
1990
).
Differential expression by nerve growth factor of two types of Ca2+-channels in rat phaeochromocytoma cell lines
.
J. Physiol., Lond.
426
,
95
116
.
Vandekerckhove
,
J.
(
1990
).
Actin-binding proteins
.
Curr. Opin. Cell Biol.
2
,
41
50
.
Van Essen
,
D. C.
(
1982
).
Neuromuscular synapse elimination
. In
Neural Development
(ed.
N. C.
Spitzer
), pp.
333
376
.
New York
:
Plenum Press
.
Wallace
,
B. G.
(
1984
).
Selective loss of neurites during differentiation of cells in the leech central nervous system
.
J. comp. Neurol.
228
,
149
153
.
Weeds
,
A.
and
Maciver
,
S.
(
1993
).
F-actin capping proteins
.
Curr. Opin. Cell Biol.
5
,
63
69
.
Wiesel
,
T. N.
and
Hubel
,
D. H.
(
1963
).
Single-cell responses in striate cortex of kittens deprived of vision in one eye
.
J. Neurophysiol.
26
,
1003
1017
.
Wolszon
,
L. R.
,
Gao
,
W.-Q.
,
Passani
,
M. B.
and
Macagno
,
E. R.
(
1994
).
Growth cone ‘collapse’ in vivo: Are inhibitory interactions mediated by gap junctions?
J. Neurosci.
14
,
999
1010
.
Wolszon
,
L. R.
and
Macagno
,
E. R.
(
1992
).
Growth cone interactions and the patterns of peripheral innervation in the leech
. In
The Nerve Growth Cone
(ed.
P. C.
Letourneau
,
S. B.
Kater
and
E. R.
Macagno
), pp.
305
335
.
New York
:
Raven Press
.
Wong
,
R. O. L.
,
Herrmann
,
K.
and
Shatz
,
C. J.
(
1991
).
Remodeling of retinal ganglion cell dendrites in the absence of action potential activity
.
J. Neurobiol.
22
,
685
697
.
Yamada
,
K. M.
,
Spooner
,
B. S.
and
Wessells
,
N. K.
(
1970
).
Axon growth: Roles of microfilaments and microtubules
.
Proc. natn. Acad. Sci. U.S.A.
66
,
1206
1212
.
Yamada
,
K. M.
,
Spooner
,
B. S.
and
Wessells
,
N. K.
(
1971
).
Ultrastructure and function of growth cones and axons of cultured nerve cells
.
J. Cell Biol.
49
,
614
635
.
Yin
,
H. L.
,
Albrecht
,
J. H.
and
Fattoum
,
A.
(
1981a
).
Identification of gelsolin, a Ca2+-dependent regulatory protein of actin gel–sol transformation and its intracellular distribution in a variety of cells and tissues
.
J. Cell Biol.
91
,
901
906
.
Yin
,
H. L.
,
Hartwig
,
J. H.
,
Maruyama
,
K.
and
Stossel
,
T. P.
(
1981b
).
Ca2+control of actin filament length
.
J. biol. Chem.
256
,
9693
9697
.
Yin
,
H. L.
,
Zaner
,
K. S.
and
Stossel
,
T. P.
(
1980
).
Ca2+control of actin gelation
.
J. biol. Chem.
255
,
9494
9500
.
Zhong
,
Y.
,
Budnik
,
V.
and
Wu
,
C.-F.
(
1992
).
Synaptic plasticity in Drosophila memory and hyperexcitable mutants: Role of cyclic AMP cascade
.
J. Neurosci.
12
,
644
651
.