The L1 CAM family of cell adhesion molecules and the ankyrin family of spectrin-binding proteins are candidates to collaborate in transcellular complexes used in diverse contexts in nervous systems of vertebrates and invertebrates. This report presents evidence for functional coupling between L1 and 440-kD ankyrinB in premyelinated axons in the mouse nervous system. L1 and 440-kD ankyrinB are colocalized in premyelinated axon tracts in the developing nervous system and are both down-regulated after myelination. AnkyrinB (−/−) mice exhibit a phenotype similar to, but more severe, than L1 (−/−) mice and share features of human patients with L1 mutations. AnkyrinB (−/−) mice exhibit hypoplasia of the corpus callosum and pyramidal tracts, dilated ventricles, and extensive degeneration of the optic nerve, and they die by postnatal day 21. AnkyrinB (−/−) mice have reduced L1 in premyelinated axons of long fiber tracts, including the corpus callosum, fimbria, and internal capsule in the brain, and pyramidal tracts and lateral columns of the spinal cord. L1 was evident in the optic nerve at postnatal day 1 but disappeared by postnatal day 7 in mutant mice while NCAM was unchanged. Optic nerve axons of ankyrinB (−/−) mice become dilated with diameters up to eightfold greater than normal, and they degenerated by day 20. These findings provide the first evidence for a role of ankyrinB in the nervous system and support an interaction between 440-kD ankyrinB and L1 that is essential for maintenance of premyelinated axons in vivo.

The L1 CAM1 family of cell adhesion molecules (comprised of L1, NgCAM, NrCAM, CHL1, and neurofascin) is a major contributor to intercellular interactions in the vertebrate nervous system (Hortsch, 1996; Brummendorf et al., 1998). L1 CAM molecules, which are represented in Drosophila by neuroglian (Bieber et al., 1989), possess variable extracellular domains composed of six Ig domains, three to five fibronectin type III domains, and a relatively conserved cytoplasmic domain (Sonderegger and Rathjen, 1992; Hortsch, 1996). Extracellular domains of L1 CAMs participate in homophilic interactions as well as a variety of interactions with soluble proteins and other cell adhesion molecules (Hortsch, 1996; Brummendorf et al., 1998). Extensive diversity in extracellular interactions of L1 CAM family members is provided by divergence between extracellular domains encoded by different genes as well as multiple alternatively spliced variants of each gene (as many as 50 estimated in the case of neurofascin) (Hassel et al., 1997). Consistent with the diverse extracellular domains, a range of functions has been attributed to the L1 CAM family, including axon fasciculation, axonal guidance, neurite extension, a role in long term potentiation, synaptogenesis, and myelination (Hortsch, 1996; Walsh and Doherty, 1997; Brummendorf et al., 1998).

Mutations in the human L1 gene, located on the X chromosome, are responsible for mental retardation, as well as hydrocephalus, abnormal development of major axon pathways in the corpus callosum and corticocospinal tract, and motor deficits (Rosenthal et al., 1992; Jouet et al., 1994; Wong et al., 1995; Fransen et al., 1997). Penetrance of L1 mutations is variable, except for mental retardation, which is the most common abnormality. The term CRASH (corpus callosum agenesis, mental retardation, adducted thumbs, spastic paraplegia, and hydrocephalus) has been used to describe general features of L1 mutations (Fransen et al., 1995). Over 35 mutations in L1 have currently been characterized, and these are distributed in all domains, including several in the cytoplasmic domain (Van Camp et al., 1996; Fransen et al., 1998a). Disruption of the L1 gene in mice results in abnormal axon guidance in the corticospinal tract, brain malformations (including dilated ventricles), and behavioral deficits (Dahme et al., 1997; Cohen et al., 1998; Fransen et al., 1998b). These observations demonstrate the physiological importance of L1 in normal development of the nervous system as well as a role of L1 in human disorders.

The cytoplasmic domains of L1 CAM family members share a conserved sequence that has been demonstrated to associate with high affinity with ankyrins, a family of spectrin-binding proteins located on the cytoplasmic surface of the plasma membrane (Davis and Bennett, 1994; Dubreuil et al., 1996; Garver et al., 1997; Hortsch et al., 1998; Zhang et al., 1998). The association between ankyrins and L1 CAMs is regulated by phosphorylation of the FIGQY tyrosine, a site which is conserved among vertebrate and invertebrate members of the L1CAM family (Garver et al., 1997; Tuvia et al., 1997). Ankyrins also associate with the voltage-gated Na channel, Na/K ATPase, anion channel, and Na/Ca exchanger. Ankyrins have been proposed to form lateral complexes between these ion channels and members of the L1 CAM family (Michaely and Bennett, 1995a,b; Bennett et al., 1997).

The ankyrin family currently includes three genes encoding ankyrinR, ankyrinB, and ankyrinG polypeptides as well as multiple alternatively spliced variants encoded by each gene (Peters and Lux, 1993; Bennett et al., 1997). 440-kD ankyrinB and 480-kD ankyrinG polypeptides are the major isoforms in the developing nervous system and are localized in premyelinated and unmyelinated axons (Kunimoto et al., 1991; Otto et al., 1991; Chan et al., 1993; Kunimoto, 1995; Lambert et al., 1997). These axonal ankyrins have a predicted ball-and-chain configuration, with an extended tail domain inserted between the COOH- terminal death domain and globular head domain. The head domain is composed of membrane- and spectrin-binding domains (Chan et al., 1993; Kordeli et al., 1995). The extended tail domains of axonal ankyrins have the potential to interact with cytoskeletal elements, and thus to couple L1 CAM molecules to cytoplasmic proteins in addition to the membrane-associated spectrin–actin network.

L1 CAMs and ankyrins are candidates to collaborate in a major structural pathway, involving both extracellular and intracellular interactions, which is used in diverse contexts in the developing and adult nervous system. This report presents evidence for functional coupling between L1 and 440-kD ankyrinB in premyelinated axon tracts in the neonatal mouse nervous system. Disruption of the ankyrinB gene in the mouse results in a phenotype similar to, but more severe, than the L1 knock-out in mice and L1 mutations in humans, as well as loss of L1 from premyelinated axon tracts. These findings provide evidence that ankyrins and L1CAMs associate in vivo, and furthermore identify the genes and alternatively spliced variants involved in L1 CAM–ankyrin interactions in premyelinated axons.

Gene Targeting

A clone containing 17 kb of the ank2 gene was isolated from a 129SVJ genomic lambda fix II library using a probe that covers nucleotides 2443– 2805 of the human ankyrinB cDNA. A novel NotI site was introduced within an exon in the spectrin-binding domain, and this site was used to introduce a cassette that included the neomycin resistance gene (pMC1neo polyA vector; Stratagene, La Jolla, CA), an in-frame HA epitope, and a stop codon. An XhoI/HindIII fragment containing the thymidine kinase gene (from the vector pIC19R) was added to the construct. ES cells were transfected by electroporation with the linearized targeting construct, neomycin- and gancycolivir-resistant clones were isolated, and homologous recombination was verified by Southern blotting (see Fig. 1 B). ES cells were injected into blastocysts from C57Bl/6 mice and implanted in C57Bl/6 pseudopregnant mothers. Male chimeric progeny were mated with C57Bl/6 females to generate heterozygous mice. Homozygous mice were produced from none-littermate heterozygous mice. ES cell culture, transfection, and blastocyst injection were performed in the laboratory of Dr. Beverly Koller (University of North Carolina, Chapel Hill, NC) (Mansour et al., 1988; Koller et al., 1991).

Magnetic Resonance Imaging

Magnetic resonance (MR) microscopy was performed on a General Electric (Fremont, CA) 7.1 Tesla unit, 15-cm bore size, equipped with shielded gradient coils. A Helmholtz coil especially designed for MR microscopy of fixed specimens at 300 MHz was used for imaging (Banson et al., 1992). Formalin-fixed brains were suspended in Fomblin (perfluoropolyether; Ausimont, Inc., Morristown, NJ) to minimize dehydration and to limit susceptibility effects at the surface. Data was acquired as anisotropic arrays with an aspect ratio of 1:1:2 with the following acquisition parameters: (a) 18.2-h scan time; (b) 20-ms echo time; (c) 1,000-ms repetition time, diffusion gradient pulse (sine bell-shaped) durations of 5 ms; (d) diffusion gradient separation of 12.2 ms; (e) large diffusion gradients of 35 Gauss/cm, resulting in a diffusion-weighting factor of 965 s/mm2, 256 × 256 × 128 digital resolution; and (f) voxel size of 39 × 39 × 78 μm. All images were analyzed using NIH Image 1.68 (public domain from The National Institutes of Health) and VoxelView Ultra (Vital Images, Inc., Fairfield, IA) for three-dimensional rendering.

Analysis of Tissues

Immunofluorescence was performed with 5–10-μm frozen brain sections, which were prepared and stained as described (Lambert et al., 1997). Affinity-purified rabbit antibodies were raised against the human COOH-terminal domain (residues 1551–1813) and 440-kD–specific tail domain (Chan et al., 1993) of ankyrinB using recombinant proteins expressed in bacteria as antigens as described (Lambert et al., 1997). Chick antibody against the 270/480-kD ankyrinG has been described elsewhere (Lambert et al., 1997). Rabbit polyclonal antibody against neurofascin was described in Davis et al. (1996). Antibody against L1 was purchased from Boehringer Mannheim Corp. (Indianapolis, IN). Electron microscopy was performed with optic nerve fixed with 4% glutaraldehyde and subsequently embedded in Epon. To determine the sizes of axons in the knock-out and the control mouse optic nerves, a sheet of paper containing a hole with dimensions of 8 × 8 cm was mounted on the electron micrographs, and the number of axons that appeared in the hole were counted. Northern and Southern blots (Sambrook et al., 1989) and immunoblots (Lambert et al., 1997) were performed as described. Antibodies used in immunoblotting experiments are anti-L1, a protein A affinity-purified polyclonal antibody (a generous gift from Dr. Patricia Manness of the University of North Carolina at Chapel Hill); polyclonal antispectrin and polyclonal antineurofascin as described in previous reports (Hu et al., 1995; Davis et al., 1996); and polyclonal anti-NCAM, purchased from Chemicon (Temecula, CA).

AnkyrinB (−/−) Mice Exhibit Brain Malformations Resembling Defects Caused by L1 Mutations in Humans and Mice

The ank2 gene encoding ankyrinB was disrupted by homologous recombination (Fig. 1,A), and ankyrinB +/+, +/−, and −/− mice (n = 325) were identified by Southern blots (Fig. 1,B) (Materials and Methods). This study is based on the first-generation progeny with a genetic background of 1:1 129SVJ to C57Bl/6. Brains of ankyrinB (−/−) mice exhibited nearly complete loss of a major 13-kb mRNA transcript encoding the 440-kD ankyrinB (Kunimoto et al., 1991) as well as loss of less abundant 8- and 5-kb transcripts encoding 220- and 150-kD ankyrinB polypeptides, respectively (Fig. 1,C). 440- and 220-kD ankyrinB polypeptides were undetectable in homogenates of brains of ankyrinB (−/−) mice by immunoblotting (Fig. 1 D). A minor 150-kD polypeptide in the ankyrinB (−/−) mice still is barely detectable in original gels, although not in the published version. 150-kD ankyrinB lacks a membrane-binding domain (Otto et al., 1991). Apparently expression of the 150-kD ankyrinB transcript is less affected by disruption of an exon in the spectrin binding domain (Materials and Methods).

440-kD ankyrinB is a major isoform in the early postnatal period, as noted previously (Kunimoto et al., 1991). AnkyrinB is also expressed in heart where the major isoform of 220 kD was undetectable in ankyrinB (−/−) mice (Fig. 1 D).

AnkyrinB (−/−) mice were born in nearly Mendelian ratios of 1:3 (81/325), demonstrating normal survival during fetal development. However, over half of the ankyrinB (−/−) animals were dead or dying (and therefore killed) on the first postnatal day, and 95% by 8 d (n = 81) (Fig. 2,B). A few mutant mice survived until postnatal day 20 (4 out of 81). Mice that did survive to 2 wk were about 25% of the weight of littermates (n = 8) and exhibited abnormal locomotion and balance (Fig. 2 A).

Three-dimensional diffusion-weighted magnetic resonance microscopy (DWM) of formalin-fixed brains from a litter of 1-d-old mice (three mutant and three normal) was performed to obtain general information on gross morphology and accurate measurements of ventricle volumes. We used DWM, which measures diffusion of free proton spins to get optimal anatomical resolution. Brain anatomy is revealed in detail with DWM because proton diffusion within regions composed of oriented axon fiber tracts and those comprised of cell bodies is markedly different (Beaulieu and Allen, 1996).

Total brain and cerebellar volumes in the ankyrinB (−/−) brains were nearly identical to those of (+/+) and (+/−) littermates (Table I). However, volumetric measurements of the ventricular systems revealed a significant sevenfold enlargement of the lateral ventricles in the ankyrinB (−/−) brains compared with controls (Table I; Fig. 3). In addition, an enlarged lateral ventricle is easily visualized in Fig. 3. Enlargement of lateral ventricles of ankyrinB (−/−) mice was not associated with occlusion of the cerebral aqueduct, which is closed in Fig. 3 but was patent in other sections (not shown). The extent of lateral ventricle dilation in neonatal ankyrinB (−/−) mice is in agreement with that reported for some strains of adult L1 (−/−) mice (Dahme et al., 1997; Fransen et al., 1998b). Fransen et al. (1998b) found a twofold enlargement of both the frontolateral and fourth ventricle in adult L1 (−/−) mice. AnkyrinB (−/−) mice did not exhibit fourth ventricle enlargement; however, they did not survive beyond day 21, and the full extent of enlargement of brain ventricles could therefore not be determined. L1 (−/−) mice in some genetic backgrounds do not exhibit enlarged ventricles (Dahme et al., 1997). Hydrocephalus is also variable in extent in humans with L1 mutations, even within the same kindred (Fransen et al., 1997). L1 hydrocephalus also is associated with a patent aqueduct of Sylvius in some cases (Fransen et al., 1995; Wong et al., 1995).

L1 mutations in mice and humans are associated with abnormalities in the pyramidal tracts and corpus callosum (Fransen et al., 1997; Cohen et al., 1998). DWM images revealed the presence of the corpus callosum and internal capsule in sagittal sections as well as pyramidal tracts in coronal sections in control mice (Fig. 3). The corpus callosum, internal capsule, and pyramidal tracts could not be resolved in brains of ankyrinB (−/−) mice (Fig. 3). Histological brain sections also reveal a reduction in the corpus callosum in ankyrinB (−/−) mice, with nearly complete absence of the anterior portion revealed in a sagittal section (Fig. 5). The corpus callosum also was greatly reduced in coronal sections of ankyrinB mutant brain (Fig. 5). Hypoplasia of the corpus callosum was observed in 16/26 (postnatal days 1–5) mutant mouse brains examined by coronal sections. Some of the variability in detection of reduction of the corpus callosum may be due to selective loss of anterior relative to posterior regions.

Loss of L1 in Premyelinated Axon Tracts in the Central Nervous System of AnkyrinB (−/−) Mice

The similarities in phenotype of ankyrinB (−/−) mice and L1 (−/−) mice suggested the possibility that ankyrinB and L1 are coexpressed in the nervous system. Evidence in support of segregation of L1 with ankyrinB and of neurofascin with ankyrinG is presented in Fig. 4. Neurofascin has previously been identified at axon initial segments and nodes of Ranvier, which are sites that also contain high local concentrations of 270/480-kD isoforms of ankyrinG (Davis et al., 1996; Lambert et al., 1997). Neurofascin and 270/480 ankyrinG are colocalized at axon initial segments in the neonatal cerebral cortex (Fig. 4). However, in neonatal mouse brain, neurofascin and 270/480 ankyrinG were absent from premyelinated axon tracts, including the corpus callosum (Fig. 4, lower panels) as well the optic nerve and other sites in the central nervous system (data not shown). L1 and 440 ankyrinB, in contrast, codistributed in premyelinated axon fiber tracts in the cerebral cortex (Fig. 4) and throughout the CNS (Figs. 5 and 6).

440-kD ankyrinB and L1 are colocalized in premyelinated axons located in long fiber tracts in brain (corpus callosum, internal capsule, fornix, striatum, and olfactory tract) (Fig. 5) as well as spinal cord (pyramidal tract and lateral columns) (Fig. 6). Similar patterns of ankyrinB staining in axon tracts were observed with antibodies against the COOH terminus, which react with all known alternatively spliced forms as with antibodies against 440-kD ankyrinB (not shown). These patterns of distribution of L1 are consistent with previous studies (Bartsch et al., 1989; Joosten and Gribnau, 1989; Martini, 1994). Brain and spinal cord sections of ankyrinB (−/−) mice exhibited no staining with antibody against the COOH terminus of ankyrinB, confirming both the efficacy of gene disruption and specificity of the antibody.

L1 staining is reduced in premyelinated axon tracts in brains and spinal cords of ankyrinB (−/−) mice (Figs. 5 and 6). The corpus callosum and internal capsule are reduced or absent in mutant animals (Fig. 5) (16/26), as noted above. However, fiber tracts in the brain that still were present and exhibited major reduction in L1 staining include the fimbria and stria terminalis (Fig. 5,A). L1 staining in the pyramidal tract and lateral columns of the spinal cord also is greatly reduced (Fig. 6). AnkyrinB (−/−) animals exhibited variability with respect to L1 staining, with about 50% having less pronounced differences. For example, L1 staining still is evident in the sagittal section shown in Fig. 5 B. However, littermate control animals always exhibited strong L1 staining.

Loss of L1 in ankyrinB (−/−) mice is also demonstrated by immunoblotting of brain tissue (Fig. 7). It was not feasible to obtain tissue samples from particular anatomical sites such as corpus callosum or optic nerve (see below), where loss of L1 is most prominent. Therefore, intact brains of a control and a mutant postnatal day 7 mouse were evaluated for levels of L1, neurofascin, NCAM, and spectrin (n = 1) (Fig. 7). Expression levels of neurofascin and NCAM exhibited no detectable change in the brain of the ankyrinB (−/−) mouse. Moreover, levels of βG-spectrin also were unchanged (Fig. 7). However, immunoblots of the same gel samples with antibody against L1 revealed a 30–40% percent reduction in amounts of 200-, 135-, and 43-kD cross-reacting polypeptides (Fig. 7). These results demonstrating partial loss of L1 are consistent with reduction of L1 in localized sites detected by immunofluorescence.

Loss of L1 staining in ankyrinB (−/−) brain and spinal cord raises questions of whether lack of ankyrinB impairs delivery of L1 to axons, whether L1-expressing axon tracts never form and are simply absent, and/or whether L1-expressing axons are present initially but subsequently degenerate. The optic nerve provides a simplified, easily identifiable anatomical site for addressing issues related to L1 behavior and axonal degeneration in ankyrinB (−/−) mice. The premyelinated optic nerve has previously been identified as a site where the 440-kD ankyrinB is the sole ankyrin (Chan et al., 1993), and also expresses high levels of L1 (Bartsch et al., 1989). 440-kD ankyrinB and L1 both exhibit a dramatic down-regulation coincident with myelination and are not expressed in adult myelinated axons of the optic nerve (Bartsch et al., 1989; Chan et al., 1993).

Transverse sections through optic nerve at postnatal days 1–3 revealed unreduced staining for L1 in ankyrinB (−/−) mice, even though ankyrinB staining was absent (n = 7) (Fig. 8). These observations establish that at least early in the postnatal period, optic nerve axons are present, and that L1 is delivered to axons in the absence of ankyrinB. However, L1 staining disappeared from optic nerves of ankyrinB (−/−) mice by around postnatal day 7, although the nerve itself was still present (Fig. 8) (n = 6). Optic nerves of normal mice retained L1 and ankyrinB (Fig. 8). In contrast to L1, levels of NCAM were unchanged in the optic nerve section of ankyrinB (−/−) mice at postnatal day 7 (Fig. 8). Neurofascin was only found at the initial segments of the premyelinated optic nerve axons and was absent from the optic nerves of both control and ankyrinB (−/−) mice. Electron micrographs (see below; Fig. 9) reveal that myelination occurred with only a few axons at this stage and was limited to a few loops of myelin. The normal loss of L1 accompanying myelination therefore had not yet begun at the time when L1 disappeared from ankyrinB (−/−) optic nerves.

Degeneration of the Optic Nerve in AnkyrinB (−/−) Mice

Electron microscopy of transverse sections of the postnatal day 1 optic nerve revealed relatively normal axon profiles in ankyrinB (−/−) mice, although the average axon cross-sectional area was increased about 20% (see Materials and Methods; data not shown). Axons were packed in tightly assembled arrays, indicating that some level of axon fasciculation occurs in the absence of ankyrinB. Sections of postnatal day 9 optic nerve reveal a major change in axons of ankyrinB (−/−) mice (Fig. 9). Many axons were increased in diameter up to eightfold compared with the normal axon and contained multivescicular bodies. Longitudinal sections demonstrated increased diameter along the length of ankyrinB (−/−) axons (not shown). Axon fasciculation also was impaired in sections of ankyrinB (−/−) optic nerve with many examples of loss of direct axon–axon contacts (Fig. 9).

All ankyrinB (−/−) mice surviving until postnatal day 21 exhibited nearly complete degeneration of the optic nerve (n = 8) (Fig. 10). The optic nerve of ankyrinB (−/−) mice thus forms initially with normal axonal delivery of L1 and subsequently degenerates. AnkyrinB and ankyrinG are coexpressed before myelination in most other nerves, including the trigeminal nerve (Lambert et al., 1997). The optic nerve, which only has ankyrinB, apparently is especially vulnerable in ankyrinB (−/−) mice.

This study describes dramatic consequences of deficiency of ankyrinB in mice that allow survival until birth but result in hypoplasia of the corpus callosum and pyramidal tracts, dilated lateral ventricles, followed by extensive degeneration of the optic nerve, and death by postnatal day 21. This phenotype is similar to, but more severe, than those observed in L1 (−/−) mice and human patients with L1 mutations. L1 was reduced or absent in ankyrinB (−/−) mutant mice in premyelinated axons of long fiber tracts in the brain and spinal cord as well as optic nerve at postnatal day 7. These findings provide the first evidence for a role of ankyrinB in the vertebrate nervous system and support a functional connection between ankyrinB and L1 that is required for maintenance of premyelinated axons in vivo.

L1 deficiency in mice is compatible with survival to adulthood (Dahme et al., 1997; Cohen et al., 1998; Fransen et al., 1998b) and represents a much less severe phenotype than ankyrinB deficiency. Other L1 CAM family members (CHL1, NrCAM, NgCAM, and neurofascin) may functionally compensate for L1 in L1 (−/−) mice. In addition, ankyrinB is likely to have additional roles unrelated to L1, particularly early in development when some L1 CAMs are tyrosine phosphorylated and therefore incapable of binding to ankyrin (Garver et al., 1997; Zisch et al., 1997; Heiland et al., 1998). Conversely, L1 and other members of the family are likely to have ankyrin-independent functions when these proteins are tyrosine phosphorylated. This study demonstrates that early in development, L1 is capable of axonal transport to the optic nerves of ankyrinB (−/−) mice, implying that its delivery to the plasma membrane and axonal targeting are independent of ankyrinB. It will be of interest to determine the phosphorylation state of L1 in the early postnatal optic nerve, where this protein is transported to axons in the absence of ankyrinB (Fig. 6).

The mechanism for loss of L1 in ankyrinB (−/−) axons is not known, but it could result from reduced stability of L1 in the absence of ankyrinB. A similar situation in which loss of a cytoskeletal protein results in reduction of integral proteins has been found in mdx mice, where absence of dystrophin results in reduction of dystrophin-associated proteins (Matsumura et al., 1992). A practical consequence of reduced expression of interacting proteins is that a characterization of patterns of gene or protein expression in the ankyrinB (−/−) mouse brain may yield heretofore unidentified proteins whose functions involve ankyrinB.

The striking increase in axonal diameter of optic nerve axons suggests the hypothesis that the absence of ankyrinB leads to loss of mechanical stability of axons. AnkyrinB could contribute to axonal membrane support by connecting the spectrin–actin network to the plasma membrane, as occurs in erythrocytes (Peters and Lux, 1993). In this case, axonal swelling would be analogous to hereditary spherocytosis of erythrocytes observed in cases of ankyrinR mutations (Eber et al., 1996). The tail domain of the 440-kD ankyrinB potentially could extend several hundred nanometers into the axoplasm and also mediate interactions with structural proteins in addition to the spectrin skeleton. L1-mediated connections between axons combined with coupling to the spectrin–actin network via ankyrin could provide an additional level of mechanical support for axon bundles. Such a transcellular network based on L1 and ankyrin is analogous, in a formal sense, to the epidermis, where keratinocytes are connected through desmosomes on their outer surfaces, and desmosomes are interconnected via intermediate filaments. The consequence in both cases is a mechanical buffering system capable of distributing shear stresses throughout the tissue. It is of interest in this regard that the brains of ankyrinB (−/−) are much more fragile and susceptible to damage during dissection compared with brains of normal mice (data not shown).

AnkyrinB and L1 are coexpressed and may interact at sites in addition to premyelinated axons. Unmyelinated axons in the molecular layer of the adult cerebellum express 440-kD ankyrinB as well as L1 (Kunimoto et al., 1991; Martini, 1994; Kunimoto, 1995). 220-kD ankyrinB is expressed in glial cells, including Schwann cells and astrocytes (Kordeli et al., 1990; Kunimoto et al., 1991; Kunimoto, 1995), and could associate with alternatively spliced forms of L1 present in these cells (Martini, 1994). L1 is expressed outside of the nervous system in cells including lymphocytes (Kowitz et al., 1992). It will be of interest in future experiments to examine L1 distribution and function in ankyrinB (−/−) mice at sites of axo–glial contact and in the immune system.

Ankyrin and L1 have both been implicated in axon guidance. L1-deficient mice exhibit defects in guidance of corticospinal axons across the pyramidal decussation (Cohen et al., 1998). Ankyrin (Unc-44) mutations in Caenorhabditis elegans result in abnormal axon outgrowth and guidance (Otsuka et al., 1995). It is not clear from the current data if ankyrinB and L1 collaborate in axonal guidance, or if ankyrinB has any role in axonal guidance. It will be of interest to trace the paths of axons in the optic nerve and corticospinal tracts of ankyrinB (−/−) mice to determine possible defects in decussation and in final arrival at neuronal targets. Consequences of ankyrinB deficiency, including degeneration of the optic nerve, could be secondary to defective targeting of axons to their normal sites accompanied by apoptosis caused by lack of retrograde transport of appropriate factors.

The accumulated L1 mutations characterized in human families provide a valuable database for structure–function studies of the L1 molecule (Fransen et al., 1998a). One missense mutation in the L1 cytoplasmic domain results in conversion of Y1229 to H and clinical disease (Van Camp et al., 1996). The Y1229H mutation of the FIGQY tyrosine is of particular interest with respect to ankyrin. Recent analysis of structural requirements for L1 CAM binding to ankyrin has demonstrated that the FIGQY tyrosine is essential for ankyrin binding (Hortsch et al., 1998; Zhang et al., 1998). Moreover, mutation of the FIGQY tyrosine to histidine abolishes ankyrin binding activity of neurofascin (Zhang et al., 1998). These considerations suggest the possibility that ankyrin binding activity of L1 is also abolished by the Y1229H mutation, and that loss of ankyrin binding results in loss of L1 function in humans.

CAM

cell adhesion molecule

DWM

diffusion-weighted MR microscopy

MR

magnetic resonance

Banson
ML
,
Cofer
GP
,
Black
R
,
Johnson
GA
A probe for specimen magnetic resonance microscopy
Investig Radiol
1992
27
57
164
Bartsch
U
,
Kirchhoff
F
,
Schachner
M
Immunohistological localization of the adhesion molecules L1, N-CAM, and MAG in the developing and adult optic nerve of mice
J Comp Neurol
1989
284
451
462
[PubMed]
Beaulieu
C
,
Allen
PS
An in vitro evaluation of the effects of local magnetic-susceptibility-induced gradients on anisotropic water diffusion in nerve
Magn Res Med
1996
36
39
44
[PubMed]
Bennett
V
,
Lambert
S
,
Davis
JQ
,
Zhang
X
Molecular architecture of the specialized axonal membrane at the node of Ranvier
Soc Gen Physiol
1997
52
107
120
[PubMed]
Bieber
AJ
,
Snow
PM
,
Hortsch
M
,
Patel
NH
,
Jacobs
JR
,
Traquina
ZR
,
Schilling
J
,
Goodman
CS
Drosophilaneuroglian: a member of the immunoglobulin superfamily with extensive homology to the vertebrate neural adhesion molecule L1
Cell
1989
59
447
460
[PubMed]
Brummendorf
T
,
Kenwrick
S
,
Rathjen
FG
Neural cell recognition molecule L1: from cell biology to human hereditary brain malformations
Curr Opin Neurobiol
1998
8
87
97
[PubMed]
Chan
W
,
Kordeli
E
,
Bennett
V
440-kD ankyrin-B: structure of the major developmentally regulated domain and selective localization in unmyelinated axons
J Cell Biol
1993
123
1463
1473
[PubMed]
Cohen
NR
,
Taylor
JS
,
Scott
LB
,
Guillery
RW
,
Soriano
P
,
Furley
AJ
Errors in corticospinal axon guidance in mice lacking the neural cell adhesion molecule L1
Curr Biol
1998
8
26
33
[PubMed]
Dahme
M
,
Bartsch
U
,
Martini
R
,
Anliker
B
,
Schachner
M
,
Mantei
N
Disruption of the mouse L1 gene leads to malformations of the nervous system
Nat Genet
1997
17
346
349
[PubMed]
Davis
JQ
,
Bennett
V
Ankyrin binding activity shared by the neurofascin/L1/NrCAM family of nervous system cell adhesion molecules
J Biol Chem
1994
269
27163
27166
[PubMed]
Davis
JQ
,
Lambert
S
,
Bennett
V
Molecular composition of the node of Ranvier: identification of ankyrin-binding cell adhesion molecules neurofascin (mucin+/third FNIII domain−) and NrCAM at nodal axon segments
J Cell Biol
1996
135
1355
1367
[PubMed]
Dubreuil
RR
,
MacVicar
G
,
Dissannyake
S
,
Liu
C
,
Homer
D
,
Hortsch
M
Neuroglian-mediated cell adhesion induces assembly of the membrane skeleton at cell contact sites
J Cell Biol
1996
133
647
655
[PubMed]
Eber
SW
,
Gonzalez
JM
,
Lux
ML
,
Scarpa
AL
,
Tse
WT
,
Dornwell
M
,
Herbers
J
,
Kugler
W
,
Ozcan
R
,
Pekrun
A
et al
Ankyrin-1 mutations are a major cause of dominant and recessive hereditary spherocytosis
Nat Genet
1996
13
214
218
[PubMed]
Fransen
E
,
Lemmon
V
,
Van Camp
G
,
Vits
L
,
Coucke
P
,
Willems
PJ
CRASH syndrome: clinical spectrum of corpus callosum hypoplasia, retardation, adducted thumbs, spastic paraparesis and hydrocephalus due to mutations in one single gene, L1
Eur J Hum Genet
1995
3
273
284
[PubMed]
Fransen
E
,
Van Camp
G
,
Vits
L
,
Willems
PJ
L1-associated diseases: clinical geneticists divide, molecular geneticists unite
Hum Mol Genet
1997
6
1625
1632
[PubMed]
Fransen
E
,
Van Camp
G
,
D'Hooge
R
,
Vits
L
,
Willems
PJ
Genotype-phenotype correlation in L1 associated diseases
J Med Genet
1998a
35
399
404
[PubMed]
Fransen
E
,
D'Hooge
R
,
Van Camp
G
,
Verhoye
M
,
Sijbers
J
,
Reyniers
E
,
Soriano
P
,
Kamiguchi
H
,
Willemsen
R
,
Koekkoek
SK
et al
L1 knockout mice show dilated ventricles, vermis hypoplasia and impaired exploration patterns
Hum Mol Genet
1998b
7
999
1009
[PubMed]
Garver
TD
,
Ren
Q
,
Tuvia
S
,
Bennett
V
Tyrosine phosphorylation at a site highly conserved in the L1 family of cell adhesion molecules abolishes ankyrin binding and increases lateral mobility of neurofascin
J Cell Biol
1997
137
703
714
[PubMed]
Hassel
B
,
Rathjen
FG
,
Volkmer
H
Organization of the neurofascin gene and analysis of developmentally regulated alternative splicing
J Biol Chem
1997
272
28742
28749
[PubMed]
Heiland
PC
,
Griffith
LS
,
Lange
R
,
Schachner
M
,
Hertlein
B
,
Traub
O
,
Schmitz
B
Tyrosine and serine phosphorylation of the neural cell adhesion molecule L1 is implicated in its oligomannosidic glycan dependent association with NCAM and neurite outgrowth
Eur J Cell Biol
1998
75
97
106
[PubMed]
Hortsch
M
The L1 family of neural cell adhesion molecules: old proteins performing new tricks
Neuron
1996
17
587
593
[PubMed]
Hortsch
M
,
Homer
D
,
Malhotra
JD
,
Chang
S
,
Frankel
J
,
Jefford
G
,
Dubreuil
RR
Structural requirements for outside-in and inside-out signaling by Drosophilaneuroglian, a member of the L1 family of cell adhesion molecules
J Cell Biol
1998
142
251
261
[PubMed]
Hu
R-J
,
Moorthy
S
,
Bennett
V
Expression of functional domains of β-G spectrin disrupts epithelial morphology in cultured cells
J Cell Biol
1995
128
1069
1080
[PubMed]
Johnson
GA
,
Benveniste
H
,
Black
RD
,
Hedlund
LW
,
Maronpot
RR
,
Smith
BR
Histology by magnetic resonance microscopy
Magn Reson
1993
9
1
30
Joosten
EA
,
Gribnau
AA
Immunocytochemical localization of cell adhesion molecule L1 in developing rat pyramidal tract
Neurosci Lett
1989
100
94
98
[PubMed]
Jouet
M
,
Rosenthal
A
,
Armstrong
G
,
MacFarlane
J
,
Stevenson
R
,
Paterson
J
,
Metzenberg
A
,
Ionasescu
V
,
Temple
K
,
Kenwrick
S
X-linked spastic paraplegia (SPG1), MASA syndrome and X-linked hydrocephalus result from mutations in the L1 gene
Nat Genet
1994
7
402
407
[PubMed]
Koller
BH
,
Kim
HS
,
Latour
AM
,
Brigman
K
,
Boucher
RC
Jr
,
Scambler
P
,
Wainwright
B
,
Smithies
O
Toward an animal model of cystic fibrosis: targeted interruption of exon 10 of the cystic fibrosis transmembrane regulator gene in embryonic stem cells
Proc Natl Acad Sci USA
1991
88
10730
10734
[PubMed]
Kordeli
E
,
Lambert
S
,
Bennett
V
Ankyrin-G. A new ankyrin gene with neural-specific isoforms localized at the axonal initial segment and node of Ranvier
J Biol Chem
1995
270
2352
2359
[PubMed]
Kowitz
A
,
Kadmon
G
,
Eckert
M
,
Schirrmacher
V
,
Schachner
M
,
Altevogt
P
Expression and function of the neural cell adhesion molecule L1 in mouse leukocytes
Eur J Immunol
1992
22
1199
1205
[PubMed]
Kunimoto
M
A neuron-specific isoform of brain ankyrin, 440-kD ankyrin-B, is targeted to the axons of rat cerebellar neurons
J Cell Biol
1995
131
1821
1829
[PubMed]
Kunimoto
M
,
Otto
E
,
Bennett
V
A new 440-kD isoform is the major ankyrin in neonatal rat brain
J Cell Biol
1991
115
1319
1331
[PubMed]
Lambert
S
,
Davis
JQ
,
Bennett
V
Morphogenesis of the node of Ranvier: co-clusters of ankyrin and ankyrin-binding integral proteins define early developmental intermediates
J Neurosci
1997
17
7025
7036
[PubMed]
Mansour
SL
,
Thomas
KR
,
Capecchi
MR
Disruption of the proto-oncogene int-2 in mouse embryo-derived stem cells: a general strategy for targeting mutations to non-selectable genes
Nature
1988
336
348
352
[PubMed]
Martini
R
Expression and functional roles of neural cell surface molecules and extracellular matrix components during development and regeneration of peripheral nerves
J Neurocytol
1994
23
1
28
[PubMed]
Matsumura
K
,
Ervasti
JM
,
Ohlendieck
K
,
Kahl
SD
,
Campbell
KP
Association of dystrophin-related protein with dystrophin-associated proteins in mdxmouse muscle
Nature
1992
360
588
591
[PubMed]
Michaely
P
,
Bennett
V
The ANK repeats of erythrocyte ankyrin form two distinct but cooperative binding sites for the erythrocyte anion exchanger
J Biol Chem
1995a
270
22050
22057
[PubMed]
Michaely
P
,
Bennett
V
Mechanism for binding site diversity on ankyrin: comparison of binding sites on ankyrin for neurofascin and the Cl−/ HCO3−anion exchanger
J Biol Chem
1995b
270
31298
31302
[PubMed]
Otsuka
AJ
,
Franco
R
,
Yang
B
,
Shim
KH
,
Tang
LZ
,
Zhang
YY
,
Boontrakulpoontawee
P
,
Jeyaprakash
A
,
Hedgecock
E
,
Wheaton
VI
,
Sobery
A
An ankyrin-related gene (unc-44) is necessary for proper axonal guidance in Caenorhabditis elegans.
J Cell Biol
1995
129
1081
1092
[PubMed]
Otto
E
,
Kunimoto
M
,
McLaughlin
T
,
Bennett
V
Isolation and characterization of cDNAs encoding human brain ankyrins reveal a family of alternatively spliced genes
J Cell Biol
1991
114
241
253
[PubMed]
Peters
LL
,
Lux
SE
Ankyrins: structure and function in normal cells and hereditary spherocytes
Semin Hematol
1993
30
85
118
[PubMed]
Rosenthal
A
,
Jouet
M
,
Kenwrick
S
Aberrant splicing of neural cell adhesion molecule L1 mRNA in a family with X-linked hydrocephalus
Nat Genet
1992
2
107
112
[PubMed]
Sambrook, J., E.F. Fritsch, and T. Maniatis. 1989. Molecular Cloning: A Laboratory Manual. 2nd edition. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY. 545 pp.
Sonderegger
P
,
Rathjen
FG
Regulation of axonal growth in the vertebrate nervous system by interactions between glycoproteins belonging to two subgroups of the immunoglobulin superfamily
J Cell Biol
1992
119
1387
1394
[PubMed]
Tuvia
S
,
Garver
TD
,
Bennett
V
The phosphorylation state of the FIGQY tyrosine of neurofascin determines ankyrin-binding activity and patterns of cell segregation
Proc Natl Acad Sci USA
1997
94
12957
12962
[PubMed]
Van Camp
G
,
Fransen
E
,
Vits
L
,
Raes
G
,
Willems
PJ
A locus-specific mutation database for the neural cell adhesion molecule L1CAM (Xq28)
Hum Mut
1996
8
391
[PubMed]
Walsh
FS
,
Doherty
P
Neural cell adhesion molecules of the immunoglobulin superfamily: role in axon growth and guidance
Annu Rev Cell Dev Biol
1997
13
425
456
[PubMed]
Wong
EV
,
Kenwrick
S
,
Willems
P
,
Lemmon
V
Mutations in the cell adhesion molecule L1 cause mental retardation
Trends Neurosci
1995
18
168
172
[PubMed]
Zhang
X
,
Davis
J
,
Carpenter
S
,
Bennett
V
Structural requirements for association of neurofascin with ankyrin
J Biol Chem
1998
273
30785
30794
[PubMed]
Zisch
AH
,
Stallcup
WB
,
Chong
LD
,
Dahlin-Huppe
K
,
Voshol
J
,
Schachner
M
,
Pasquale
EB
Tyrosine phosphorylation of L1 family adhesion molecules: implication of the Eph kinase Cek5
J Neurosci
1997
47
655
665

Dr. Beverly Koller is gratefully acknowledged for performing ES cell manipulations as well as advice in design of the targeting construct. Susan Hester of the Duke Microscopy facility performed electron microscopy. Scott Carpenter is acknowledged for technical assistance.

MRM was performed at the Duke Center for In Vivo Microscopy, National Institutes of Health (NIH) National Center for Research Resources No. P41RR05959. This study was funded in part by a grant from the NIH.

Paula Scotland and Daixing Zhou each made major contributions.

Address all correspondence to Vann Bennett, Box 3892, Duke University Medical Center, Durham, NC 27710. Tel.: (919) 684-3538. Fax: (919) 684-3590. E-mail: benne012@mc.duke.edu