Comprehensive Physiology Wiley Online Library

Of Membrane Stability and Mosaics: The Spectrin Cytoskeleton

Full Article on Wiley Online Library



Abstract

The sections in this article are:

1 The Red Cell Membrane Skeleton
2 How Does the Spectrin Membrane Skeleton Stabilize the Red Cell?
3 The Trilayer Couple—Spectrin as A Membrane Organizer
4 Components of the Erythrocyte Membrane Skeleton
4.1 Spectrin
4.2 Actin
4.3 Ankyrin
4.4 Protein 4.1
4.5 Adducin
4.6 Dematin (Protein 4.9)
4.7 Pallidin (Protein 4.2)
4.8 p55 (an Erythrocyte Membrane‐Associated Guanylate Kinase)
4.9 Stomatin
4.10 Tropomyosin and Tropomodulin
4.11 Dynamin
4.12 Interactions with Phospholipids
5 The Spectrin Skeleton of Non‐Erythroid Cells
5.1 Spatial and Temporal Polarization
6 Proteins Interacting with Spectrin in Non‐Erythroid Cells
6.1 Cytoskeletal Elements
6.2 Adhesion Proteins
7 Evolving Concepts
8 Conclusions: The Linked Mosaic Model
Figure 1. Figure 1.

The erythrocyte membrane skeleton A. Model of erythrocyte membrane skeleton. The spectrin‐actin subcortical lattice is anchored to the membrane via at least three different integral membrane proteins, as well as by at least one direct attachment. Which of these attachments drives membrane assembly is uncertain, although current evidence favors ankyrin‐independent linkages. B. Major proteins of the red cell membrane and the cortical membrane skeleton. SDS‐PAGE analysis of purified human red cell membranes yields approximately 16 major protein bands (lane 1). Extraction of such membranes with 0.1 mM EDTA at pH 8 or above liberates spectrin and actin and a fraction of the cell's protein 4.1, dematin, and adducin (lane 2). The membranes remaining after the cortical skeleton is removed retain all of the integral proteins and many of the linker proteins (lane 3). The nomenclature according to their order on SDS‐PAGE gels as well as their assigned names is given.

Figure 2. Figure 2.

Four views of the spectrin membrane skeleton. A. The erythrocyte cytoskeleton in situ. This sample was prepared from paraformaldehyde‐fixed ghosts attached to coated coverslips and sheered to expose the skeleton, followed by quick‐freeze, deepetch, rotary‐replication (QFDERR). The arrows point to unquenched water left in the sample. The average filament length is 29–37 nm, which is about a third of the extended length of the spectrin heterodimer. This information together with immunogold labeling has suggested that in situ the skeleton is highly compact, with many side‐to‐side associations of spectrin as well as a large fraction of spectrin hexamers and octamers as well as tetramers 518. B. A view of the spread isolated skeleton prepared by extraction of ghosts with 2.5% Triton X‐100, followed by spreading onto a formvar film followed by QFDERR. After this treatment, the compact skeleton seen in situ is highly expanded. Small arrows point to apparent actin sites; large arrows point to spectrin hexamer junctions. Arrowheads point to laterally associated filaments. The double arrowhead designates an ankyrin‐AE1 complex 519. (Panels A and B were adapted with permission from the references cited.) C. Quick‐freeze deepetch replica of the inner face of a red cell membrane in situ that has been extracted with NP‐40 and NaCl to remove most of the ankyrin and protein 4.1, exposing the network of intertwined spectrin and 67 nm actin nuggets. Compared with (A), this view depicts a much looser skeleton. It is likely that the salt and detergent treatment used in this preparation has loosened the side‐to‐side associations of spectrin as well as removed many associated proteins. D. Image of an expanded skeleton prepared by negative staining. The stretched spectrin filaments (tetramers) predominate in these preparations, although hexamers are seen 279. The preparation shown here was used to determine by Fourier analysis the supercoiling of the spectrin heterotetramer (see Fig. 4).

Micrograph provided compliments of Dr. J. Heuser and reprinted with permission from Coleman et al. 91 Adapted with permission from McGough and Josephs 323
Figure 3. Figure 3.

Alterations in the spectrin skeleton alter the shape and stability of erythrocytes. Scanning electron microscopic images: A. Normal red cell. B. Defects in ankyrin or AE1 often lead to spherocytosis. C. Inherited defects in the self‐association of spectrin often lead to hereditary elliptocytosis. D. Increased intracellular Ca2+, lowered pH, or ATP depletion induce echinocytic transformations.

Figure 4. Figure 4.

Fourier filtered images of the spectrin heterotetramer in situ. These images were derived from negative stained electron micrographs of the stretched spectrin cortical skeleton such as the micrograph shown in Figure 2D. This analysis revealed a coiled‐coil quaternary conformation of spectrin in situ. Note that the spectrin heterotetramer can be linearly deformed, accounting for its proposed “spring‐like” properties. Shown are three filtered images demonstrating how the pitch and length of the spectrin helix change with length. Each turn of the helix is estimated to contain four spectrin repeat units. Upon extraction from the membrane, spectrin heterodimers and heterotetramers assume a much more open and loose conformation (cf. Fig. 12).

Adapted with permission from McGough and Josephs 323
Figure 5. Figure 5.

Effects of embedded proteins as predicted by the bilayer couple hypothesis. Integral membrane proteins can have a major effect on the shape and stability of the bilayer. Control by a cortical skeleton ameliorates this potential effect. A. Cylindrical embedded proteins can impart curvature to the membrane by expanding one face preferentially. They thus will tend to cluster in regions of curvature that reduce their net interaction energy with the bilayer. A cortical skeleton aids in maintaining a homogeneous distribution of these molecules, and thereby contributes to membrane stability. C. Released of the constraints of a cortical skeleton, asymmetric embedded proteins can induce acute membrane curvatures and instability.

Figure 6. Figure 6.

By controlling the distribution of embedded proteins, the cortical skeleton controls cell shape. The equilibrium erythrocyte shape and the lateral density of the embedded proteins and cortical skeleton has been calculated (A) without the stabilizing effect of the skeleton or (B) with a cortical skeleton. In these calculations, the relative volume was held constant at 0.6, and the relative surface areas of the two leaves of the bilayer was set at 1.0465; cs = 4; k = 0.005; p = 500. Note that in case A, the equilibrium shape is elliptical, whereas in B the equilibrium shape is axisymmetric and biconcave. Also note the markedly inhomogeneous distribution of protein in the case without skeletal stabilization. Views are presented along the minor axis; the ellipse semiaxis ratio in A is 1.18.

Adapted with permission from Svetina et al. 493
Figure 7. Figure 7.

Increasing the number of embedded molecules may induce instability in the cell membrane. The relative membrane free energy is calculated as a function of the number of embedded proteins for two different shapes, the discocyte and a shape consisting of a spherical mother cell and two spherical daughter vesicles. For both cells, v = 0.6; kr/kc = 3; cs = 2, k = 0.005. For the discocyte, Δao = 1.038; for the mother cell and vesicles, net Δa = 1.7105, rmother = 0.6985, and rdaughter = 0.5060. Note that as the number of embedded proteins (p) rises, the cell tends toward vesiculation, a consequence resisted by the presence of a cortical skeleton (cf. Fig. 6).

Adapted with permission from Svetina et al. 493
Figure 8. Figure 8.

Spectrin typically is a multifunctional heterodimer that self‐associates. The functional unit is most often an αβ‐heterodimer. Each subunit contains non‐homologous NH2‐ and COOH‐terminal domains (domains I and III, respectively), joined by a central domain (domain II) composed of multiple homologous repeats of ∼106 residues each. Within the molecules are many sites that interact with other proteins or regulator molecules.

Figure 9. Figure 9.

The spectrins are the prototypical members of a gene superfamily that includes α‐actinin and dystrophin. The derived protein sequences of the repetitive unit found in the central domain of the four documented spectrin genes, α‐actinin, and dystrophin were first compared within each protein to arrive at a consensus repeat structure for each protein. These consensus repeats were then compared with each other to estimate their overall degree of relatedness. A. A relatedness tree in which the aggregate distance separating the different proteins is proportional to their similarity. This analysis indicates that the superfamily has three major branches, with the spectrins being most related to each other, then to α‐actinin, and only distantly to dystropin. B. Sequence comparison of the consensus repeat units for each protein in the superfamily. Note the strong conservation of the trp at position 18 in the repetitive unit. A similar residue occurs at position 21 in dystrophin, although the alignment algorithm does not score this in this comparison.

Morrow and Rimm, unpublished obervations
Figure 10. Figure 10.

Spectrin contains many regions of sequence divergence associated with specialized function. The sequences of either (A) JI spectrin or (B) βI spectrin were compared with themselves by “dot blot” analysis. The parallel lines spaced at 106 residues are indicative of the repetitive structure of spectrin. Note the nonrepetitive structure in domains I and III of both spectrins, as well as the many regions of divergence even with domain II. Sites marked by * indicate the approximate location of several mutations that cause hemolytic disease (reviewed in 301). Note that all known functional domains (labeled), as well as disease causing mutations, are located in regions of sequence divergence. The role of many such regions remains undetermined, although it is likely that they may harbor activities necessary for some of spectrin's myriad functions (see Table 2).

Figure 11. Figure 11.

The repetitive unit of spectrin is a stable triple helix, capable of refolding spontaneously. A. Spectrin's characteristic pattern of resistance to proteolytic digestion by trypsin is recovered spontaneously after denaturation with urea. This resistance derives from spectrin's secondary and tertiary structure, which limits access of the protease to over 300 potential sites in each subunit 473. In these experiments, carried out by Dr. William Knowles and Vincent Marchesi at Yale, αI/βI spectrin was denatured with increasing amounts of urea; then, for half the samples, the urea was diluted or dialyzed away, and the protein subjected to 60 min of trypsin proteolysis at 0°C. Note the spontaneous recovery of its resistance to trypsin, indicative that the protein can faithfully recover at least most aspects of its secondary and tertiary structure after denaturation (Drs. Knowles and Marchesi, personal communication). B. The putative structure of two spectrin repeat units as derived from the crystal coordinates of Drosophila αI spectrin's 14th repeat unit. It is probable that helix C joins helix A with minimal discontinuity.

Figure 12. Figure 12.

The self‐association of αβ‐spectrin heterodimers to form tetramers is mediated by domain 1 of α‐spectrin and repeat 17 of β‐spectrin. A. Spectrin αβ heterodimers undergo a concentration‐dependent self‐association to form tetramers and larger oligomers. This process can be most clearly demonstrated by non‐denaturing gel electrophoresis, in which each oligomeric species migrates as a separate band. The different forms of spectrin can also be demonstrated by electron microscopy after rotary shadowing. As the concentration of spectrin incubated in vitro is diminished from 24 mg/ml (left lane) to 12, 6, and 3 mg/ml respectively, spectrin's state of self‐association changes from larger oligomeric species towards the dimer. Oligomerization to species beyond the tetramer is a property more pronounced in αI/βI spectrin than in αII/βII spectrin.

Adapted with permission from Morrow and Marchesi 356. B. The joining of two αβ‐heterodimers to form a spectrin tetramer involves the paired association of the amino‐terminal extension of α‐spectrin that is homologous to helix C with the incomplete repeat 17 of β‐spectrin. This association process thus essentially recreates a complete spectrin triple helical repeat motif (cf. Fig. 11) 235
Figure 13. Figure 13.

Domain maps of spectrin generated by trypsin digestion. αI/βI spectrin extracted from erythrocytes was subjected to mild trypsin digestion at 0°C, cleaving the protein into a reproducible number of well‐characterized large fragments. A. The alignment of the various tryptic fragments, as determined by high‐resolution peptide mapping 473. Also shown is the alignment of some of the fragments generated by 2‐nitro‐5‐thiocyanobenzoic acid (NTCB) digestion of αI/βI spectrin. B. The pattern following limited trypsin digestion. Peptide fragments are resolved by two‐dimensional IEF‐SDS‐PAGE analysis, and visualized by Coomassie blue. The pattern that results is of limited complexity, facilitating the identification of the sites of functional domains, post‐translational modifications, and inherited mutations in this very large protein.

Adapted with permission from Speicher et al. 473
Figure 14. Figure 14.

The nomenclature of the spectrins. A. There are four known genes for spectrin; these encode spectrins αI, αII, βI, and βII. Multiple isoforms arise from each of these genes except αI (at least none have so far been discovered). The various isoforms for each spectrin are annotated by the addition of a Σ followed by arabic numbers. Non‐mammalian spectrins are referenced to this nomenclature by homology or, if not homologous, by special notation 307,543. B. Alternative transcripts in αII spectrin. Three sites of alternative mRNA splicing have been identified in αII spectrin 76,77,346. Note: numbering is based on the lung fibroblast sequence, which does not contain insert 2.

Figure 15. Figure 15.

The calpain cleavage of αIIβII spectrin (fodrin) activates its susceptibility to regulation by Ca2+ and calmodulin. Diagram depicting the synergistic action of calpain and calmodulin on the self‐association and actin‐binding properties of spectrin. A. Intact αIIβII spectrin binds F‐actin and cross‐links actin filaments by forming a stable heterotetramer. Calcium and calmodulin are without effect on these interactions. B. Proteolysis of the α subunit, creating the breakdown products α‐bdp1 and α‐bdp2, has no immediate effect on the actin binding of spectrin or its self‐associative properties, and the spectrin molecule remains assembled and tetrameric under native conditions. The exact (calculated) size of the major proteolytic products of αII spectrin, α‐bdp1 and α‐bdp2, are 135,832 and 148,495 Da respectively. C. The binding of calmodulin to α‐cleaved spectrin dissociates the tetramer to dimers and reduces its ability to bind actin. This process is rapidly reversible and requires calcium. (Note: this is the only physiologic condition yet identified that will lead to dissociation of αIIβII spectrin heterotetramers.) D. Continued action of calpain in the presence of calmodulin leads to cleavage of the β‐subunit, generating β‐bdp1 and β‐bdp2. This cleavage causes dissociation of the heterodimer and irreversible loss of actin binding and crosslinking properties. The placement of these fragments relative to the intact molecule is as indicated. Note that the two fragments comprising the actin‐binding end of the molecule remain non‐covalently associated, while the other fragments dissociate.

Adapted from Harris and Morrow 187 with permission
Figure 16. Figure 16.

Tripartite domain structure of ankyrin. Ankyrin isolated from erythrocytes (ANKI or AnkR) contains three independently folded domains. The ankyrin repeats domain (ARD) consists of 24 nonidentical repeats of 33 residues each. The ARD motif is a common feature of many proteins. This domain in ankyrin is responsible for binding to AE1, Na,K‐ATPase, and typically other membrane proteins. The central domain is involved with ankyrin's interactions with spectrin. The region of the most critical residues are shaded black. The least well understood of ankyrin's domains is the putative regulator domain. Alternative transcripts involving this domain alter ankyrin's other interactions. Also shown is the derived structure of a recently cloned ankyrin with wide tissue distribution (Ank G, ANK3, ANKIII) 117,250,400.

Figure 17. Figure 17.

Domain structure of protein 4.1. Many isoforms of protein 4.1 exist, all arising by a complex pattern of alternative mRNA spicing over a region encompassing 23 exons 94. The functional domain structure of the erythroid protein is shown here, based on several studies 75,93,94,98,203,204,205,272,284,391. The 10 kDa spectrin‐actin binding region appears to be an erythroid‐specific spiciform, suggesting that this activity may not be fundamental to protein 4.1's function. The other members of the 4.1 family (ezrin, moesin, radixin) share their strongest homologies in the 30 kD NH2‐terminal region. The predicted secondary structure is derived from the algorithms of Garnier et al. 157.

Figure 18. Figure 18.

Adducin isoforms and alternative transcripts. Multiple transcripts of both α‐adducin and β‐adducin have been identified. All of these so far involve alternative transcripts involving the COOH‐terminal third of the molecule (aka the “tail region”). This region is very proteolytically sensitive.

Figure 19. Figure 19.

Stomatin contains a hydrophobic 28‐residue sequence thought to intercalate into the bilayer. A schematic view of stomatin suggests a typical transmembrane protein. However, both the NH2‐and COOH‐ termini are accessible from the cytoplasmic side of the membrane 199,442. Therefore, stomatin may assume an unusual structure with a putative hairpin loop inserting into the bilayer. Stomatin appears to regulate Na+/K+ leaks, and binds to adducin 465.

Figure 20. Figure 20.

Spectrin in non‐erythroid cells is often highly organized. A. Rat myotubes after extraction with saponin and stained with rhodamine‐bungarotoxin, which stains the acetylcholine receptor (AChR) clusters. B. Same preparation stained with Mab VIIF7, which reacts with βI spectrin. Note the coincident distribution, indicating that this spectrin is codistributed with the AChR,. C. βIΣ2 spectrin is concentrated at the postsynaptic density in the molecular layer of the rat cerebellum. Immunoperoxidase labeling EM. Bar = 0.5. μm D. Immunofluorescent image of MDCK cell stained with an antibody to Ank G119(ANKIIΣ4). The punctate eccentric cytoplasmic staining adjacent to the nucleus marks a form of βI spectrin associated with the Golgi apparatus 20,117.

Adapted with permission from Bloch and Morrow 41 Adapted with permission from 307
Figure 21. Figure 21.

A hypothetical model of the cadherin‐based transmembrane adhesion complex and its relationship to the associated cortical cytoskeleton. The Greek letters α and β indicate the corresponding catenins. Although there is evidence that α‐catenin can bind both spectrin 286 and actin 428 and that is exists as a dimer, the linkages shown are hypothetical. It is not known if catenin/spectrin and catenin/actin interactions can occur simultaneously. It is possible that α‐catenin not only links the cadherin complex to the cytoskeleton but also participates in actin bundling in other parts of the cell or in other cell types. There is no good evidence for a direct interaction between E‐cadherin and spectrin, but they co‐localize and are suspected to be joined by α‐catenin. The interactions between β‐ and α‐catenin are presented as an example of how β‐catenin (a member of the arm family) binds directly to both E‐cadherin and α‐catenin. It is not yet known which, if any, other members of the arm family can participate in this interaction. Plakoglobin, previously known as γ‐catenin, may participate, but cannot bind the same cadherin molecule (although theoretically both could bind a cadherin dimer). α‐Actinin, another member of the spectrin superfamily (see Fig. 9), has also been implicated in this complex, but its exact role remains uncertain; and hence it is not included in this figure. Both p120 (CAS) and Ptpaseμ also bind to cadherin, but neither appears to bind in same region as β‐catenin or to interact directly with spectrin.

Figure 22. Figure 22.

The linked mosaic model of spectrin action. The fundamental role of the spectrin skeleton is to control lateral order in the plane of the membrane. Through its capacity to bind multiple ligands selectively, and to self‐associate through hetero‐and homo‐typic interactions, end‐on and side to side, spectrin alone can form ordered arrays of mosaics of limited size. Associated with these nascent arrays are various embedded and soluble proteins. It is hypothesized that the spectrin mosaics with their associated proteins are joined by linking interactions, most commonly involving actin or other filament systems, to form larger arrays. Collectively, these linked mosaics serve to control the lateral organization of integral membrane proteins, and to recruit selected cytoplasmic components (e.g., signal transduction molecules, kinases) in a way that enhances receptor efficiencies, creates signal transduction systems, and stabilizes the membrane against spontaneous and deleterious changes in the distribution of its integral membrane proteins. This proposed role applies with equal validity to internal as well as surface membranes. Perhaps for this reason spectrin and ankyrin are also associated with internal membranes.



Figure 1.

The erythrocyte membrane skeleton A. Model of erythrocyte membrane skeleton. The spectrin‐actin subcortical lattice is anchored to the membrane via at least three different integral membrane proteins, as well as by at least one direct attachment. Which of these attachments drives membrane assembly is uncertain, although current evidence favors ankyrin‐independent linkages. B. Major proteins of the red cell membrane and the cortical membrane skeleton. SDS‐PAGE analysis of purified human red cell membranes yields approximately 16 major protein bands (lane 1). Extraction of such membranes with 0.1 mM EDTA at pH 8 or above liberates spectrin and actin and a fraction of the cell's protein 4.1, dematin, and adducin (lane 2). The membranes remaining after the cortical skeleton is removed retain all of the integral proteins and many of the linker proteins (lane 3). The nomenclature according to their order on SDS‐PAGE gels as well as their assigned names is given.



Figure 2.

Four views of the spectrin membrane skeleton. A. The erythrocyte cytoskeleton in situ. This sample was prepared from paraformaldehyde‐fixed ghosts attached to coated coverslips and sheered to expose the skeleton, followed by quick‐freeze, deepetch, rotary‐replication (QFDERR). The arrows point to unquenched water left in the sample. The average filament length is 29–37 nm, which is about a third of the extended length of the spectrin heterodimer. This information together with immunogold labeling has suggested that in situ the skeleton is highly compact, with many side‐to‐side associations of spectrin as well as a large fraction of spectrin hexamers and octamers as well as tetramers 518. B. A view of the spread isolated skeleton prepared by extraction of ghosts with 2.5% Triton X‐100, followed by spreading onto a formvar film followed by QFDERR. After this treatment, the compact skeleton seen in situ is highly expanded. Small arrows point to apparent actin sites; large arrows point to spectrin hexamer junctions. Arrowheads point to laterally associated filaments. The double arrowhead designates an ankyrin‐AE1 complex 519. (Panels A and B were adapted with permission from the references cited.) C. Quick‐freeze deepetch replica of the inner face of a red cell membrane in situ that has been extracted with NP‐40 and NaCl to remove most of the ankyrin and protein 4.1, exposing the network of intertwined spectrin and 67 nm actin nuggets. Compared with (A), this view depicts a much looser skeleton. It is likely that the salt and detergent treatment used in this preparation has loosened the side‐to‐side associations of spectrin as well as removed many associated proteins. D. Image of an expanded skeleton prepared by negative staining. The stretched spectrin filaments (tetramers) predominate in these preparations, although hexamers are seen 279. The preparation shown here was used to determine by Fourier analysis the supercoiling of the spectrin heterotetramer (see Fig. 4).

Micrograph provided compliments of Dr. J. Heuser and reprinted with permission from Coleman et al. 91 Adapted with permission from McGough and Josephs 323


Figure 3.

Alterations in the spectrin skeleton alter the shape and stability of erythrocytes. Scanning electron microscopic images: A. Normal red cell. B. Defects in ankyrin or AE1 often lead to spherocytosis. C. Inherited defects in the self‐association of spectrin often lead to hereditary elliptocytosis. D. Increased intracellular Ca2+, lowered pH, or ATP depletion induce echinocytic transformations.



Figure 4.

Fourier filtered images of the spectrin heterotetramer in situ. These images were derived from negative stained electron micrographs of the stretched spectrin cortical skeleton such as the micrograph shown in Figure 2D. This analysis revealed a coiled‐coil quaternary conformation of spectrin in situ. Note that the spectrin heterotetramer can be linearly deformed, accounting for its proposed “spring‐like” properties. Shown are three filtered images demonstrating how the pitch and length of the spectrin helix change with length. Each turn of the helix is estimated to contain four spectrin repeat units. Upon extraction from the membrane, spectrin heterodimers and heterotetramers assume a much more open and loose conformation (cf. Fig. 12).

Adapted with permission from McGough and Josephs 323


Figure 5.

Effects of embedded proteins as predicted by the bilayer couple hypothesis. Integral membrane proteins can have a major effect on the shape and stability of the bilayer. Control by a cortical skeleton ameliorates this potential effect. A. Cylindrical embedded proteins can impart curvature to the membrane by expanding one face preferentially. They thus will tend to cluster in regions of curvature that reduce their net interaction energy with the bilayer. A cortical skeleton aids in maintaining a homogeneous distribution of these molecules, and thereby contributes to membrane stability. C. Released of the constraints of a cortical skeleton, asymmetric embedded proteins can induce acute membrane curvatures and instability.



Figure 6.

By controlling the distribution of embedded proteins, the cortical skeleton controls cell shape. The equilibrium erythrocyte shape and the lateral density of the embedded proteins and cortical skeleton has been calculated (A) without the stabilizing effect of the skeleton or (B) with a cortical skeleton. In these calculations, the relative volume was held constant at 0.6, and the relative surface areas of the two leaves of the bilayer was set at 1.0465; cs = 4; k = 0.005; p = 500. Note that in case A, the equilibrium shape is elliptical, whereas in B the equilibrium shape is axisymmetric and biconcave. Also note the markedly inhomogeneous distribution of protein in the case without skeletal stabilization. Views are presented along the minor axis; the ellipse semiaxis ratio in A is 1.18.

Adapted with permission from Svetina et al. 493


Figure 7.

Increasing the number of embedded molecules may induce instability in the cell membrane. The relative membrane free energy is calculated as a function of the number of embedded proteins for two different shapes, the discocyte and a shape consisting of a spherical mother cell and two spherical daughter vesicles. For both cells, v = 0.6; kr/kc = 3; cs = 2, k = 0.005. For the discocyte, Δao = 1.038; for the mother cell and vesicles, net Δa = 1.7105, rmother = 0.6985, and rdaughter = 0.5060. Note that as the number of embedded proteins (p) rises, the cell tends toward vesiculation, a consequence resisted by the presence of a cortical skeleton (cf. Fig. 6).

Adapted with permission from Svetina et al. 493


Figure 8.

Spectrin typically is a multifunctional heterodimer that self‐associates. The functional unit is most often an αβ‐heterodimer. Each subunit contains non‐homologous NH2‐ and COOH‐terminal domains (domains I and III, respectively), joined by a central domain (domain II) composed of multiple homologous repeats of ∼106 residues each. Within the molecules are many sites that interact with other proteins or regulator molecules.



Figure 9.

The spectrins are the prototypical members of a gene superfamily that includes α‐actinin and dystrophin. The derived protein sequences of the repetitive unit found in the central domain of the four documented spectrin genes, α‐actinin, and dystrophin were first compared within each protein to arrive at a consensus repeat structure for each protein. These consensus repeats were then compared with each other to estimate their overall degree of relatedness. A. A relatedness tree in which the aggregate distance separating the different proteins is proportional to their similarity. This analysis indicates that the superfamily has three major branches, with the spectrins being most related to each other, then to α‐actinin, and only distantly to dystropin. B. Sequence comparison of the consensus repeat units for each protein in the superfamily. Note the strong conservation of the trp at position 18 in the repetitive unit. A similar residue occurs at position 21 in dystrophin, although the alignment algorithm does not score this in this comparison.

Morrow and Rimm, unpublished obervations


Figure 10.

Spectrin contains many regions of sequence divergence associated with specialized function. The sequences of either (A) JI spectrin or (B) βI spectrin were compared with themselves by “dot blot” analysis. The parallel lines spaced at 106 residues are indicative of the repetitive structure of spectrin. Note the nonrepetitive structure in domains I and III of both spectrins, as well as the many regions of divergence even with domain II. Sites marked by * indicate the approximate location of several mutations that cause hemolytic disease (reviewed in 301). Note that all known functional domains (labeled), as well as disease causing mutations, are located in regions of sequence divergence. The role of many such regions remains undetermined, although it is likely that they may harbor activities necessary for some of spectrin's myriad functions (see Table 2).



Figure 11.

The repetitive unit of spectrin is a stable triple helix, capable of refolding spontaneously. A. Spectrin's characteristic pattern of resistance to proteolytic digestion by trypsin is recovered spontaneously after denaturation with urea. This resistance derives from spectrin's secondary and tertiary structure, which limits access of the protease to over 300 potential sites in each subunit 473. In these experiments, carried out by Dr. William Knowles and Vincent Marchesi at Yale, αI/βI spectrin was denatured with increasing amounts of urea; then, for half the samples, the urea was diluted or dialyzed away, and the protein subjected to 60 min of trypsin proteolysis at 0°C. Note the spontaneous recovery of its resistance to trypsin, indicative that the protein can faithfully recover at least most aspects of its secondary and tertiary structure after denaturation (Drs. Knowles and Marchesi, personal communication). B. The putative structure of two spectrin repeat units as derived from the crystal coordinates of Drosophila αI spectrin's 14th repeat unit. It is probable that helix C joins helix A with minimal discontinuity.



Figure 12.

The self‐association of αβ‐spectrin heterodimers to form tetramers is mediated by domain 1 of α‐spectrin and repeat 17 of β‐spectrin. A. Spectrin αβ heterodimers undergo a concentration‐dependent self‐association to form tetramers and larger oligomers. This process can be most clearly demonstrated by non‐denaturing gel electrophoresis, in which each oligomeric species migrates as a separate band. The different forms of spectrin can also be demonstrated by electron microscopy after rotary shadowing. As the concentration of spectrin incubated in vitro is diminished from 24 mg/ml (left lane) to 12, 6, and 3 mg/ml respectively, spectrin's state of self‐association changes from larger oligomeric species towards the dimer. Oligomerization to species beyond the tetramer is a property more pronounced in αI/βI spectrin than in αII/βII spectrin.

Adapted with permission from Morrow and Marchesi 356. B. The joining of two αβ‐heterodimers to form a spectrin tetramer involves the paired association of the amino‐terminal extension of α‐spectrin that is homologous to helix C with the incomplete repeat 17 of β‐spectrin. This association process thus essentially recreates a complete spectrin triple helical repeat motif (cf. Fig. 11) 235


Figure 13.

Domain maps of spectrin generated by trypsin digestion. αI/βI spectrin extracted from erythrocytes was subjected to mild trypsin digestion at 0°C, cleaving the protein into a reproducible number of well‐characterized large fragments. A. The alignment of the various tryptic fragments, as determined by high‐resolution peptide mapping 473. Also shown is the alignment of some of the fragments generated by 2‐nitro‐5‐thiocyanobenzoic acid (NTCB) digestion of αI/βI spectrin. B. The pattern following limited trypsin digestion. Peptide fragments are resolved by two‐dimensional IEF‐SDS‐PAGE analysis, and visualized by Coomassie blue. The pattern that results is of limited complexity, facilitating the identification of the sites of functional domains, post‐translational modifications, and inherited mutations in this very large protein.

Adapted with permission from Speicher et al. 473


Figure 14.

The nomenclature of the spectrins. A. There are four known genes for spectrin; these encode spectrins αI, αII, βI, and βII. Multiple isoforms arise from each of these genes except αI (at least none have so far been discovered). The various isoforms for each spectrin are annotated by the addition of a Σ followed by arabic numbers. Non‐mammalian spectrins are referenced to this nomenclature by homology or, if not homologous, by special notation 307,543. B. Alternative transcripts in αII spectrin. Three sites of alternative mRNA splicing have been identified in αII spectrin 76,77,346. Note: numbering is based on the lung fibroblast sequence, which does not contain insert 2.



Figure 15.

The calpain cleavage of αIIβII spectrin (fodrin) activates its susceptibility to regulation by Ca2+ and calmodulin. Diagram depicting the synergistic action of calpain and calmodulin on the self‐association and actin‐binding properties of spectrin. A. Intact αIIβII spectrin binds F‐actin and cross‐links actin filaments by forming a stable heterotetramer. Calcium and calmodulin are without effect on these interactions. B. Proteolysis of the α subunit, creating the breakdown products α‐bdp1 and α‐bdp2, has no immediate effect on the actin binding of spectrin or its self‐associative properties, and the spectrin molecule remains assembled and tetrameric under native conditions. The exact (calculated) size of the major proteolytic products of αII spectrin, α‐bdp1 and α‐bdp2, are 135,832 and 148,495 Da respectively. C. The binding of calmodulin to α‐cleaved spectrin dissociates the tetramer to dimers and reduces its ability to bind actin. This process is rapidly reversible and requires calcium. (Note: this is the only physiologic condition yet identified that will lead to dissociation of αIIβII spectrin heterotetramers.) D. Continued action of calpain in the presence of calmodulin leads to cleavage of the β‐subunit, generating β‐bdp1 and β‐bdp2. This cleavage causes dissociation of the heterodimer and irreversible loss of actin binding and crosslinking properties. The placement of these fragments relative to the intact molecule is as indicated. Note that the two fragments comprising the actin‐binding end of the molecule remain non‐covalently associated, while the other fragments dissociate.

Adapted from Harris and Morrow 187 with permission


Figure 16.

Tripartite domain structure of ankyrin. Ankyrin isolated from erythrocytes (ANKI or AnkR) contains three independently folded domains. The ankyrin repeats domain (ARD) consists of 24 nonidentical repeats of 33 residues each. The ARD motif is a common feature of many proteins. This domain in ankyrin is responsible for binding to AE1, Na,K‐ATPase, and typically other membrane proteins. The central domain is involved with ankyrin's interactions with spectrin. The region of the most critical residues are shaded black. The least well understood of ankyrin's domains is the putative regulator domain. Alternative transcripts involving this domain alter ankyrin's other interactions. Also shown is the derived structure of a recently cloned ankyrin with wide tissue distribution (Ank G, ANK3, ANKIII) 117,250,400.



Figure 17.

Domain structure of protein 4.1. Many isoforms of protein 4.1 exist, all arising by a complex pattern of alternative mRNA spicing over a region encompassing 23 exons 94. The functional domain structure of the erythroid protein is shown here, based on several studies 75,93,94,98,203,204,205,272,284,391. The 10 kDa spectrin‐actin binding region appears to be an erythroid‐specific spiciform, suggesting that this activity may not be fundamental to protein 4.1's function. The other members of the 4.1 family (ezrin, moesin, radixin) share their strongest homologies in the 30 kD NH2‐terminal region. The predicted secondary structure is derived from the algorithms of Garnier et al. 157.



Figure 18.

Adducin isoforms and alternative transcripts. Multiple transcripts of both α‐adducin and β‐adducin have been identified. All of these so far involve alternative transcripts involving the COOH‐terminal third of the molecule (aka the “tail region”). This region is very proteolytically sensitive.



Figure 19.

Stomatin contains a hydrophobic 28‐residue sequence thought to intercalate into the bilayer. A schematic view of stomatin suggests a typical transmembrane protein. However, both the NH2‐and COOH‐ termini are accessible from the cytoplasmic side of the membrane 199,442. Therefore, stomatin may assume an unusual structure with a putative hairpin loop inserting into the bilayer. Stomatin appears to regulate Na+/K+ leaks, and binds to adducin 465.



Figure 20.

Spectrin in non‐erythroid cells is often highly organized. A. Rat myotubes after extraction with saponin and stained with rhodamine‐bungarotoxin, which stains the acetylcholine receptor (AChR) clusters. B. Same preparation stained with Mab VIIF7, which reacts with βI spectrin. Note the coincident distribution, indicating that this spectrin is codistributed with the AChR,. C. βIΣ2 spectrin is concentrated at the postsynaptic density in the molecular layer of the rat cerebellum. Immunoperoxidase labeling EM. Bar = 0.5. μm D. Immunofluorescent image of MDCK cell stained with an antibody to Ank G119(ANKIIΣ4). The punctate eccentric cytoplasmic staining adjacent to the nucleus marks a form of βI spectrin associated with the Golgi apparatus 20,117.

Adapted with permission from Bloch and Morrow 41 Adapted with permission from 307


Figure 21.

A hypothetical model of the cadherin‐based transmembrane adhesion complex and its relationship to the associated cortical cytoskeleton. The Greek letters α and β indicate the corresponding catenins. Although there is evidence that α‐catenin can bind both spectrin 286 and actin 428 and that is exists as a dimer, the linkages shown are hypothetical. It is not known if catenin/spectrin and catenin/actin interactions can occur simultaneously. It is possible that α‐catenin not only links the cadherin complex to the cytoskeleton but also participates in actin bundling in other parts of the cell or in other cell types. There is no good evidence for a direct interaction between E‐cadherin and spectrin, but they co‐localize and are suspected to be joined by α‐catenin. The interactions between β‐ and α‐catenin are presented as an example of how β‐catenin (a member of the arm family) binds directly to both E‐cadherin and α‐catenin. It is not yet known which, if any, other members of the arm family can participate in this interaction. Plakoglobin, previously known as γ‐catenin, may participate, but cannot bind the same cadherin molecule (although theoretically both could bind a cadherin dimer). α‐Actinin, another member of the spectrin superfamily (see Fig. 9), has also been implicated in this complex, but its exact role remains uncertain; and hence it is not included in this figure. Both p120 (CAS) and Ptpaseμ also bind to cadherin, but neither appears to bind in same region as β‐catenin or to interact directly with spectrin.



Figure 22.

The linked mosaic model of spectrin action. The fundamental role of the spectrin skeleton is to control lateral order in the plane of the membrane. Through its capacity to bind multiple ligands selectively, and to self‐associate through hetero‐and homo‐typic interactions, end‐on and side to side, spectrin alone can form ordered arrays of mosaics of limited size. Associated with these nascent arrays are various embedded and soluble proteins. It is hypothesized that the spectrin mosaics with their associated proteins are joined by linking interactions, most commonly involving actin or other filament systems, to form larger arrays. Collectively, these linked mosaics serve to control the lateral organization of integral membrane proteins, and to recruit selected cytoplasmic components (e.g., signal transduction molecules, kinases) in a way that enhances receptor efficiencies, creates signal transduction systems, and stabilizes the membrane against spontaneous and deleterious changes in the distribution of its integral membrane proteins. This proposed role applies with equal validity to internal as well as surface membranes. Perhaps for this reason spectrin and ankyrin are also associated with internal membranes.

References
 1. Aberle, H., S. Butz, J. Stappert, H. Weissig, R. Kemler, and H. Hoschuetzky. Assembly of the cadherin‐catenin complex in vitro with recombinant proteins. J. Cell Sci. 107: 3655–3663, 1994.
 2. Aderem, A. Signal transduction and the actin cytoskeleton: the roles of MARCKS and profilin. Trends Biochem. Sci. 17: 438–43, 1992.
 3. Aghib, D. F., and P. D. McCrea. The E‐cadherin complex contains the src substrate p120. Exp. Cell Res. 218: 359–69, 1995.
 4. Agre, P., E. P. Orringer, and V. Bennett. Deficient red‐cell spectrin in severe, recessively inherited spherocytosis. N. Engl. J. Med. 306: 1155–1161, 1982.
 5. Alloisio, N., N. Dalla Venezia, A. Rana, K. Andrabi, P. Texier, F. Gilsanz, J. P. Cartron, J. Delaunay, and A. H. Chishti. Evidence that red blood cell protein p55 may participate in the skeleton‐membrane linkage that involves protein 4.1 and glycophorin C. Blood 82: 1323–1327, 1993.
 6. Amin, K. M., A. Scarpa, J. C. Winkelmann, P. J. Curtis, and B. G. Forget. The exon‐intron organization of the human erythroid beta‐spectrin gene. Genomics 18: 118–125, 1993.
 7. Anderson, J. M., A. S. Fanning, L. Lapierre, and C. M. Van Itallie. Zonula occludens ZO‐1 and ZO‐2: membrane‐associated guanylate kinase homologues (MAGuks) of the tight junction. Biochem. Soc. Trans. 23: 470–475, 1995.
 8. Anderson, J. P., and J. S. Morrow. The interaction of calmodulin with human erythrocyte spectrin: inhibition of protein 4.1 stimulated actin binding. J. Biol. Chem. 262: 6365–6372, 1987.
 9. Andrews, R. K., and J. E. Fox. Identification of a region in the cytoplasmic domain of the platelet membrane glycoprotein lb‐IX complex that binds to purified actin‐binding protein. J. Biol. Chem. 267: 18605–18611, 1992.
 10. Appleyard, S. T., M. J. Dunn, V. Dubowitz, M. L. Scott, S. L. Pittman, and D. M. Shotton. Monoclonal antibodies detect a spectrin‐like protein in normal and dystrophic human skeletal muscle. Proc. Natl. Acad. Sci. U.S.A. 81: 776–780, 1984.
 11. Atkinson, M.A.L., J. S. Morrow, and V. T. Marchesi. The polymeric state of actin in the human erythrocyte cytoskeleton. J. Cell Biochem. 18: 493–505, 1982.
 12. Aunis, D., and M.‐F. Bader. The cytoskeleton as a barrier to exocytosis in secretory cell. J. Exp. Biol. 139: 253–266, 1988.
 13. Axton, J. M., F. L. Shamanski, L. M. Young, D. S. Henderson, J. B. Boyd, and W. T. Orr. The inhibitor of DNA replication encoded by the Drosophila gene plutonium is a small, ankyrin repeat protein. EMBO J. 13: 462–70, 1994.
 14. Azim, A. C., J. H. Knoll, A. H. Beggs, and A. H. Chishti. Isoform cloning, actin binding, and chromosomal localization of human erythroid dematin, a member of the villin superfamily. J. Biol. Chem. 270: 17407–17413, 1995.
 15. Babcock, G. G., and V. M. Fowler. Isoform‐specific interaction of tropomodulin with skeletal muscle and erythrocyte tropomyosins. J. Biol. Chem. 269: 27510–27518, 1994.
 16. Bähler, M. F., and P. Greengard. Synapsin 1 bundles F‐actin in a phosphorylation‐dependent manner. Nature 326: 704–707, 1987.
 17. Baines, A. J., and V. Bennett. Synapsin I is a microtubulebundling protein. Nature 319: 145–147, 1986.
 18. Baltensperger, K., L. M. Kozma, A. D. Cherniack, J. K. Klarlund, A. Chawla, U. Banerjee, and M. P. Czech. Binding of the ras activator son of sevenless to insulin receptor substrate‐1 signaling complexes. Science 260: 1950–1952, 1993.
 19. Baron, M. D., M. D. Davison, P. Jones, and D. R. Critchley. The sequence of chick a‐actinin reveals homologies to spectrin and calmodulin. J. Biol. Chem. 262: 17623–17630, 1987.
 20. Beck, K. A., J. A. Buchanan, V. Malhotra, and W. J. Nelson. Golgi spectrin: identification of an erythroid beta‐spectrin homolog associated with the Golgi complex. J. Cell Biol. 127: 707–723, 1994.
 21. Becker, P. J., J. S. Morrow, and S. E. Lux. Abnormal oxidant sensitivity and beta chain structure of spectrin in hereditary spherocytosis associated with defective spectrin‐protein 4.1 binding. J. Clin. Invest. 80: 557–565, 1987.
 22. Becker, P. S., and S. E. Lux. Hereditary spherocytosis and related disorders. Clin. Haematol. 14: 15–43, 1985.
 23. Becker, P. S., M. A. Schwartz, J. S. Morrow, and S. E. Lux. Radiolabel‐transfer cross‐linking demonstrates that protein 4.1 binds to the N‐terminal region of beta spectrin and to actin in binary interactions. Eur. J. Biochem. 193: 827–836, 1990.
 24. Becker, P. S., W. T. Tse, S. E. Lux, and B. G. Forget. Beta spectrin kissimmee: a spectrin variant associated with autosomal dominant hereditary spherocytosis and defective binding to protein 4.1. J. Clin. Invest. 92: 612–616, 1993.
 25. Beggs, A. H., T. J. Byers, J. H. Knoll, F. M. Boyce, et al. Cloning and characterization of two human skeletal muscle alpha‐actinin genes located on chromosomes 1 and 11. J. Biol. Chem. 267: 9281–9288, 1992.
 26. Behrens, J., L. Vakaet, R. Friis, E. Winterhager, R. F. Van, M. M. Mareel, and W. Birchmeier. Loss of epithelial differentiation and gain of invasiveness correlates with tyrosine phosphorylation of the E‐cadherin/beta‐catenin complex in cells transformed with a temperature‐sensitive v‐SRC gene. J. Cell Biol. 120: 757–766, 1993.
 27. Belkin, A. M., N. I. Zhidkova, and V. E. Koteliansky. Localization of talin in skeletal and cardiac muscles. FEBS Lett. 200: 32–36, 1986.
 28. Bennett, V. Ankyrins. Adaptors between diverse plasma membrane proteins and the cytoplasm. J. Biol. Chem. 267: 8703–8706, 1982.
 29. Bennett, V., and D. Branton. Selective association of spectrin with the cytoplasmic surface of human erythrocyte plasma membranes. Quantitative determination with purified (32P) spectrin. J. Biol. Chem. 252: 2753–2763, 1977.
 30. Bennett, V., and J. Davis. Erythrocyte ankyrin: immunoreactive analogues are associated with mitotic structures in cultured cells and with microtubules in brain. Proc. Natl. Acad. Sci. U.S.A. 78: 7550–7554, 1981.
 31. Bennett, V., and D. M. Gilligan. The spectrin‐based membrane skeleton and micron‐scale organization of the plasma membrane. Annu. Rev. Cell Biol. 9: 27–66, 1993.
 32. Bennett, V., and S. Lambert. The spectrin skeleton: from red cells to brain. J. Clin. Invest. 87: 1483–1489, 1991.
 33. Bennett, V., and P. J. Stenbuck. Identification and partial purification of ankyrin, the high affinity membrane attachment site for human erythrocyte spectrin. J. Biol. Chem. 254: 2533–2541, 1979.
 34. Bennett, V., and P. J. Stenbuck. The membrane attachment protein for spectrin is associated with band 3 in human erythrocyte membranes. Nature 280: 468–73, 1979.
 35. Bianchi, G., P. Ferrarri, D. Trizio, M. Ferrandi, L. Torielli, B. R. Barber, and E. Polli. Red blood cell abnormalities and spontaneous hypertension in the rat: a genetically determined link. Hypertension 7: 319–325, 1985.
 36. Bianchi, G., G. Tripodi, G. Casari, S. Salardi, B. R. Barber, R. Garcia, P. Leoni, L. Torielli, D. Cusi, M. Ferrandi, et al. Two point mutations within the adducin genes are involved in blood pressure variation. Proc. Natl. Acad. Sci. U.S.A. 91: 3999–4003, 1994.
 37. Birchmeier, W., K. M. Weidner, and J. Behrens. Molecular mechanisms leading to loss of differentiation and gain of invasiveness in epithelial cells. J. Cell Sci. Suppl. 17: 159–164, 1993.
 38. Birkenmeier, C. S., R. A. White, L. L. Peters, E. J. Hall, S. E. Lux, and J. E. Barker. Complex patterns of sequence variation and multiple 5' and 3' ends are found among transcripts of the erythroid ankyrin gene. J. Biol. Chem. 268: 9533–9540, 1993.
 39. Bitbol, M., C. Dempsey, A. Watts, and P. F. Devaux. Weak interaction of spectrin with phosphatidylcholine‐phosphatidylserine. FEBS Lett. 244: 217–222, 1989.
 40. Black, J. D., S. T. Koury, R. B. Bankert, and E. A. Repasky. Heterogeneity in lymphocyte spectrin distribution: ultrastructural identification of a new spectrin‐rich cytoplasmic structure. J. Cell Biol. 106: 97–109, 1988.
 41. Bloch, R. J., and J. S. Morrow. An unusual β‐spectrin associated with clustered acetylcholine receptors. J. Cell Biol. 108: 481–493, 1989.
 42. Bloom, M. L., B. K. Lee, C. S. Birkenmeier, Y. Ma, W. E. Zimmer, S. R. Goodman, E. M. Eicher, and J. E. Barker. Brain beta spectrin isoform 235 (Spnb‐2) maps to mouse chromosome 11. Mamm. Genome 3: 293–295, 1992.
 43. Bonder, E. M., D. J. Fishkind, N. M. Cotran, and D. A. Begg. The cortical actin‐membrane cytoskeleton of unfertilized sea urchin eggs: analysis of the spatial organization and relationship of filamentous actin, nonfilamentous actin, and egg spectrin. Dev. Biol. 134: 327–341, 1989.
 44. Bourguignon, L. Y., A. Chu, H. Jin, and N. R. Brandt. Ryanodine receptor‐ankyrin interaction regulates internal Ca2+ release in mouse T‐lymphoma cells. J. Biol. Chem. 270: 17917–17922, 1995.
 45. Bourguignon, L. Y., N. Iida, and H. Jin. The involvement of the cytoskeleton in regulating IP3 receptor‐mediated internal Ca2+ release in human blood platelets. Cell Biol. Int. 17: 751–758, 1993.
 46. Bourguignon, L. Y., and H. Jin. Identification of the ankyrinbinding domain of the mouse T‐lymphoma cell inositol 1,4,5‐trisphosphate (IP3) receptor and its role in the regulation of IP3‐mediated internal Ca2+ release. J. Biol. Chem. 270: 7257–60, 1995.
 47. Bourguignon, L. Y., H. Jin, N. Iida, N. R. Brandt, and S. H. Zhang. The involvement of ankyrin in the regulation of inositol 1,4,5‐trisphosphate receptor‐mediated internal Ca2+ release from Ca2+ storage vesicles in mouse T‐lymphoma cells. J. Biol. Chem. 268: 7290–7297, 1993.
 48. Bourguignon, L. Y., E. L. Kalomiris, and V. B. Lokeshwar. Acylation of the lymphoma transmembrane glycoprotein, GP85, may be required for GP85‐ankyrin interaction. J. Biol. Chem. 266: 11761–11765, 1991.
 49. Bourguignon, L. Y., V. B. Lokeshwar, J. He, X. Chen, and G. J. Bourguignon. A CD44‐like endothelial cell transmembrane glycoprotein (GP116) interacts with extracellular matrix and ankyrin. Mol. Cell Biol. 12: 4464–4471, 1992.
 50. Bourguignon, L. Y., S. J. Suchard, and E. L. Kalomiris. Lymphoma Thy‐1 glycoprotein is linked to the cytoskeleton via a 4.1‐like protein. J. Cell Biol. 103: 2529–2540, 1986.
 51. Bourguignon, L. Y., S. J. Suchard, M. L. Nagpal, and J. J. Glenney. A T‐lymphoma transmembrane glycoprotein (gp180) is linked to the cytoskeletal protein, fodrin. J. Cell Biol. 101: 477–487, 1985.
 52. Bourguignon, L. Y., G. Walker, S. J. Suchard, and K. Balazovich. A lymphoma plasma membrane‐associated protein with ankyrin‐like properties. J. Cell Biol. 102: 2115–2124, 1986.
 53. Brady‐Kalnay, S. M., A. J. Flint, and N. K. Tonks. Homophilic binding of PTP mu, a receptor‐type protein tyrosine phosphatase, can mediate cell‐cell aggregation. J. Cell Biol. 122: 961–972, 1993.
 54. Brady‐Kalnay, S. M., D. L. Rimm, and N. K. Tonks. Receptor protein tyrosine phosphatase PTPmu associates with cadherins and catenins in vivo. J. Cell Biol. 130: 977–986, 1995.
 55. Branton, D., C. M. Cohen, and J. Tyler. Interaction of cytoskeletal proteins on the human erythrocyte membrane. Cell 24: 24–32, 1981.
 56. Bretscher, A. Rapid phosphorylation and reorganization of ezrin and spectrin accompany morphological changes induced in A‐431 cells by epidermal growth factor. J. Cell Biol. 108: 921–930, 1989.
 57. Bryant, P. J., K. L. Watson, R. W. Justice, and D. F. Woods. Tumor suppressor genes encoding proteins required for cell interactions and signal transduction in Drosophila. Dev. Suppl. 239–249, 1993.
 58. Button, E., C. Shapland, and D. Lawson. Actin, its associated proteins and metastasis. [Review]. Cell Motil. Cytoskeleton 30: 247–251, 1995.
 59. Buxton, R. S., and A. I. Magee. Structure and interactions of desmosomal and other cadherins. Semin. Cell Biol. 3: 157–167, 1992.
 60. Byers, T., E. Brandin, R. Lue, E. Winograd, and D. Branton. The complete sequence of Drosophila beta‐spectrin reveals supra‐motifs. Proc. Nat. Acad. Sci. U.S.A. 89: 6187–6191, 1992.
 61. Byers, T. J., A. Husain‐Chishti, R. R. Dubreuil, D. Branton, and L. S. Goldstein. Sequence similarity of the amino‐terminal domain of Drosophila beta spectrin to alpha actinin and dystrophin. J. Cell Biol. 109: 1633–1641, 1989.
 62. Calvert, R., P. Bennett, and W. Gratzer. Properties and structural role of the subunits of human spectrin. Eur. J. Biochem. 107: 355–361, 1980.
 63. Carboni, J., C. L. Howe, A. B. West, K. W. Barwick, M. S. Mooseker, and J. S. Morrow. Characterization of intestinal brush border cytoskeletal proteins of normal and neoplastic human epithelial cells: a comparison with the avian brush border. Am. J. Pathol. 129: 589–600, 1987.
 64. Carlier, M. F., C. Simon, R. Cassoly, and L. A. Pradel. Interaction between microtubule‐associated protein tau and spectrin. Biochimie 66: 305–311, 1984.
 65. Cartron, J. P., and C. Rahuel. Human erythrocyte glycophorins: protein and gene structure analyses. Transfus. Med. Rev. 6: 63–92, 1992.
 66. Casari, G., C. Barlassina, D. Cusi, L. Zagato, R. Muirhead, M. Righetti, P. Nembri, K. Amar, M. Gtti, F. Macciardi, G. Binelli, and G. Bianchi. Association of the alpha adducin locus with essential hypertension. Hypertension 25: 320–326, 1995.
 67. Chasis, J. A., P. Agre, and N. Mohandas. Decreased membrane mechanical stability and in vivo loss of surface area reflect spectrin deficiencies in hereditary spherocytosis. J. Clin. Invest. 82: 617–623, 1988.
 68. Chasis, J. A., L. Coulombel, J. Conboy, S. McGee, K. Andrews, Y. W. Kan, and N. Mohandas. Differentiation‐associated switches in protein 4.1 expression. Synthesis of multiple structural isoforms during normal human erythropoiesis. J. Clin. Invest. 91: 329–438, 1993.
 69. Chasis, J. A., M. E. Reid, R. H. Jensen, and N. Mohandas. Signal transduction by glycophorin A: role of extracellular and cytoplasmic domains in a modulatable process. J. Cell Biol. 107: 1351–1357, 1988.
 70. Cho, K. O., C. A. Hunt, and M. B. Kennedy. The rat brain postsynaptic density fraction contains a homolog of the Drosophila discs‐large tumor suppressor protein. Neuron 9: 929–942, 1992.
 71. Chu, Z. L., A. Wickrema, S. B. Krantz, and J. C. Winkelmann. Erythroid‐specific processing of human beta spectrin I pre‐mRNA. Blood 84: 1992–1999, 1994.
 72. Cianci, C. D., E.‐T. Chen, and J. S. Morrow. Cooperative interactions in the spectrin‐based cortical cytoskeleton: a kinetic study. Federation Proc. 240, 1987.
 73. Cianci, C. D., P. G. Gallagher, B. G. Forget, and J. S. Morrow. Unique subsets of proteins bind the SH3 domain of ãI spectrin and Grb2 in differentiating mouse erythroleukemia (MEL) cells. Mol. and Cell. Biol. 5: 271a (abstract), 1995.
 74. Cianci, C. D., M. Giorgi, and J. S. Morrow. Phosphorylation of ankyrin down‐regulates its cooperative interaction with spectrin and protein 3. J. Cell Biochem. 37: 301–315, 1988.
 75. Cianci, C. D., A. S. Harris, S. Mische, and J. S. Morrow. The calmodulin binding site of protein 4.1. submitted, 1996.
 76. Cianci, C. D., J. C. Kim, J. McLaughlin, P. R. Stabach, C. R. Lombardo, and J. S. Morrow. Spectrin isoform diversity and assembly in non‐erythroid cells. Cell. Mol. Biol. Lett. 1: 79–99, 1996.
 77. Cianci, C. D., and J. S. Morrow. Cloning and characterization of human fetal brain alpha II spectrin. GenBank Accession number U26396, 1995.
 78. Cicchetti, P., B. J. Mayer, G. Thiel, and D. Baltimore. Identification of a protein that binds to the SH3 region of abl and is similar to bcr and GAP‐rho. Science 257: 803–806, 1992.
 79. Cioe, L., P. Laurila, P. Meo, K. Krebs, S. Goodman, and P. J. Curtis. Cloning and nucleotide sequence of a mouse erythrocyte beta‐spectrin cDNA. Blood 70: 915–920, 1987.
 80. Clark, M. B., Y. Ma, M. L. Bloom, J. E. Barker, I. S. Zagon, W. E. Zimmer, and S. R. Goodman. Brain alpha erythroid spectrin: identification, compartmentalization, and beta spectrin associations. Brain Res. 663: 223–236, 1994.
 81. Clark, S. W., and D. I. Meyer. Centractin is an actin homologue associated with the centrosome [see comments]. Nature 359: 246–250, 1992.
 82. Clark, S. W., O. Staub, I. B. Clark, E. L. Holzbaur, B. M. Paschal, R. B. Vallee, and D. I. Meyer. Beta‐centractin: characterization and distribution of a new member of the centractin family of actin‐related proteins. Mol. Biol. Cell 5: 1301–1310, 1994.
 83. Cohen, C. M., E. Dotimas, and C. Korsgren. Human erythrocyte membrane protein band 4.2 (pallidin). Semin. Hematol. 30: 119–137, 1993.
 84. Cohen, C. M., and S. F. Foley. The role of band 4.1 in the association of actin with erythrocyte membranes. Biochim. Biophys. Acta 688: 691–701, 1982.
 85. Cohen, C. M., and S. F. Foley. Spectrin‐dependent and ‐ independent association of F‐actin with the erythrocyte membrane. J. Cell Biol. 86: 694–698, 1980.
 86. Cohen, C. M., S. F. Foley, and C. Korsgren. A protein immunologically related to erythrocyte band 4.1 is found on stress fibres on non‐erythroid cells. Nature 299: 648–650, 1982.
 87. Cohen, C. M., and C. Korsgren. Band 4.1 causes spectrin‐actin gels to become thixiotropic. Biochem. Biophys. Res. Commun. 97: 1429–1435, 1980.
 88. Cohen, C. M., J. M. Tyler, and D. Branton. Spectrin‐actin associations studied by electron microscopy of shadowed preparations. Cell 21: 875–883, 1980.
 89. Cole, N., and G. B. Ralston. Enhancement of self‐association of human spectrin by polyethylene glycol. Int. J. Biochem. 26: 799–804, 1994.
 90. Coleman, T. R., D. J. Fishkind, M. S. Mooseker, and J. S. Morrow. Contributions of the beta‐subunit to spectrin structure and function. Cell Motil. Cytoskeleton 12: 248–263, 1989.
 91. Coleman, T. R., D. J. Fishkind, M. S. Mooseker, and J. S. Morrow. Functional diversity among spectrin isoforms. Cell Motil. Cytoskeleton 12: 225–247, 1989.
 92. Coleman, T. R., A. S. Harris, S. M. Mische, M. S. Mooseker, and J. S. Morrow. Beta spectrin bestows protein 4.1 sensitivity on spectrin‐actin interactions. J. Cell Biol. 104: 519–526, 1987.
 93. Conboy, J., Y. W. Kan, S. B. Shohet, and N. Mohandas. Molecular cloning of protein 4.1, a major structural element of the human erythrocyte membrane skeleton. Proc. Natl. Acad. Sci. U.S.A. 83: 9512–9516, 1986.
 94. Conboy, J. G. Structure, function, and molecular genetics of erythroid membrane skeletal protein 4.1 in normal and abnormal red blood cells. Semin. Hematol. 30: 58–73, 1993.
 95. Conboy, J. G., J. Y. Chan, J. A. Chasis, Y. W. Kan, and N. Mohandas. Tissue‐ and development‐specific alternative RNA splicing regulates expression of multiple isoforms of erythroid membrane protein 4.1. J. Biol. Chem. 266: 8273–8280, 1991.
 96. Cook, T. A., R. Urrutia, and M. A. McNiven. Identification of dynamin 2, an isoform ubiquitously expressed in rat tissue. Proc. Natl. Acad. Sci. U.S.A. 91: 644–648, 1994.
 97. Correas, I. Characterization of isoforms of protein 4.1 present in the nucleus. Biochem. J. 279: 581–585, 1991.
 98. Correas, I., T. L. Leto, D. W. Speicher, and V. T. Marchesi. Identification of the functional site of erythrocyte protein 4.1 involved in spectrin‐actin associations. J. Biol. Chem. 261: 3310–3315, 1986.
 99. Cowin, P., H. P. Kapprell, W. W. Franke, J. Tamkun, and R. O. Hynes. Plakoglobin: a protein common to different kinds of intercellular adhering junctions. Cell 46: 1063–1073, 1986.
 100. Craig, S. W., and J. V. Pardo. Gamma actin, spectrin, and intermediate filament proteins colocalize with vinculin at costameres, myofibril‐to‐sarcolemma attachment sites. Cell Motil. 3: 449–462, 1983.
 101. Dahl, S. C., R. W. Geib, M. T. Fox, M. Edidin, and D. Branton. Rapid capping in alpha‐spectrin‐deficient MEL cells from mice afflicted with hereditary hemolytic anemia. J. Cell Biol. 125: 1057–1065, 1994.
 102. Daniel, J. M., and A. B. Reynolds. The tyrosine kinase substrate p120cas binds directly to E‐cadherin but not to the adenomatous polyposis coli protein or alpha‐catenin. Mol. Cell Biol. 15: 4819–4824, 1995.
 103. Davis, J., and V. Bennett. Brain spectrin. Isolation of subunits and formation of hybrids with erythrocyte spectrin subunits. J. Biol. Chem. 258: 7757–7766, 1983.
 104. Davis, J. Q., and V. Bennett. The anion exchanger and Na + K(+)‐ATPase interact with distinct sites on ankyrin in in vitro assays. J. Biol. Chem. 265: 17252–17256, 1990.
 105. Davis, J. Q., and V. Bennett. Ankyrin‐binding activity of nervous system cell adhesion molecules expressed in adult brain. J. Cell Sci. Suppl. 17: 109–117, 1993.
 106. Davis, J. Q., and V. Bennett. Brain ankyrin. A membrane‐associated protein with binding sites for spectrin, tubulin, and the cytoplasmic domain of the erythrocyte anion channel. J. Biol. Chem. 259: 13550–13559, 1984.
 107. Davis, J. Q., T. McLaughlin, and V. Bennett. Ankyrin‐binding proteins related to nervous system cell adhesion molecules: candidates to provide transmembrane and intercellular connections in adult brain. J. Cell Biol. 121: 121–33, 1993.
 108. Davis, L., and V. Bennett. Purification of a calmodulin‐sensitive spectrin‐binding membrane glycoprotein from brain. Mol. Biol. Cell 6: 1566A, 1995.
 109. Davis, L. H., and V. Bennett. Identification of two regions of beta G spectrin that bind to distinct sites in brain membranes. J. Biol. Chem. 269: 4409–4416, 1994.
 110. Davis, L. H., and V. Bennett. Mapping the binding sites of human erythrocyte ankyrin for the anion exchanger and spectrin. J. Biol. Chem. 265: 10589–10596, 1990.
 111. Davis, L. H., J. Q. Davis, and V. Bennett. Ankyrin regulation: an alternatively spliced segment of the regulatory domain functions as an intramolecular modulator. J. Biol. Chem. 267: 18966–18972, 1992.
 112. De Arruda, M. V., S. Watson, C. S. Lin, J. Leavitt, and P. Matsudaira. Fimbrin is a homologue of the cytoplasmic phosphoprotein plastin and has domains homologous with calmodulin and actin gelation proteins. J. Cell Biol. 111: 1069–1070, 1990.
 113. Delaunay, J., N. Alloisio, L. Morle, and F. Baklouti. The genetic disorders of the red cell skeleton. Nouv. Rev. Er. Hematol. 33: 63–70, 1991.
 114. DeSilva, T. M., K. C. Peng, K. D. Speicher, and D. W. Speicher. Analysis of human red cell spectrin tetramer (head‐to‐head) assembly using complementary univalent peptides. Biochemistry 31: 10872–10878, 1992.
 115. Devarajan, P., and J. S. Morrow. The spectrin cytoskeleton and organization of polarized epithelial cell membranes. In: Membrane Protein—Cytoskeleton Interactions, edited by W. J. Nelson. New York: Academic Press, 1996, Vol. 43, p. 97–128.
 116. Devarajan, P., D. A. Scaramuzzino, and J. S. Morrow. Ankyrin binds to two distinct cytoplasmic domains of Na,K‐ATPase ã subunit. Proc. Natl. Acad. Sci. U.S.A. 91: 2965–2969, 1994.
 117. Devarajan, P., P. R. Stabach, A. S. Mann, T. Ardito, M. Kashgarian, and J. S. Morrow. Identification of a small cytoplasmic ankyrin (AnkG119) in kidney and muscle that binds βIc* spectrin and associates with the Golgi apparatus. J. Cell Biol. 133: 819–830, 1996.
 118. Diakowski, W., and A. F. Sikorski. Brain spectrin interacts with membrane phospholipids. Acta Biochim. Pol. 41: 153–154, 1994.
 119. Ding, D., S. Parkhurst, and H. Lipshitz. Different genetic requirements for anterior RNA localization revealed by the distribution of adducin‐like transcripts during Drosophila oogenesis. Proc. Natl. Acad. Sci. U.S.A. 90: 2512–2516, 1993.
 120. Ding, Y., J. R. Casey, and R. R. Kopito. The major kidney AE1 isoform does not bind ankyrin (Ank1) in vitro. An essential role for the 79 NH2‐terminal amino acid residues of band 3. J. Biol. Chem. 269: 32201–32208, 1994.
 121. Discher, D. E., R. Winardi, P. O. Schischmanoff, M. Parra, J. G. Conboy, and N. Mohandas. Mechanochemistry of protein 4.1's spectrin‐actin‐binding domain: ternary complex interactions, membrane binding, network integration, structural strengthening. J. Cell Biol. 130: 897–907, 1995.
 122. Dmytrenko, G. M., D. W. Pumplin, and R. J. Bloch. Dystrophin in a membrane skeletal network: localization and comparison to other proteins. J. Neurosci. 13: 547–558, 1993.
 123. Dong, L., C. Chapline, B. Mousseau, L. Fowler, K. Ramsay, J. Stevens, and S. Jaken. 35H, a sequence isolated as a protein kinase C binding protein, is a novel member of the adducin family. J. Biol. Chem. 270: 25534–25540, 1995.
 124. Dubreuil, R. R. Structure and evolution of the actin crosslinking proteins. Bioessays 13: 219–26, 1991.
 125. Dubreuil, R. R., E. Branton, J. H. Reisberg, L. S. Goldstein, and D. Branton. Structure, calmodulin‐binding, and calcium‐binding properties of recombinant alpha spectrin polypeptides. J. Biol. Chem. 266: 7189–7193, 1991.
 126. Dubreuil, R. R., T. J. Byers, A. L. Sillman, D. Bar‐Zvi, L. S. Goldstein, and D. Branton. The complete sequence of Drosophila alpha‐spectrin: conservation of structural domains between alpha‐spectrins and alpha‐actinin. J. Cell Biol. 109: 2197–2205, 1989.
 127. Dubreuil, R. R., T. J. Byers, C. T. Stewart, and D. P. Kiehart. A beta‐spectrin isoform from Drosophila (beta H) is similar in size to vertebrate dystrophin. J. Cell Biol. 111: 1849–1858, 1990.
 128. Dubreuil, R. R., and J. Yu. Ankyrin and beta‐spectrin accumulate independently of alpha‐spectrin in Drosophila. Proc. Natl. Acad. Sci. U.S.A. 91: 10285–10289, 1994.
 129. Eber, S. W., W. M. Lande, T. A. Iarocci, W. C. Mentzer, P. Hohn, J. S. Wiley, and W. Schroter. Hereditary stomatocytosis: consistent association with an integral membrane protein deficiency. Br. J. Haematol. 72: 452–455, 1989.
 130. Egan, S. E., B. W. Giddings, M. W. Brooks, L. Buday, A. M. Sizeland, and R. A. Weinberg. Association of Sos Ras exchange protein with Grb2 is implicated in tyrosine kinase signal transduction and transformation [see comments]. Nature 363: 45–51, 1993.
 131. Elgsaeter, A., and A. Mikkelsen. Shapes and shape changes in vitro in normal red blood cells. Biochim. Biophys. Acta 1071: 273–290, 1991.
 132. Elgsaeter, A., D. M. Shotton, and D. Branton. Intramembrane particle aggregation in erythrocyte ghosts. Il. The influence of spectrin aggregation. Biochim. Biophys. Acta 426: 101–122, 1976.
 133. Fach, B. L., S. F. Graham, and R. A. Keates. Association of fodrin with brain microtubules. Can. J. Biochem. Cell Biol. 63: 372–381, 1985.
 134. Fairbanks, G., T. L. Steck, and D. F. Wallach. Electrophoretic analysis of the major polypeptides of the human erythrocyte membrane. Biochemistry 10: 2606–2617, 1971.
 135. Faire, K., F. Trent, J. M. Tepper, and S. M. Bonder. Analysis of dynamin isoforms in mammalian brain: dynamin‐1 expression is spatially and temporally regulated during postnatal development. Proc. Natl. Acad. Sci. U.S.A. 89: 8376–8380, 1992.
 136. Falzone, C. J., Y. H. Kao, J. Zhao, D. A. Bryant, and J. T. Lecomte. Three‐dimensional solution structure of PsaE from the cyanobacterium Synechococcus sp. strain PCC 7002, a photosystem I protein that shows structural homology with SH3 domains. Biochemistry 33: 6052–6062, 1994.
 137. Faquin, W. C., C. A. Husain, and D. Branton. Expression of dematin (protein 4.9) during avian erythropoiesis. Eur. J. Cell Biol. 53: 48–58, 1990.
 138. Faraday, C. D., and R. M. Spanswick. Evidence for a membrane skeleton in higher plants. A spectrin‐like polypeptide co‐isolates with rice root plasma membranes. FEBS Lett. 318: 313–6, 1993.
 139. Fearon, E. R., T. Finkel, M. L. Gillison, S. P. Kennedy, J. F. Casella, G. F. Tomaselli, J. S. Morrow, and C. V. Dang. Karyoplasmic interaction selection strategy (KISS): A general strategy to detect protein‐protein interactions in mammalian cells. Proc. Natl. Acad. Sci. U.S.A. 89: 7958–7962, 1992.
 140. Ferrarri, P., L. Torielli, M. Ferrandi, and G. Bianchi. Volumes and Na transports in intact red blood cells, resealed ghosts, and inside‐out vesicles of Milan‐hypertensive rats. In: Membrane Pathology, edited by G. Bianchi, E. Carafoli and A. Scarpe, New York: Ann NY Acad Sci, 1986, p. 561–566.
 141. Fishkind, D. J., E. M. Bonder, and D. A. Begg. Isolation and characterization of sea urchin egg spectrin: calcium modulation of the spectrin‐actin interaction. Cell Motil. Cytoskeleton 7: 304–314, 1987.
 142. Fishkind, D. J., E. M. Bonder, and D. A. Begg. Sea urchin spectrin in oogenesis and embryogenesis: a multifunctional integrator of membrane‐cytoskeletal interactions. Dev. Biol. 142: 453–464, 1990.
 143. Fishkind, D. J., E. M. Bonder, and D. A. Begg. Subcellular localization of sea urchin egg spectrin: evidence for assembly of the membrane‐skeleton on unique classes of vesicles in eggs and embryos. Dev. Biol. 142: 439–452, 1990.
 144. Fowler, V. M., and V. Bennett. Erythrocyte membrane tropomyosin. Purification and properties. J. Biol. Chem. 259: 5978–5989, 1984.
 145. Fowler, V. M., and V. Bennett. Tropomyosin: a new component of the erythrocyte membrane skeleton. Prog. Clin. Biol. Res. 159: 57–71, 1984.
 146. Fowler, V. M., M. A. Sussmann, P. G. Miller, B. E. Flucher, and M. P. Daniels. Tropomodulin is associated with the free (pointed) ends of the thin filaments in rat skeletal muscle. J. Cell Biol. 120: 411–420, 1993.
 147. Franck, Z., R. Gary, and A. Bretscher. Moesin, like ezrin, colocalizes with actin in the cortical cytoskeleton in cultured cells, but its expression is more variable. J. Cell Sci. 105: 219–231, 1993.
 148. Frappier, T., J. Derancourt, and L. A. Pradel. Actin and neurofilament binding domain of brain spectrin beta subunit. Eur. J. Biochem. 205: 85–91, 1992.
 149. Frappier, T., F. Regnouf, and L. A. Pradel. Binding of brain spectrin to the 70‐kDa neurofilament subunit protein. Eur. J. Biochem. 169: 651–657, 1987.
 150. Frappier, T., F. Stetzkowski‐Marden, and L. A. Pradel. Interaction domains of neurofilament light chain and brain spectrin. Biochem. J. 275 (Pt 2): 521–527, 1991.
 151. Funayama, N., A. Nagafuchi, N. Sato, S. Tsukita, and S. Tsukita. Radixin is a novel member of the band 4.1 family. J. Cell Biol. 115: 1039–1048, 1991.
 152. Gallagher, P. G., and B. G. Forget. Structure, organization and expression of the human band 7.2b gene, a candidate gene for hereditary hydrocytosis. J. Biol. Chem. 270: 26358–26363, 1995.
 153. Gallagher, P. G., W. T. Tse, A. L. Scarpa, S. E. Lux, and B. G. Forget. Large numbers of alternatively spliced isoforms of the regulatory region of human erythrocyte ankyrin. Trans. Assoc. Am. Physicians 105: 268–277, 1992.
 154. Gallagher, P. G., M. Upender, D. C. Ward, and B. G. Forget. The gene for human erythrocyte membrane protein band 7.2 (EPB72) maps to 9q33‐q34 centromeric to the Philadelphia chromosome translocation breakpoint region. Genomics 18: 167–169, 1993.
 155. Gardner, K., and V. Bennett. Modulation of spectrin‐actin assembly by erythrocyte adducin. Nature 328: 359–362, 1987.
 156. Gardner, K., and V. Bennett. A new erythrocyte membrane‐associated protein with calmodulin binding activity. Identification and purification. J. Biol. Chem. 261: 1339–1348, 1986.
 157. Garnier, J., D. J. Osguthorpe, and B. Robson. Analysis of the accuracy and implications of simple methods for predicting the secondary structure of globular proteins. J. Mol. Biol. 120: 97–120, 1978.
 158. Geiduschek, J. B., and S. J. Singer. Molecular changes in the membrane of mouse erythroid cells accompanying differentiation. Cell 16: 149–163, 1979.
 159. Georgatos, S. D., and V. T. Marchesi. The binding of vimentin to human erythrocyte membranes: a model system for the study of intermediate filament‐membrane interactions. J. Cell Biol. 100: 1955–1961, 1985.
 160. Georgatos, S. D., D. C. Weaver, and V. T. Marchesi. Site specificity in vimentin‐membrane interactions: intermediate filament subunits associate with the plasma membrane via their head domains. J. Cell Biol. 100: 1962–1967, 1985.
 161. Giebelhaus, D. H., D. W. Eib, and R. T. Moon. Antisense RNA inhibits expression of membrane skeleton protein 4.1 during embryonic development of Xenopus. Cell 53: 601–615, 1988.
 162. Gilligan, D., and V. Bennett. Alternative splicing of adducin beta subunit. J. Cell Biol. 115: 42a, 1991.
 163. Gilligan, D. M., and V. Bennett. Localization of adducin iso‐forms in brain and kidney. Mol. Biol. Cell 5: 421 (abstract), 1994.
 164. Gilligan, D. M., J. Lieman, and V. Bennett. Assignment of the human beta‐adducin gene (ADD2) to 2p13‐p14 by in situ hybridization. Genomics 28: 610–612, 1995.
 165. Glantz, S. B., and J. S. Morrow. The spectrin‐actin cytoskeleton and membrane integrity in hypoxia. In: Tissue Oxygen Deprivation: Developmental, Molecular and Integrated Function, edited by G. G. Haddad and G. Lister, New York: Marcel Dekker, 1995, p. 153–192.
 166. Glenney, J. R., Jr., and P. Glenney. Fodrin is the general spectrin‐like protein found in most cells whereas spectrin and the TW protein have a restricted distribution. Cell 34: 503–512, 1983.
 167. Goldberg, Y. P., B. Y. Lin, S. E. Andrew, J. Nasir, R. Graham, M. L. Glaves, G. Hutchinson, J. Theilmann, D. G. Ginzinger, K. Schappert, L. Clarke, J. M. Rommens, and M. R. Hayden. Cloning and mapping of the alpha‐adducin gene close to D4S95 and assessment of its relationship to Huntington disease. Hum. Mol. Genet. 1: 669–675, 1992.
 168. Goodman, S. R., K. E. Krebs, C. F. Whitfield, B. M. Riederer, and I. S. Zagon. Spectrin and related molecules. CRC Crit. Rev. Biochem. 23: 171–234, 1988.
 169. Goodman, S. R., I. S. Zagon, C. F. Whitfield, L. A. Casoria, P. J. McLaughlin, and T. L. Laskiewicz. A spectrin‐like protein from mouse brain membranes: immunological and structural correlations with erythrocyte spectrin. Cell Motil. 3: 635–647, 1983.
 170. Goodman, S. R., W. E. Zimmer, M. B. Clark, I. S. Zagon, J. E. Barker, and M. L. Bloom. Brain spectrin: of mice and men. Brain Res. Bull. 36: 593–606, 1995.
 171. Gorlin, J. B., R. Yamin, S. Egan, M. Stewart, T. P. Stossel, D. J. Kwiatkowski, and J. H. Hartwig. Human endothelial actin‐binding protein (ABP‐280, nonmuscle filamin): a molecular leaf spring. J. Cell. Biol. 111: 1089–1105, 1990.
 172. Gout, I., R. Dhand, I. D. Hiles, M. J. Fry, G. Panayotou, P. Das, O. Truong, N. F. Totty, J. Hsuan, G. W. Booker, et al. The GTPase dynamin binds to and is activated by a subset of SH3 domains. Cell 75: 25–36, 1993.
 173. Gregorio, C. C., and V. M. Fowler. Mechanisms of thin filament assembly in embryonic chick cardiac myocytes: tropomodulin requires tropomyosin for assembly. J. Cell Biol. 129: 683–695, 1995.
 174. Gregorio, C. C., R. T. Kubo, R. B. Bankert, and E. A. Repasky. Translocation of spectrin and protein kinase C to a cytoplasmic aggregate upon lymphocyte activation. Proc. Natl. Acad. Sci. U.S.A. 89: 4947–4951, 1992.
 175. Gregorio, C. C., E. A. Repasky, V. M. Fowler, and J. D. Black. Dynamic properties of ankyrin in T lymphocytes: colocalization with spectrin and protein kinase C beta. J. Cell Biol. 125: 345–358, 1994.
 176. Gregorio, C. C., A. Weber, M. Bondad, C. R. Pennise, and V. M. Fowler. Requirement of pointed‐end capping by tropomodulin to maintain actin filament length in embryonic chick cardiac myocytes. Nature 377: 83–86, 1995.
 177. Gundersen, D., J. Orlowski, and E. Rodriquez‐Boulan. Apical polarity of Na, K‐ATPase in retinal pigment epithelium is linked to a reversal of the ankyrin‐fodrin submembrane cytoskeleton. J. Biol. Chem. 112: 863–872, 1991.
 178. Hall, T. G., and V. Bennett. Regulatory domains of erythrocyte ankyrin. J. Biol. Chem. 262: 10537–10545, 1987.
 179. Hanspal, M., J. S. Hanspal, R. Kalraiya, S. C. Liu, K. E. Sahr, D. Howard, and J. Palek. Asynchronous synthesis of membrane skeletal proteins during terminal maturation of murine erythroblasts. Blood 80: 530–539, 1992.
 180. Hanspal, M., J. S. Hanspal, R. Kalraiya, and J. Palek. The expression and synthesis of the band 3 protein initiates the formation of a stable membrane skeleton in murine Rauscher‐transformed erythroid cells. Eur. J. Cell Biol. 58: 313–318, 1992.
 181. Hanspal, M., R. Kalraiya, J. Hanspal, K. E. Sahr, and J. Palek. Erythropoietin enhances the assembly of alpha, beta spectrin heterodimers on the murine erythroblast membranes by increasing beta spectrin synthesis. J. Biol. Chem. 266: 15626–15630, 1991.
 182. Harell, D., and M. Morrison. Two‐dimensional separation of erythrocyte membrane proteins. Arch. Biochem. Biophys. 193: 158–168, 1979.
 183. Harris, A. S., C. D. Cianci, S. M. Mische, and J. S. Morrow. Characterization of human erythrocyte protein 4.1‐calmodulin (CAM) interactions. In: Proceedings of the Symposium on Cytoskeleton and Cell Regulation. Steamboat Springs, CO: 237, 1990.
 184. Harris, A. S., D. Croall, and J. S. Morrow. The calmodulin‐binding site in fodrin is near the site of calcium‐dependent protease‐I cleavage. J. Biol. Chem. 263: 15754–15761, 1988.
 185. Harris, A. S., D. E. Croall, and J. S. Morrow. Calmodulin regulates fodrin susceptibility to cleavage by calcium‐dependent protease I. J. Biol. Chem. 264: 17401–17408, 1989.
 186. Harris, A. S., L. A. D. Green, K. J. Ainger, and J. S. Morrow. Mechanisms of cytoskeletal regulation : Functional differences correlate with antigenic dissimilarity in human brain and erythrocyte spectrin. Biochim. Biophys. Acta 830: 147–158, 1985.
 187. Harris, A. S., and J. S. Morrow. Calmodulin and calcium‐dependent protease I coordinately regulate the interaction of fodrin with actin. Proc. Natl. Acad. Sci. U.S.A. 87: 3009–3013, 1990.
 188. Harris, A. S., and J. S. Morrow. Proteolytic processing of human brain alpha spectrin (Fodrin): Identification of a hypersensitive site. J. Neurosci. 8: 2640–2651, 1988.
 189. Hartwig, J. H. Spectrin superfamily, subfamily 1: The spectrin family. Protein Profile 1: 715–749, 1994.
 190. Hartwig, J. H., and D. J. Kwiatkowski. Actin‐binding proteins. Curr. Opin. Cell Biol. 3: 87–97, 1991.
 191. Haslam, R. J., H. B. Koide, and B. A. Hemmings. Pleckstrin domain homology [letter]. Nature 363: 309–310, 1993.
 192. Hayes, N. V., F. E. Holmes, J. Grantham, and A. J. Baines. A60, an axonal membrane‐skeletal spectrin‐binding protein. Biochem. Soc. Trans. 23: 54–58, 1995.
 193. Hayette, S., D. Dhermy, M. E. dos Santos, M. Bozon, D. Drenckhahn, N. Alloisio, P. Texier, J. Delaunay, and L. Morle. A deletional frameshift mutation in protein 4.2 gene (allele 4.2 Lisboa) associated with hereditary hemolytic anemia. Blood 85: 250–256, 1995.
 194. Hemming, N. J., D. J. Anstee, W. J. Mawby, M. E. Reid, and M. J. Tanner. Localization of the protein 4.1‐binding site on human erythrocyte glycophorins C and D [published erratum appears in Biochem. J. 1994 Jun 15;300(Pt 3):920]. Biochem. J. 299: 191–196, 1994.
 195. Hemming, N. J., D. J. Anstee, M. A. Staricoff, M.J.A. Tanner, and N. Mohandas. Identification of the membrane attachment sites for protein 4.1 in the human erythrocyte. J. Biol. Chem. 270: 5360–5366, 1995.
 196. Henniker, A., and G. B. Ralston. Reinvestigation of the thermodynamics of spectrin self‐association. Biophys. Chem. 52: 251–258, 1994.
 197. Herman, I. M. Actin isoforms. [Review]. Curr. Opin. Cell Biol. 5: 48–55, 1993.
 198. Herrmann, H., and G. Wiche. Plectin and IFAP‐300K are homologous proteins binding to microtubule‐associated proteins 1 and 2 and to the 240‐kilodalton subunit of spectrin. J. Biol. Chem. 262: 1320–1325, 1987.
 199. Hiebl‐Dirschmied, C. M., G. R. Adolf, and R. Prohaska. Isolation and partial characterisation of the human band 7 integral membrane protein. Biochim. Biophys. Acta 1065: 195–202, 1991.
 200. Hiebl‐Dirschmied, C. M., B. Entler, C. Glotzmann, I. Maurer‐Fogy, C. Stratowa, and R. Prohaska. Cloning and nucleotide sequence of cDNA encoding human erythrocyte band 7 integral membrane protein. Biochim. Biophys. Acta 1090: 123–124, 1991.
 201. Hirokawa, N., R. E. Cheney, and M. Willard. Location of a protein of the fodrin‐spectrin‐TW260/240 family in the mouse intestinal brush border. Cell 32: 953–65, 1983.
 202. Hock, R. S., G. Davis, and D. W. Speicher. Purification of human smooth muscle filamin and characterization of structural domains and functional sites. Biochemistry 29: 9441–9451, 1990.
 203. Horne, W. C., S. C. Huang, P. S. Becker, T. K. Tang, and E. J. Benz, Jr.. Tissue‐specific alternative splicing of protein 4.1 inserts an exon necessary for formation of the ternary complex with erythrocyte spectrin and F‐actin. Blood 82: 2558–2563, 1993.
 204. Horne, W. C., T. L. Leto, and V. T. Marchesi. Differential phosphorylation of multiple sites in protein 4.1 and protein 4.9 by phorbol ester‐activated and cyclic AMP‐dependent protein kinases. J. Biol. Chem. 260: 9073–9076, 1985.
 205. Horne, W. C., W. C. Prinz, and E. K. Tang. Identification of two cAMP‐dependent phosphorylation sites on erythrocyte protein 4.1. Biochim. Biophys. Acta 1055: 87–92, 1990.
 206. Hoschuetzky, H., H. Aberle, and R. Kemler. Beta‐catenin mediates the interaction of the cadherin‐catenin complex with epidermal growth factor receptor. J. Cell Biol. 127: 1375–1380, 1994.
 207. Howe, C. L., L. M. Sacramone, M. S. Mooseker, and J. S. Morrow. Mechanisms of cytoskeletal regulation: modulation of membrane affinity in avian brush border and erythrocyte spectrins. J. Cell Biol. 101: 1379–1385, 1985.
 208. Hu, R. J., S. Moorthy, and V. Bennett. Expression of functional domains of beta G‐spectrin disrupts epithelial morphology in cultured cells. J. Cell Biol. 128: 1069–80, 1995.
 209. Huang, J. P., C. J. Tang, G. H. Kou, V. T. Marchesi, E. J. Benz, Jr., and T. K. Tang. Genomic structure of the locus encoding protein 4.1. Structural basis for complex combinational patterns of tissue‐specific alternative RNA splicing. J. Biol. Chem. 268: 3758–66, 1993.
 210. Huang, M., and M. Chalfie. Gene interactions affecting mechanosensory transduction in Caenorhabditis elegans [see comments]. Nature 367: 467–470, 1994.
 211. Huang, M., G. Gu, E. L. Ferguson, and M. Chalfie. A stomatin‐like protein necessary for mechanosensation in C. elegans. Nature 378: 292–295, 1995.
 212. Huebner, K., A. P. Palumbo, M. Isobe, C. A. Kozak, S. Monaco, G. Rovera, C. M. Croce, and P. J. Curtis. The alpha‐spectrin gene is on chromosome 1 in mouse and man. Proc. Natl. Acad. Sci. U.S.A. 82: 3790–3793, 1985.
 213. Hughes, C. A., and V. Bennett. Adducin: a physical model with implications for function in assembly of spectrin‐actin complexes. J. Biol. Chem. 270: 18990–18996, 1995.
 214. Hulsken, J., W. Birchmeier, and J. Behrens. E‐cadherin and APC compete for the interaction with beta‐catenin and the cytoskeleton. J. Cell Biol. 127: 2061–2069, 1994.
 215. Husain‐Chishti, A., W. Faquin, C. C. Wu, and D. Branton. Purification of erythrocyte dematin (protein 4.9) reveals an endogenous protein kinase that modulates actin‐bundling activity. J. Biol. Chem. 264: 8985–8991, 1989.
 216. Husain‐Chishti, A., A. Levin, and D. Branton. Abolition of actin‐bundling by phosphorylation of human erythrocyte protein 4.9. Nature 334: 718–721, 1988.
 217. Hyvonen, M., M. J. Macias, M. Nilges, H. Oschkinat, M. Saraste, and M. Wilmanns. Structure of the binding site for inositol phosphates in a PH domain. EMBO J. 14: 4676–4685, 1995.
 218. Ideguchi, H., J. Nishimura, H. Nawata, and N. Hamasaki. A genetic defect of erythrocyte band 4.2 protein associated with hereditary spherocytosis. Br. J. Haematol. 74: 347–353, 1990.
 219. Iglic, A., S. Svetina, and B. Zeks. Depletion of membrane skeleton in red blood cell vesicles. Biophys J. 69: 274–279, 1995.
 220. Iglic, A., S. Svetina, and B. Zeks. A role of membrane skeleton in discontinuous red blood cell shape transformations. Cell Mol. Biol. Lett. 1: 137–144, 1996.
 221. Iida, N., V. B. Lokeshwar, and L. Y. Bourguignon. Mapping the fodrin binding domain in CD45, a leukocyte membrane‐associated tyrosine phosphatase. J. Biol. Chem. 269: 28576–28583, 1994.
 222. Jesaitis, A. J., G. M. Bokoch, J. O. Tolley, and R. A. Allen. Lateral segregation of neutrophil chemotactic receptors into actin‐ and fodrin‐rich plasma membrane microdomains depleted in guanyl nucleotide regulatory proteins. J. Cell Biol. 107: 921–928, 1988.
 223. Jesaitis, A. J., J. O. Tolley, and R. A. Allen. Receptorcytoskeleton interactions and membrane traffic may regulate chemoattractant‐induced superoxide production in human granulocytes. J. Biol. Chem. 261: 13662–13669, 1986.
 224. Jesaitis, L. A., and D. A. Goodenough. Molecular characterization and tissue distribution of ZO‐2, a tight junction protein homologous to ZO‐1 and the Drosophila discs‐large tumor suppressor protein. J. Cell Biol. 124: 949–961, 1994.
 225. Johnson, G. V., J. M. Litersky, and R. S. Jope. Degradation of microtubule‐associated protein 2 and brain spectrin by calpain: a comparative study. J. Neurochem. 56: 1630–1638, 1991.
 226. Joseph, S. K., and S. Samanta. Detergent solubility of the inositol trisphosphate receptor in rat brain membranes. Evidence for association of the receptor with ankyrin. J. Biol. Chem. 268: 6477–6486, 1993.
 227. Joshi, R., and V. Bennett. Mapping the domain structure of human erythrocyte adducin. J. Biol. Chem. 265: 13130–13136, 1990.
 228. Joshi, R., D. M. Gilligan, E. Otto, T. McLaughlin, and V. Bennett. Primary structure and domain organization of human alpha and beta adducin. J. Cell Biol. 115: 665–675, 1991.
 229. Kahana, E., J. C. Pinder, K. S. Smith, and W. B. Gratzer. Fluorescence quenching of spectrin and other red cell membrane cytoskeletal proteins. Relation to hydrophobic binding sites. Biochem. J. 282: 75–80, 1992.
 230. Kaiser, H. W., E. O'Keefe, and V. Bennett. Adducin: Ca++‐dependent association with sites of cell‐cell contact. J. Cell Biol. 109: 557–569, 1989.
 231. Kapfhamer, D., D. E. Miller, S. Lambert, V. Bennett, T. W. Glover, and M. Burmeister. Chromosomal localization of the ankyrinG gene (ANK3/Ank3) to human 10q21 and mouse 10. Genomics 27: 189–91, 1995.
 232. Karinch, A. M., W. E. Zimmer, and S. R. Goodman. The identification and sequence of the actin‐binding domain of human red blood cell beta‐spectrin. J. Biol. Chem. 265: 11833–11840, 1990.
 233. Kelly, G. M., B. D. Zelus, and R. T. Moon. Identification of a calcium‐dependent calmodulin‐binding domain in Xenopus membrane skeleton protein 4.1. J. Biol. Chem. 266: 12469–12473, 1991.
 234. Kennedy, S. P., S. L. Warren, B. G. Forget, and J. S. Morrow. Ankyrin binds to the 15th repetitive unit of erythroid and non‐erythroid β‐spectrin. J. Cell Biol. 115: 267–277, 1991.
 235. Kennedy, S. P., S. A. Weed, B. G. Forget, and J. S. Morrow. A partial structural repeat forms the heterodimer self‐association site of all β‐spectrins. J. Biol. Chem. 269: 11400–11408, 1994.
 236. Kennedy, S. P., S. A. Weed, J. Winkleman, and J. S. Morrow. The self‐association site of recombinant human erythroid and muscle β‐spectrin. J. Cell Biol. 115: 42a, 1991.
 237. Kikuchi, A., K. Kaibuchi, Y. Hori, H. Nonaka, T. Sakoda, M. Kawamura, T. Mizuno, and Y. Takai. Molecular cloning of the human cDNA for a stimulatory GDP/GTP exchange protein for c‐Ki‐ras p21 and smg p21. Oncogene 7: 289–93, 1992.
 238. Kim, E., M. Niethammer, A. Rothschild, Y. N. Jan, and M. Sheng. Clustering of Shaker‐type K+ channels by interaction with a family of membrane‐associated guanylate kinases. Nature 378: 85–88, 1995.
 239. Kim, S. K. Tight junction, membrane‐associated guanylate kinases and cell signaling. Curr. Opin. Cell Biol. 7: 641–649, 1995.
 240. Kistner, U., B. M. Wenzel, R. W. Veh, C. Cases‐Langhoff, A. M. Garner, U. Appeltauer, B. Voss, E. D. Gundelfinger, and C. C. Garner. SAP90, a rat presynaptic protein related to the product of the Drosophila tumor suppressor gene dlg‐A. J. Biol. Chem. 268: 4580–4583, 1993.
 241. Klotz, K. N., and A. J. Jesaitis. Neutrophil chemoattractant receptors and the membrane skeleton. Bioessays 16: 193–198, 1994.
 242. Knowles, W. J., W. J. Knowles, J. S. Morrow, D. W. Speicher, H. S. Zarkowsky, N. Mohandas, W. C. Mentzer, S. B. Shohet, and V. T. Marchesi. Spectrin from patients with hereditary pyropoikilocytosis have self‐association defects and an abnormal tryptic digestion map. In: ICN/UCLA Symposium on Differentiation and Function of Hematopoietic Cell Surfaces, Keystone, CO, February. Keystone: ICN/UCLA, 1981.
 243. Knowles, W. J., J. S. Morrow, D. W. Speicher, H. S. Zarkowsky, N. Mohandas, W. C. Mentzer, S. B. Shohet, and V. T. Marchesi. Molecular and functional changes in spectrin from patients with hereditary pyropoikilocytosis. J. Clin. Invest. 71: 1867–1877, 1983.
 244. Knudsen, K. A., A. P. Soler, K. R. Johnson, and M. J. Wheelock. Interaction of alpha‐actinin with the cadherin/catenin cell‐cell adhesion complex via alpha‐catenin. J. Cell. Biol. 130: 67–77, 1995.
 245. Knudsen, K. A., and M. J. Wheelock. Plakoglobin, or an 83‐kD homologue distinct from beta‐catenin, interacts with E‐cadherin and N‐cadherin. J. Cell Biol. 118: 671–679, 1992.
 246. Kobayashi, T., S. Ohno, Y. C. Park‐Matsumoto, N. Kameda, and T. Baba. Developmental studies of dystrophin and other cytoskeletal proteins in cultured muscle cells. Microsc. Res. Tech. 30: 437–457, 1995.
 247. Koenig, M., A. P. Monaco, and L. M. Kunkel. The complete sequence of dystrophin predicts a rod‐shaped cytoskeletal protein. Cell 53: 219–226, 1988.
 248. Koob, R., M. Zimmermann, W. Schoner, and D. Drenckhahn. Colocalization and coprecipitation of ankyrin and Na +,K +‐ATPase in kidney epithelial cells. Eur. J. Cell Biol. 45: 230–237, 1988.
 249. Kordeli, E., J. Davis, B. Trapp, and V. Bennett. An isoform of ankyrin is localized at nodes of Ranvier in myelinated axons of central and peripheral nerves. J. Cell Biol. 110: 1341–1352, 1990.
 250. Kordeli, E., S. Lambert, and V. Bennett, Ankyrin, G. A new ankyrin gene with neural‐specific isoforms localized at the axonal initial segment and node of Ranvier. J. Biol. Chem. 270: 2352–2359, 1995.
 251. Korsgren, C., and C. M. Cohen. cDNA sequence, gene sequence, and properties of murine pallidin (band 4.2), the protein implicated in the murine pallid mutation. Genomics 21: 478–485, 1994.
 252. Korsgren, C., J. Lawler, S. Lambert, D. Speicher, and C. M. Cohen. Complete amino acid sequence and homologies of human erythrocyte membrane protein band 4.2. Proc. Natl. Acad. Sci. U.S.A. 87: 613–617, 1990.
 253. Kotula, L., T. M. DeSilva, D. W. Speicher, and P. J. Curtis. Functional characterization of recombinant human red cell α‐spectrin polypeptides containing the tetramer binding site. J. Biol. Chem. 268: 14788–14793, 1993.
 254. Kotula, L., L. D. Laury‐Kleintop, L. Showe, K. Sahr, A. J. Linnenbach, B. Forget, and P. J. Curtis. The exon‐intron organization of the human erythrocyte alpha‐spectrin gene. Genomics 9: 131–140, 1991.
 255. Koury, M. J., M. C. Bondurant, and S. S. Rana. Changes in erythroid membrane proteins during erythropoietin‐mediated terminal differentiation. J. Cell Physiol. 133: 438–448, 1987.
 256. Kralj‐Iglic, V., S. Svetina, and B. Zeks. The existence of non‐axisymmetric bilayer vesicle shapes predicted by the bilayer couple model. Eur. Biophys. J. 22: 97–103, 1993.
 257. Kralj‐Iglic, V., S. Svetina, and B. Zeks. Lateral distribution of membrane constituents and cellular shapes. Eur. Biophys. J. submitted, 1996.
 258. Kretsinger, R. H., and S. Nakayama. Evolution of EF‐hand calcium‐modulated proteins. IV. Exon shuffling did not determine the domain compositions of EF‐hand proteins. J. Mol. Evol. 36: 477–488, 1993.
 259. Kunimoto, M., E. Otto, and V. Bennett. A new 440‐kD isoform is the major ankyrin in neonatal rat brain. J. Cell Biol. 115: 1319–1331, 1991.
 260. Lambert, S., and V. Bennett. Postmitotic expression of ankyrinR and beta R‐spectrin in discrete neuronal populations of the rat brain. J. Neurosci. 13: 3725–3735, 1993.
 261. Lambert, S., H. Yu, J. T. Prchal, J. Lawler, P. Ruff, D. Speicher, M. C. Cheung, Y. W. Kan, and J. Palek. cDNA sequence for human erythrocyte ankyrin. Proc. Natl. Acad. Sci. U.S.A. 87: 1730–1734, 1990.
 262. Lande, W. M., and W. C. Mentzer. Haemolytic anaemia associated with increased cation permeability. Clin. Hematol. 14: 89–103, 1985.
 263. Lande, W. M., P. W. Thiemann, and W. M. Mentzer. Missing band 7 membrane protein in two patients with high Na, low K erythrocytes. J. Clin. Invest. 70: 1273–1280, 1982.
 264. Langley, R. C., Jr., and C. M. Cohen. Association of spectrin with desmin intermediate filaments. J. Cell Biochem. 30: 101–109, 1986.
 265. Laurila, P., L. Cioe, C. A. Kozak, and P. J. Curtis. Assignment of mouse beta‐spectrin gene to chromosome 12. Somat. Cell Mol. Genet. 13: 93–97, 1987.
 266. Lazarides, E., and W. J. Nelson. Erythrocyte and brain forms of spectrin in cerebellum: distinct membrane‐cytoskeletal domains in neurons. Science 220: 1295–1296, 1983.
 267. Lazarides, E., and C. Woods. Biogenesis of the red blood cell membrane‐skeleton and the control of erythroid morphogenesis. Annu. Rev. Cell Biol. 5: 427–452, 1989.
 268. Lee, J. K., J. D. Black, E. A. Repasky, R. T. Kubo, and R. B. Bankert. Activation induces a rapid reorganization of spectrin in lymphocytes. Cell 55: 807–816, 1988.
 269. Lee, J. K., R. S. Coyne, R. R. Dubreuil, L. S. Goldstein, and D. Branton. Cell shape and interaction defects in alpha‐spectrin mutants of Drosophila melanogaster. J. Cell Biol. 123: 1797–809, 1993.
 270. Lee, J. K., and E. A. Repasky. Cytoskeletal polarity in mammalian lymphocytes in situ. Cell Tissue Res. 247: 195–202, 1987.
 271. Lehnert, M. E., and H. F. Lodish. Unequal synthesis and differential degradation of alpha and beta spectrin during murine erythroid differentiation. J. Cell Biol. 107: 413–426, 1988.
 272. Leto, T. L., D. Fortugno‐Erikson, D. Barton, T. L. Yang‐Fong, U. Francke, A. S. Harris, J. S. Morrow, V. T. Marchesi, and E. J. Benz, Jr.. Comparison of non‐erythroid α‐spectrin genes reveals strict homology among diverse species. Mol. Cell. Biol. 8: 1–9, 1988.
 273. Leto, T. L., and V. T. Marchesi. A structural model of human erythrocyte protein 4.1. J. Biol. Chem. 259: 4603–4608, 1984.
 274. LeVine, H. I., and N. E. Sayoun. Involvement of fodrin‐binding proteins in the structure of the neuronal postsynaptic density and regulation by phosphorylation. Biochem. Biophys. Res. Commun. 138: 59–65, 1986.
 275. Levine, J., and M. Willard. Fodrin: axonally transported polypeptides associated with the internal periphery of many cells. J. Cell Biol. 90: 631–643, 1981.
 276. Li, X.‐L., and V. Bennett. Identification of the spectrin subunit and domains required for formation of adducin/spectrin/actin complexes. Mol. Biol. Cell 6: 269a (abstract), 1995.
 277. Li, Z. P., E. P. Burke, J. S. Frank, V. Bennett, and K. D. Philipson. The cardiac Na+ ‐Ca2+ exchanger binds to the cytoskeletal protein ankyrin. J. Biol. Chem. 268: 11489–11491, 1993.
 278. Lin, B., J. Nasir, H. McDonald, R. Graham, J. M. Rommens, Y. P. Goldberg, and M. R. Hayden. Genomic organization of the human alpha‐adducin gene and its alternately spliced isoforms. Genomics 25: 93–99, 1995.
 279. Liu, S. C., L. H. Derick, and J. Palek. Visualization of the hexagonal lattice in the erythrocyte membrane skeleton. J. Cell Biol. 104: 527–536, 1987.
 280. Lokeshwar, V. B., and L. Y. Bourguignon. The lymphoma transmembrane glycoprotein GP85 (CD44) is a novel guanine nucleotide‐binding protein which regulates GP85 (CD44)‐ankyrin interaction. J. Biol. Chem. 267: 22073–22078, 1992.
 281. Lokeshwar, V. B., and L. Y. Bourguignon. Post‐translational protein modification and expression of ankyrin‐binding site(s) in GP85 (Pgp‐1/CD44) and its biosynthetic precursors during T‐lymphoma membrane biosynthesis. J. Biol. Chem. 266: 17983–17989, 1991.
 282. Lokeshwar, V. B., and L. Y. Bourguignon. Tyrosine phosphatase activity of lymphoma CD45 (GP180) is regulated by a direct interaction with the cytoskeleton. J. Biol. Chem. 267: 21551–21557, 1992.
 283. Lokeshwar, V. B., N. Fregien, and L. Y. Bourguignon. Ankyrin‐binding domain of CD44(GP85) is required for the expression of hyaluronic acid‐mediated adhesion function. J. Cell Biol. 126: 1099–1109, 1994.
 284. Lombardo, C. R., and P. S. Low. Calmodulin modulates protein 4.1 binding to human erythrocyte membranes. Biochim. Biophys. Acta 1196: 139–144, 1994.
 285. Lombardo, C. R., D. L. Rimm, S. P. Kennedy, B. G. Forget, and J. S. Morrow. Ankyrin independent membrane sites for non‐erythroid spectrin. Mol. Biol. Cell 4 (suppl) 57a, 1993.
 286. Lombardo, C. R., D. L. Rimm, E. Koslov, and J. S. Morrow. Human recombinant alpha‐catenin binds to spectrin. Mol. Biol. Cell 5 (suppl): 47a, 1994.
 287. Lombardo, C. R., S. A. Weed, S. P. Kennedy, B. G. Forget, and J. S. Morrow. βII‐Spectrin (fodrin) and βIç2‐spectrin (muscle) contain NH2‐ & COOH‐terminal membrane association domains (MAD1 & MAD2). J. Biol. Chem. 269: 29212–29219, 1994.
 288. Lombardo, C. R., B. M. Willardson, and P. S. Low. Localization of the protein 4.1‐binding site on the cytoplasmic domain of erythrocyte membrane band 3. J. Biol. Chem. 267: 9540–9546, 1992.
 289. López, J. A., B. Leung, C. C. Reynolds, C. Q. Li, and J.E.B. Fox. Efficient plasma membrane expression of a functional platelet glycoprotein Ib‐IX complex requires the presence of its three subunits. J. Biol. Chem. 267: 12851–12859, 1992.
 290. Lorenz, M., B. Bisikirska, B. Hanus‐Lorenz, K. Strzalka, and A. F. Sikorski. Proteins reacting with anti‐spectrin antibodies are present in Chlamydomonas cells. Cell Biol Int. 19: 83–90, 1995.
 291. Low, P. S., D. P. Allen, T. F. Zioncheck, P. Chari, B. M. Willardson, R. L. Geahlen, and M. L. Harrison. Tyrosine phosphorylation of band 3 inhibits peripheral protein binding. J. Biol. Chem. 262: 4592–4596, 1987.
 292. Low, P. S., B. M. Willardson, N. Mohandas, M. Rossi, and S. Shohet. Contribution of the band 3‐ankyrin interaction to erythrocyte membrane mechanical stability. Blood 77: 1581–1586, 1991.
 293. Lowenstein, E. J., R. J. Daly, A. G. Batzer, W. Li, B. Margolis, R. Lammers, A. Ullrich, E. Y. Skolnik, D. Bar‐Sagi, and J. Schlessinger. The SH2 and SH3 domain‐containing protein GRB2 links receptor tyrosine kinases to ras signaling. Cell 70: 431–442, 1992.
 294. Lu, P. W., C. J. Soong, and M. Tao. Phosphorylation of ankyrin decreases its affinity for spectrin tetramer. J. Biol. Chem. 260: 14958–14964, 1985.
 295. Lue, R. A., S. M. Marfatia, D. Branton, and A. H. Chishti. Cloning and characterization of hdlg: the human homologue of the Drosophila discs large turnor suppressor binds to protein 4.1. Proc. Natl. Acad. Sci. U.S.A. 91: 9818–9822, 1994.
 296. Luna, E. J. Molecular links between the cytoskeleton and membranes. Curr. Opin. Cell Biol. 3: 120–126, 1991.
 297. Luna, E. J., and A. L. Hitt. Cytoskeleton‐plasma membrane interactions. Science 258: 955–964, 1992.
 298. Luna, E. J., G. H. Kidd, and D. Branton. Identification by peptide analysis of the spectrin‐binding protein in human erythrocytes. J. Biol. Chem. 254: 2526–2532, 1979.
 299. Lundberg, S., V. P. Lehto, and L. Backman. Characterization of calcium binding to spectrins. Biochemistry 31: 5665–5671, 1992.
 300. Lux, S. E., K. M. John, and V. Bennett. Analysis of cDNA for human erythrocyte ankyrin indicates a repeated structure with homology to tissue‐differentiation and cell‐cycle control proteins. Nature 344: 36–42, 1990.
 301. Lux, S. E., and J. Palek. Disorders of the red cell membrane. In: Blood: Principles and Practice of Hematology, edited by R. I. Handin, S. E. Lux and T. P. Stossel, Philadelphia: JB Lippincott, 1995, p. 1701–1816.
 302. Ma, Y., W. E. Zimmer, B. M. Riederer, M. L. Bloom, J. E. Barker, and S. R. Goodman. The complete amino acid sequence for brain beta spectrin (beta fodrin): relationship to globin sequences [published errata appear in Brain Res. Mol. Brain Res. 1993 Oct;20:179 and 1994 Jan;21:181]. Brain Res. Mol. Bain Res. 18: 87–99, 1993.
 303. MacDonald, R. I., A. Musacchio, R. A. Holmgren, and M. Saraste. Invariant tryptophan at a shielded site promotes folding of the conformational unit of spectrin. Proc. Natl. Acad. Sci. U.S.A. 91: 1299–1303, 1994.
 304. Macias, M. J., A. Musacchio, H. Ponstingl, M. Nilges, M. Saraste, and H. Oschkinat. Structure of the pleckstrin homology domain from beta‐spectrin. Nature 369: 675–677, 1994.
 305. Mak, A. S., G. Roseborough, and H. Baker. Tropomyosin from human erythrocyte membrane polymerizes poorly but binds F‐actin effectively in the presence and absence of spectrin. Biochim. Biophys. Acta 912: 157–166, 1987.
 306. Maksymiw, R., S. F. Sui, H. Gaub, and E. Sackmann. Electrostatic coupling of spectrin dimers to phosphatidylserine containing lipid lamellae. Biochemistry 26: 2983–2990, 1987.
 307. Malchiodi‐Albedi, F., M. Ceccarini, J. C. Winkelmann, J. S. Morrow, and T. C. Petrucci. The 270 kDa splice variant of erythrocyte β‐spectrin (βIç2) segregates in vivo and in vitro to specific domains of cerebellar neurons. J. Cell Sci. 106: 67–78, 1993.
 308. Mangeat, P. H., and K. Burridge. Immunoprecipitation of nonerythrocyte spectrin within live cells following microinjection of specific antibodies: relation to cytoskeletal structures. J. Cell Biol. 98: 1363–1377, 1984.
 309. Marchesi, S. L., Mutant cytoskeletal proteins in hemolytic disease. In: Ordering the membrane‐cytoskeleton trilayer, edited by M. S. Mooseker and J. S. Morrow, New York: Academic Press, 1991, p. 155–174.
 310. Marchesi, V. T., and G. E. Palade. Inactivation of adenosine triphosphatase and disruption of red cell membranes by trypsin: protective effect of adenosine triphosphate. Proc. Natl. Acad. Sci. U.S.A. 58: 991–995, 1967.
 311. Marchesi, V. T., and E. Steers. Selective solubilization of a protein component of the red cell membrane. Science 159: 203–204, 1968.
 312. Marchesi, V. T., T. W. Tillack, R. L. Jackson, J. P. Segrest, and R. E. Scott. Chemical characterization and surface orientation of the major glycoprotein of the human erythrocyte membrane. Proc. Natl. Acad. Sci. U.S.A. 69: 1445–1449, 1972.
 313. Marfatia, S. M., R. A. Leu, D. Branton, and A. H. Chishti. Identification of the protein 4.1 binding interface on glycophorin C and p55, a homologue of the Drosophila discs‐large tumor suppressor protein. J. Biol. Chem. 270: 715–719, 1995.
 314. Marfatia, S. M., R. A. Lue, D. Branton, and A. H. Chishti. In vitro binding studies suggest a membrane‐associated complex between erythroid p55, protein 4.1, and glycophorin C. J. Biol. Chem. 269: 8631–8634, 1994.
 315. Mariani, M., D. Maretzki, and H. U. Lutz. A tightly membrane‐associated subpopulation of spectrin is 3H‐palmitoylated. J. Biol. Chem. 268: 12996–3001, 1993.
 316. Markin, V. S. Lateral organization of membranes and cell shapes. Biophys. J. 36: 1–19, 1981.
 317. Marrs, J. A., C. Andersson‐Fisone, M. C. Jeong, L. Cohen‐Gould, C. Zurzolo, I. R. Nabi, E. Rodriguez‐Boulan, and W. J. Nelson. Plasticity in epithelial cell phenotype: modulation by expression of different cadherin cell adhesion molecules. J. Cell Biol. 129: 507–519, 1995.
 318. Masuda, T., N. Fujimaki, E. Ozawa, and H. Ishikawa. Confocal laser microscopy of dystrophin localization in guinea pig skeletal muscle fibers. J. Cell Biol. 119: 543–548, 1992.
 319. Matsudaira, P. Modular organization of actin crosslinking proteins. Trends Biochem. Sci. 16: 87–92, 1991.
 320. Matsuoka, Y., C. A. Hughes, and V. Bennett. Spectin/actin assembly activity of adducin is regulated by phosphorylation of the MARCKS‐related domain by protein kinase C. Mol. Cell. Biol. 6: 269a (abstract), 1995.
 321. Matsuyoshi, N., M. Hamaguchi, S. Taniguchi, A. Nagafuchi, S. Tsukita, and M. Takeichi. Cadherin‐mediated cell‐cell adhesion is perturbed by v‐src tyrosine phosphorylation in metastatic fibroblasts. J. Cell Biol. 118: 703–714, 1992.
 322. Mayer, B. J., R. Ren, K. L. Clark, and D. Baltimore. A putative modular domain present in diverse signaling proteins [letter]. Cell 73: 629–630, 1993.
 323. McGough, A. M., and R. Josephs. On the structure of erythrocyte spectrin in partially expanded membrane skeletons. Proc. Natl. Acad. Sci. U.S.A. 87: 5208–5212, 1990.
 324. McManus, M. J., D. C. Connolly, and N. J. Maihle. Tissue‐and transformation‐specific phosphotyrosyl proteins in v‐erbB‐tranformed cells. J. Virol. 69: 3631–3638, 1995.
 325. McNeill, H., M. Ozawa, R. Kemler, and W. J. Nelson. Novel function of the cell adhesion molecule uvomorulin as an inducer of cell surface polarity. Cell 62: 309–316, 1990.
 326. McNeill, H., T. A. Ryan, S. J. Smith, and W. J. Nelson. Spatial and temporal dissection of immediate and early events following cadherin‐mediated epithelial cell adhesion. J. Cell Biol. 120: 1217–1226, 1993.
 327. Mentzer, W. C., B. H. Lubin, and S. Emmons. Correction of the permeability defect in hereditary stomatocytosis by dimethyl adipimate. N. Engl. J. Med. 294: 1200–1205, 1976.
 328. Mercier, F., H. Reggio, G. Devilliers, D. Bataille, and P. Manget. Membrane‐cytoskeleton dynamics in rat parietal cells: mobilization of actin and spectrin upon stimulation of gastric acid secretion. J. Biol. Chem. 108: 441–453, 1989.
 329. Michaely, P., and V. Bennett. The membrane‐binding domain of ankyrin contains four independently folded subdomains, each comprised of six ankyrin repeats. J. Biol. Chem. 268: 22703–22709, 1993.
 330. Michalak, K., M. Bobrowska, K. Bialkowska, J. Szopa, and A. F. Sikorski. Interaction of erythrocyte spectrin with some nonbilayer phospholipids. Gen. Physiol. Biophys. 13: 57–62, 1994.
 331. Michalak, K., M. Bobrowska, and A. F. Sikorski. Interaction of bovine erythrocyte spectrin with aminophospholipid liposomes. Gen. Physiol. Biophys. 12: 163–170, 1993.
 332. Michalak, K., M. Bobrowska, and A. F. Sikorski. Investigation of spectrin binding to phospholipid vesicles using a isoindole fluorescent probe. Thermal properties of the bound and unbound protein. Gen. Physiol. Biophys. 9: 615–624, 1990.
 333. Michaud, D., G. Guillet, P. A. Rogers, and P. M. Charest. Identification of a 220 kDa membrane‐associated plant cell protein immunologically related to human beta‐spectrin. FEBS Lett. 294: 77–80, 1991.
 334. Minetti, C., F. Bentrame, G. Marcenaro, and E. Bonilla. Dystrophin at the plasma membrane of human muscle fibers shows a costameric localization. Neuromusc. Dis. 2: 99–109, 1992.
 335. Mische, S. M., M. S. Mooseker, and J. S. Morrow. Erythrocyte adducin: a calmodulin regulated actin bundling protein that stimulates spectrin‐actin binding. J. Cell Biol. 105: 2837–2845, 1987.
 336. Mische, S. M., and J. S. Morrow. Post‐translational regulation of the erythrocyte cortical cytoskeleton. Protoplasma 145: 167–175, 1988.
 337. Mohandas, N. Molecular basis for red cell membrane viscoelastic properties. Biochem. Soc. Trans. 20: 776–782, 1992.
 338. Mohandas, N., and J. A. Chasis. Red blood cell deformability, membrane material properties and shape: regulation by transmembrane, skeletal and cytosolic proteins and lipids. Semin Hematol. 30: 171–192, 1993.
 339. Mohandas, N., and E. Evans. Mechanical properties of the red cell membrane in relation to molecular structure and genetic defects. Annu. Rev. Biophys. Biomol. Struct. 23: 787–818, 1994.
 340. Mohandas, N., A. C. Greenquist, and S. B. Shohet. Bilayer balance and regulation of red cell shape changes. J. Supramolecular Structure Cellular Biochem. 9: 453–458, 1978.
 341. Mohandas, N., R. Winardi, D. Knowles, A. Leung, M. Parra, E. George, J. Conboy, and J. Chasis. Molecular basis for membrane rigidity of hereditary ovalocytosis: A novel mechanism involving the cytoplasmic domain of band 3. J. Clin. Invest. 89: 686–692, 1992.
 342. Molday, L. L., N. J. Cook, U. B. Kaupp, and R. S. Molday. The cGMP‐gated cation channel of bovine rod photoreceptor cells is associated with a 240‐kDa protein exhibiting immunochemical cross‐reactivity with spectrin. J. Biol. Chem. 265: 18690–18695, 1990.
 343. Mombers, C., P. W. van Dijck, L. L. van Deenen, J. de Gier, and A. J. Verkleij. The interaction of spectrin‐actin and synthetic phospholipids. Biochim. Biophys. Acta 470: 152–160, 1977.
 344. Mombers, C., A. J. Verkleij, J. de Gier, and L. L. van Deenen. The interaction of spectrin‐actin and synthetic phospholipids. II. The interaction with phosphatidylserine. Biochim. Biophys. Acta 551: 271–281, 1979.
 345. Moon, R. T., and E. Lazarides. beta‐Spectrin limits alpha‐spectrin assembly on membranes following synthesis in a chicken erythroid cell lysate. Nature 305: 62–65, 1983.
 346. Moon, R. T., and A. P. McMahon. Generation of diversity in nonerythroid spectrins. Multiple polypeptides are predicted by sequence analysis of cDNAs encompassing the coding region of human nonerythroid alpha‐spectrin. J. Biol. Chem. 265: 4427–4433, 1990.
 347. Moon, R. T., J. Ngai, B. J. Wold, and E. Lazarides. Tissue‐specific expression of distinct spectrin and ankyrin transcripts in erythroid and nonerythroid cells. J. Cell Biol. 100: 152–160, 1985.
 348. Morgans, C. W., and R. R. Kopito. Association of the brain anion exchanger, AE3, with the repeat domain of ankyrin. J. Cell Sci. 105: 1137–1142, 1993.
 349. Moriyama, R., C. R. Lombardo, R. F. Workman, and P. S. Low. Regulation of linkages between the erythrocyte membrane and its skeleton by 2,3‐diphosphoglycerate. J. Biol. Chem. 268: 10990–10996, 1993.
 350. Morris, M. B., and S. E. Lux. Characterization of the binary interaction between human erythrocyte protein 4.1 and actin. Eur. J. Biochem. 231: 644–650, 1995.
 351. Morrow, J. S., and R. A. Anderson. Shaping the too fluid bilayer. Lab. Invest. 54: 237–240, 1986.
 352. Morrow, J. S., C. Cianci, T. Ardito, A. Mann, and M. T. Kashgarian. Ankyrin links fodrin to alpha Na/K ATPase in Madin‐Darby canine kidney cells and in renal tubule cells. J. Cell Biol. 108: 455–465, 1989.
 353. Morrow, J. S., C. D. Cianci, S. P. Kennedy, and S. L. Warren. Polarized assembly of spectrin and ankyrin in epithelial cells. In: Ordering the Membrane Cytoskeleton Trilayer, edited by M. S. Mooseker and J. S. Morrow, New York: Academic Press, 1991, p. 227–244.
 354. Morrow, J. S., W. Haigh, and V. T. Marchesi. Spectrin oligomers: a structural feature of the erythrocyte cytoskeleton. J. Supra. Struc. Cell Biochem. 17: 275–287, 1981.
 355. Morrow, J. S., A. S. Harris, P. E. Shile, L. A. D. Green, and K. J. Ainger. Structural and functional domains of human brain and muscle spectrin: relationship to human erythrocyte spectrin. Proceedings of the International Symposium on Contractile Proteins in Muscle and Non‐Muscle Cell Systems and Their Morphophysiopathology (Sassari, Italy). 1: 125–133, 1985.
 356. Morrow, J. S., and V. T. Marchesi. Self‐assembly of spectrin oligomers in vitro: a basis for a dynamic cytoskeleton. J. Cell Biol. 88: 463–468, 1981.
 357. Morrow, J. S., D. W. Speicher, W. J. Knowles, C. J. Hsu, and V. T. Marchesi. Identification of functional domains of human erythrocyte spectrin. Proc. Natl. Acad. Sci. U.S.A. 77: 6592–6596, 1980.
 358. Muller, B. M., U. Kistner, R. W. Veh, C. Cases‐Langhoff, B. Becker, E. D. Gundelfinger, and C. C. Garner. Molecular characterization and spatial distribution of SAP97, a novel presynaptic protein homologous to SAP90 and the Drosophila discs‐large tumor suppressor protein. J. Neurosci. 15: 2354–2366, 1995.
 359. Murachi, T. Intracellular regulatory system involving calpain and calpastatin. Biochem. Int. 18: 263–294, 1989.
 360. Musacchio, A., T. Gibson, V. P. Lehto, and M. Saraste. SH3—an abundant protein domain in search of a function. FEBS Lett. 307: 55–61, 1992.
 361. Musacchio, A., M. Noble, R. Pauptit, R. Wierenga, and M. Saraste. Crystal structure of a Src‐homology 3 (SH3) domain. Nature 359: 851–855, 1992.
 362. Nagafuchi, A., S. Ishihara, and S. Tsukita. The roles of catenins in the cadherin‐mediated cell adhesion: functional analysis of E‐cadherin‐alpha catenin fusion molecules. J. Cell Biol. 127: 235–245, 1994.
 363. Nagafuchi, A., Y. Shirayoshi, K. Okazaki, K. Yasuda, and M. Takeichi. Transformation of cell adhesion properties by exogenously introduced E‐cadherin cDNA. Nature 329: 341–343, 1987.
 364. Nagafuchi, A., and M. Takeichi. Cell binding function of E‐cadherin is regulated by the cytoplasmic domain. Eur. Mol. Biol. Organization J. 7: 3679–3684, 1988.
 365. Nakashima, K., and E. Beutler. Comparison of structure and function of human erythrocyte and human muscle actin. Proc. Natl. Acad. Sci. U.S.A. 76: 935–938, 1979.
 366. Nakata, T., A. Iwamoto, Y. Noda, R. Takemura, H. Yoshikura, and N. Hirokawa. Predominant and developmentally regulated expression of dynamin in neurons. Neuron 7: 461–469, 1991.
 367. Nakayama, S., N. D. Moncrief, and R. H. Kretsinger. Evolution of EF‐hand calcium‐modulated proteins. II. Domains of several subfamilies have diverse evolutionary histories. J. Mol. Evol. 34: 416–48, 1992.
 368. Nathke, I. S., L. Hinck, J. R. Swedlow, J. Papkoff, and W. J. Nelson. Defining interactions and distributions of cadherin and catenin complexes in polarized epithelial cells. J. Cell Biol. 125: 1341–1352, 1994.
 369. Nelson, W. J. Regulation of cell surface polarity in renal epithelia. Pediatr. Nephrol. 7: 599–604, 1993.
 370. Nelson, W. J., and R. W. Hammerton. A membranecytoskeletal complex containing Na+,K+‐ATPase, ankyrin, and fodrin in Madin‐Darby canine kidney (MDCK) cells: implications for the biogenesis of epithelial cell polarity. J. Cell Biol. 108: 893–902, 1989.
 371. Nelson, W. J., and E. Lazarides. Expression of the beta subunit of spectrin in nonerythroid cells. Proc. Natl. Acad. Sci. U.S.A. 80: 363–367, 1983.
 372. Nelson, W. J., and E. Lazarides. Goblin (ankyrin) in striated muscle: identification of the potential membrane receptor for erythroid spectrin in muscle cells. Proc. Natl. Acad. Sci. U.S.A. 81: 3292–3296, 1984.
 373. Nelson, W. J., and E. Lazarides. Posttranslational control of membrane‐skeleton (ankyrin and alpha beta‐spectrin) assembly in early myogenesis. J. Cell Biol. 100: 1726–1735, 1985.
 374. Nelson, W. J., and E. Lazarides. Switching of subunit composition of muscle spectrin during myogenesis in vitro. Nature 304: 364–368, 1983.
 375. Nelson, W. J., and P. J. Veshnock. Ankyrin binding to (Na+ + K+) ATPase and implications for the organization of membrane domains in polarized cells. Nature 328: 533–536, 1987.
 376. Nelson, W. J., and P. J. Veshnock. Dynamics of membrane‐skeleton (fodrin) organization during development of polarity in Madin‐Darby canine kidney epithelial cells. J. Cell Biol. 103: 1751–1765, 1986.
 377. Nose, A., K. Tsuji, and M. Takeichi. Localization of specificity determining sites in cadherin cell adhesion molecules. Cell 61: 147–155, 1990.
 378. Obar, R. A., C. A. Collins, J. A. Hammarback, H. S. Shpetner, and R. B. Vallee. Molecular cloning of the microtubule‐associated mechanochemical enzyme dynamin reveals homology with a new family of GTP‐binding proteins. Nature 347: 256–261, 1990.
 379. Ochiai, A., S. Akimoto, Y. Kanai, T. Shibata, T. Oyama, and S. Hirohashi. c‐erbB‐2 gene product associates with catenins in human cancer cells. Biochem. Biophys. Res. Commun. 205: 73–78, 1994.
 380. Ohanian, V., L. C. Wolfe, K. M. John, J. C. Pinder, S. E. Lux, and W. B. Gratzer. Analysis of the ternary interaction of the red cell membrane skeletal proteins spectrin, actin, and 4.1. Biochemistry 23: 4416–4420, 1984.
 381. Omann, G. M., W. N. Swann, Z. G. Oades, C. A. Parkos, A. J. Jesaitis, and L. A. Sklar. N‐formylpeptide‐receptor dynamics, cytoskeletal activation, and intracellular calcium response in human neutrophil cytoplasts. J. Immunol. 139: 3447–3455, 1987.
 382. Otsuka, A. J., R. Franco, B. Yang, K. H. Shim, L. Z. Tang, Y. Y. Zhang, P. Boontrakulpoontawee, A. Jeyaprakash, E. Hedgecock, V. I. Wheaton, et al. An ankyrin‐related gene (unc‐44) is necessary for proper axonal guidance in Caenorhabditis elegans. J. Cell Biol. 129: 1081–1092, 1995.
 383. Otto, E., M. Kunimoto, T. McLaughlin, and V. Bennett. Isolation and characterization of cDNAs encoding human brain ankyrins reveal a family of alternatively spliced genes. J. Cell Biol. 114: 241–253, 1991.
 384. Ozawa, M., H. Baribault, and R. Kemler. The cytoplasmic domain of the cell adhesion molecule uvomorulin associates with three independent proteins structurally related in different species. Eur. Mol. Biol. Org. J. 8: 1711–1717, 1989.
 385. Ozawa, M., and R. Kemler. Molecular organization of the uvomorulin‐catenin complex. J. Cell Biol. 116: 989–996, 1992.
 386. Ozawa, M., M. Ringwald, and R. Kemler. Uvomorulin‐catenin complex formation is regulated by a specific domain in the cytoplasmic region of the cell adhesion molecule. Proc. Natl. Acad. Sci. U.S.A. 87: 4246–4250, 1990.
 387. Palfrey, H. C., and A. Waseem. Protein kinase C in the human erythrocyte. Translocation to the plasma membrane and phosphorylation of bands 4.1 and 4.9 and other membrane proteins. J. Biol. Chem. 260: 16021–16029, 1985.
 388. Pardo, J. V., J. D'Angelo Siliciano, and S. W. Craig. A vinculin‐containing cortical lattice in skeletal muscle: transverse lattice elements (“costameres”) mark sites of attachment between myofibrils and sarcolemma. Proc. Natl. Acad. Sci. U.S.A. 80: 1008–1012, 1983.
 389. Pardo, J. V., J. D. Siliciano, and S. W. Craig. Vinculin is a component of an extensive network of myofibril‐sarcolemma attachment regions in cardiac muscle fibers. J. Cell Biol. 97: 1081–1088, 1983.
 390. Pasternack, G. R., R. A. Anderson, T. L. Leto, and V. T. Marchesi. Interactions between protein 4.1 and band 3. An alternative binding site for an element of the membrane skeleton. J. Biol. Chem. 260: 3676–3683, 1985.
 391. Pasternack, G. R., and R. H. Racusen. Erythrocyte protein 4.1 binds and regulates myosin. Proc. Natl. Acad. Sci. U.S.A. 86: 9712–9716, 1989.
 392. Pauly, J. L., R. B. Bankert, and E. A. Repasky. Immunofluorescent patterns of spectrin in lymphocyte cell lines. J. Immunol. 136: 246–253, 1986.
 393. Pawson, T., and J. Schlessinger. SH2 and SH3 domains. Current Bio. 3: 434–442, 1993.
 394. Peifer, M., S. Berg, and A. B. Reynolds. A repeating amino acid motif shared by proteins with diverse cellular roles [letter]. Cell 76: 789–791, 1994.
 395. Piefer, M., P. D. McCrea, K. J. Green, E. Wieschaus, and B. M. Gumbiner. The vertebrate adhesive junction proteins beta‐catenin and plakoglobin and the Drosophila segment polarity gene armadillo form a multigene family with similar properties. J. Cell Biol. 118: 681–691, 1992.
 396. Perrin, D., and H. D. Soling. No evidence for calpain I involvement in fodrin rearrangements linked to regulated secretion. FEBS Lett. 311: 302–304, 1992.
 397. Pesacreta, T. C., T. J. Byers, R. Dubreuil, D. P. Kiehart, and D. Branton. Drosophila spectrin: the membrane skeleton during embryogenesis. J. Cell Biol. 108: 1697–709, 1989.
 398. Peters, L. L., E. M. Eicher, A. C. Azim, and A. H. Chishti. The gene encoding the erythrocyte membrane skeleton protein dematin (Epb4.9) maps to mouse chromosome 14. Genomics 26: 634–635, 1995.
 399. Peters, L. L., E. M. Eicher, T. C. Hoock, K. M. John, M. Yialamas, and S. E. Lux. Evidence that the mouse platelet storage pool deficiency mutation mocha is a defect in a new member of the ankyrin gene family. Blood 82: 340a (abstract), 1993.
 400. Peters, L. L., K. M. John, F. M. Lu, E. M. Eicher, A. Higgins, M. Yialamas, L. C. Turtzo, A. J. Otsuka, and S. E. Lux. Ank3 (epithelial ankyrin), a widely distributed new member of the ankyrin gene family and the major ankyrin in kidney, is expressed in alternatively spliced forms, including forms that lack the repeat domain. J. Cell Biol. 130: 313–330, 1995.
 401. Peters, L. L., and S. E. Lux. Ankyrins: structure and function in normal cells and hereditary spherocytes. Semin. Hematol. 30: 85–118, 1993.
 402. Peters, L. L., R. A. Shivdasani, S.‐C. Liu, M. Hanspal, K. M. John, J. M. Gonzalez, C. Brugnara, B. Gwynn, N. Mohandas, S. L. Alper, S. H. Orkin, and S. E. Lux. Anion exchanger 1 (Band 3) is required to prevent erythrocyte membrane surface loss but not to form the membrane skeleton. Cell 86: 917–927, 1996.
 403. Petrucci, T. C., and J. S. Morrow. Synapsin‐I: An actin bundling protein under phosphorylation control. J. Cell Biol. 105: 1355–1363, 1987.
 404. Petrucci, T. P., and J. S. Morrow. The actin and tubulin binding domains of Synapsin la and lb. Biochemistry 30: 413–422, 1990.
 405. Pinto‐Correia, C., E. G. Goldstein, V. Bennett, and J. S. Sobel. Immunofluorescence localization of an adducin‐like protein in the chromosomes of mouse oocytes. Dev. Biol. 146: 301–311, 1991.
 406. Pitcher, J. A., J. Inglese, J. B. Higgins, J. L. Arriza, P. J. Casey, C. Kim, J. L. Benovic, M. M. Kwatra, M. G. Caron, and R. J. Lefkowitz. Role of beta gamma subunits of G proteins in targeting the beta‐adrenergic receptor kinase to membranebound receptors. Science 257: 1264–1267, 1992.
 407. Platt, O. S., S. E. Lux, and J. F. Falcone. A highly conserved region of human erythrocyte ankyrin contains the capacity to bind spectrin. J. Biol. Chem. 268: 24421–24426, 1993.
 408. Pollard, T. D. Purification of a high molecular weight actin filament gelation protein from Acanthamoeba that shares antigenic determinants with vertebrate spectrins. J. Cell Biol. 99: 1970–1980, 1984.
 409. Pollerberg, G. E., K. Burridge, K. E. Krebs, S. R. Goodman, and M. Schachner. The 180‐kD component of the neural cell adhesion molecule N‐CAM is involved in a cell‐cell contacts and cytoskeleton‐membrane interactions. Cell Tissue Res. 250: 227–236, 1987.
 410. Ponting, C. P., and C. Philips. DHR domains in syntrophins, neuronal NO synthases and other intracellular proteins. Trends Biochem. Sci. 20: 102–103, 1995.
 411. Porter, G. A., The membrane skeleton of costameres. In: Physiology. Baltimore: University of Maryland School of Medicine, 1993.
 412. Porter, G. A., G. M. Dmytrenko, J. C. Winkelmann, and R. J. Bloch. Dystrophin colocalizes with β‐spectrin in distinct subsarcolemmal domains in mammalian skeletal muscle. J. Cell Biol. 117: 997–1005, 1992.
 413. Pradhan, D., C. Lombardo, P. R. Stabach, and J. S. Morrow. Surface plasmon resonance detects a strong heterodimer nucleation site in domain I of beta‐spectrin. in preparation, 1996.
 414. Pradhan, D., P. Williamson, and R. A. Schlegel. Bilayer/cytoskeleton interactions in lipid‐symmetric erythrocytes assessed by a photoactivatable phospholipid analogue. Biochemistry 30: 7754–7758, 1991.
 415. Prchal, J. T., T. Papayannopoulou, and S.‐H. Yoon. Patterns of spectrin transcripts in erythroid and non‐erythroid cells. J. Cell Physiol. 144: 287–294, 1990.
 416. Pumplin, D. W. The membrane skeleton of acetylcholine receptor domains in rat myotubes contains antiparallel homodimers of β‐spectrin in filaments quantitatively resembling those of erythrocytes. J. Cell Sci. 108: 3145–3154, 1995.
 417. Quinn, M. T., C. A. Parkos, and A. J. Jesaitis. The lateral organization of components of the membrane skeleton and superoxide generation in the plasma membrane of stimulated human neutrophils. Biochim. Biophys. Acta 987: 83–94, 1989.
 418. Ralston, G. B. The concentration dependence of the activity coefficient of the human spectrin heterodimer. A quantitative test of the Adams‐Fujita approximation. Biophys. Chem. 52: 51–61, 1994.
 419. Rana, A. P., P. Ruff, G. J. Maalouf, D. W. Speicher, and A. H. Chishti. Cloning of human erythroid dematin reveals another member of the villin family. Proc. Natl. Acad. Sci. U.S.A. 90: 6651–6655, 1993.
 420. Reid, M. E., Y. Takakuwa, J. Conboy, G. Tchernia, and N. Mohandas. Glycophorin C content of human erythrocyte membrane is regulated by protein 4.1. Blood 75: 2229–2234, 1990.
 421. Reid, M. E., Y. Takakuwa, G. Tchernia, R. H. Jensen, J. A. Chasis, and N. Mohandas. Functional role for glycophorin C and its interaction with the human red cell membrane skeletal component, protein 4.1. Prog. Clin. Biol. Res. 319: 553–571; discussion 572–578, 1989.
 422. Ren, R., B. J. Mayer, P. Cicchetti, and D. Baltimore. Identification of a ten‐amino acid proline‐rich SH3 binding site. Science 259: 1157–1161, 1993.
 423. Reynolds, A. B., J. Daniel, P. D. McCrea, M. J. Wheelock, J. Wu, and Z. Zhang. Identification of a new catenin: the tyrosine kinase substrate p120cas associates with E‐cadherin complexes. Mol. Cell Biol. 14: 8333–8342, 1994.
 424. Reynolds, A. B., L. Herbert, J. L. Cleveland, S. T. Berg, and J. R. Gaut. p120, a novel substrate of protein tyrosine kinase receptors and of p60v‐src, is related to cadherin‐binding factors beta‐catenin, plakoglobin and armadillo. Oncogene 7: 2439–2445, 1992.
 425. Ridley, A. J., and A. Hall. The small GTP‐binding protein rho regulates the assembly of focal adhesions and actin stress fibers in response to growth factors. Cell 70: 389–399, 1992.
 426. Riederer, B. M., and S. R. Goodman. Association of brain spectrin isoforms with microtubules. FEBS Lett. 277: 49–52, 1990.
 427. Riederer, B. M., I. S. Zagon, and S. R. Goodman. Brain spectrin (240/235) and brain spectrin (240/235E): two distinct spectrin subtypes with different locations within mammalian neural cells. J. Cell Biol. 102: 2088–2096, 1986.
 428. Rimm, D. L., E. R. Koslov, P. Kebriaei, C. D. Cianci, and J. S. Morrow. α1(E)‐catenin is a novel actin binding and bundling protein mediating the attachment of F‐actin to the membrane adhesion complex. Proc. Natl. Acad. Sci. U.S.A. 92: 8813–8817, 1995.
 429. Rimm, D. L., and J. S. Morrow. Molecular cloning of human E‐cadherin suggests a novel subdivision of the cadherin super‐family. Biochem. Biophys. Res. Commun. 200: 1754–1761, 1994.
 430. Robinson, M. S. The role of clathrin, adaptors and dynamin in endocytosis. Curr. Opin. Cell Biol. 6: 538–544, 1994.
 431. Rodriguez‐Boulan, E., and W. J. Nelson. Morphogenesis of the polarized epithelial cell phenotype. Science 245: 718–725, 1989.
 432. Roof, D., A. Hayes, G. Hardenbergh, and M. Adamian. A 52 kD cytoskeletal protein from retinal rod photoreceptors is related to erythrocyte dematin. Invest. Ophthalmol. Vis. Sci. 32: 582–593, 1991.
 433. Rotin, D., D. Bar‐Sagi, H. O'Brodovich, J. Merilainen, V. P. Lehto, C. M. Canessa, B. C. Rossier, and G. P. Downey. An SH3 binding region in the epithelial Na+ channel (alpha rE‐NaC) mediates its localization at the apical membrane. EMBO J 13: 4440–4450, 1994.
 434. Rozakis, A. M., J. McGlade, G. Mbamalu, G. Pelicci, R. Daly, W. Li, A. Batzer, S. Thomas, J. Brugge, P. G. Pelicci, et al. Association of the Shc and Grb2/Sem5 SH2‐containing proteins is implicated in activation of the Ras pathway by tyrosine kinases. Nature 360: 689–692, 1992.
 435. Rubenfeld, B., B. Souza, I. Albert, O. Muller, S. H. Chamberlain, F. R. Masiarz, S. Munemitsu, and P. Polakis. Association of the APC gene product with beta‐catenin. Science 262: 1731–1734, 1993.
 436. Rubinfeld, B., I. Albert, B. Souza, S. Munemitsu, and P. Polakis. Interaction of the APC tumor suppressor protein with catenins. Mol. Biol. Cell 6: 117a, 1995.
 437. Ruff, P., D. W. Speicher, and A. Husain‐Chishti. Molecular identification of a major palmitoylated erythrocyte membrane protein containing the src homology 3 motif. Proc. Natl. Acad. Sci. U.S.A. 88: 6595–6599, 1991.
 438. Sacco, P. A., T. M. McGranahan, M. J. Wheelock, and K. R. Johnson. Identification of plakoglobin domains required for association with N‐cadherin and alpha‐catenin. J. Biol. Chem. 270: 20201–20206, 1995.
 439. Sahr, K. E., P. Laurila, L. Kotula, A. L. Scarpa, E. Coupal, T. L. Leto, A. J. Linnenbach, J. C. Winkelmann, D. W. Speicher, V. T. Marchesi, P. J. Curtis, and B. G. Forget. The complete cDNA and polypeptide sequences of human erythroid α‐spectrin. J. Biol. Chem. 265: 4434–4443, 1990.
 440. Salardi, S., R. Modica, M. Ferrandi, P. Ferrari, L. Torielli, and G. Bianchi. Characterization of erythrocyte adducin from the Milan hypertensive strain of rats. J. Hypertens. 64: S196–S198, 1988.
 441. Salardi, S., B. Saccardo, G. Borsani, R. Modica, M. Ferrandi, M. G. Tripodi, M. Soria, P. Ferrari, F. E. Baralle, A. Sidoli, and G. Bianchi. Erythrocyte adducin differential properties in the normotensive and hypertensive rats of the milan strain. Characterization of spleen adducin m‐RNA. Am. J. Hypertens. 2: 229–237, 1989.
 442. Salzer, U., H. Ahorn, and R. Prohaska. Identification of the phosphorylation site on human erythrocyte band 7 integral membrane protein: implications for a monotopic protein structure. Biochim. Biophys. Acta 1151: 149–152, 1993.
 443. Sato, N., N. Funayama, A. Nagafuchi, S. Yonemura, S. Tsukita, and S. Tsukita. A gene family consisting of ezrin, radixin and moesin. Its specific localization at actin filament/plasma membrane association sites. J. Cell Sci. 103: 131–143, 1992.
 444. Scaife, R., and R. L. Margolis. Biochemical and immunochemical analysis of rat brain dynamin interaction with microtubules and organelles in vivo and in vitro. J. Cell Biol. 111: 3023–3033, 1990.
 445. Scaramuzzino, D. A., and J. S. Morrow. Calmodulin‐binding domain of recombinant erythrocyte beta‐adducin. Proc. Natl. Acad. Sci. U.S.A. 90: 3398–3402, 1993.
 446. Schafer, D. A., S. R. Gill, J. A. Cooper, J. E. Heuser, and T. A. Schroer. Ultrastructural analysis of the dynactin complex: an actin‐related protein is a component of a filament that resembles F‐actin. J. Cell Biol. 126: 403–412, 1994.
 447. Schneider, R. The human protooncogene ret: a communicative cadherin. Trends Biochem. 17: 468–469, 1992.
 448. Schofield, A. E., M. J. Tanner, J. C. Pinder, B. Clough, P. M. Bayley, G. B. Nash, A. R. Dluzewski, D. M. Reardon, T. M. Cox, R. J. Wilson, et al. Basis of unique red cell membrane properties in hereditary ovalocytosis. J. Mol. Biol. 223: 949–958, 1992.
 449. Schroer, T. A., E. Fyrberg, J. A. Cooper, R. H. Waterston, D. Helfman, T. D. Pollard, and D. I. Meyer. Actin‐related protein nomenclature and classification. J. Cell Biol. 127: 1777–1778, 1994.
 450. Sears, D. E., V. T. Marchesi, and J. S. Morrow. A calmodulin and α‐subunit binding domain in human erythrocyte beta‐spectrin. Biochim. Biophys. Acta 870: 432–442, 1986.
 451. Seubert, P., M. Baudry, S. Dudek, and G. Lynch. Calmodulin stimulates the degradation of brain spectrin by calpain. Synapse 1: 20–24, 1987.
 452. Shapiro, L., A. M. Fannon, P. D. Kwong, A. Thompson, M. S. Lehmann, G. Grubel, J. F. Legrand, N. J. Als, D. R. Colman, and W. A. Hendrickson. Structural basis of cell‐cell adhesion by cadherins. Nature 374: 327–337, 1995.
 453. Sheetz, M. P. DNase‐I‐dependent dissociation of erythrocyte cytoskeletons. J. Cell Biol. 81: 266–270, 1979.
 454. Sheetz, M. P., and S. J. Singer. Biological membranes as bilayer couples. A molecular mechanism of drug‐erythrocyte interactions. Proc. Natl. Acad. Sci. U.S.A. 71: 4457–4461, 1974.
 455. Shen, B. W., R. Josephs, and T. L. Steck. Ultrastructure of the intact skeleton of the human erythrocyte membrane. J. Cell Biol. 102: 997–1006, 1986.
 456. Shibamoto, S., M. Hayakawa, K. Takeuchi, T. Hori, K. Miyazawa, N. Kitamura, K. R. Johnson, M. J. Wheelock, N. Matsuyoshi, M. Takeichi, et al. Association of p120, a tyrosine kinase substrate, with E‐cadherin/catenin complexes. J. Cell Biol. 128: 949–957, 1995.
 457. Shimkets, R. A., D. G. Warnock, C. M. Bositis, C. Nelson‐Williams, J. H. Hansson, M. Schambelan, J. R. Gill, Jr., S. Ulick, R. V. Milora, J. W. Findling, R. Lifton, et al. Liddle's syndrome: heritable human hypertension caused by mutations in the beta subunit of the epithelial sodium channel. Cell 79: 407–414, 1994.
 458. Shpetner, H. S., and R. B. Vallee. Dynamin is a GTPase stimulated to high levels of activity by microtubules. Nature 355: 733–735, 1992.
 459. Shpetner, H. S., and R. B. Vallee. Identification of dynamin, a novel mechanochemical enzyme that mediates interactions between microtubules. Cell 59: 421–432, 1989.
 460. Shuster, C. B., and I. M. Herman. Indirect association of ezrin with F‐actin: isoform specificity and calcium sensitivity. J. Cell Biol. 128: 837–848, 1995.
 461. Siegel, D. L., and D. Branton. Partial purification and characterization of an actin‐bundling protein, band 4.9, from human erythrocytes. J. Cell Biol. 100: 775–785, 1985.
 462. Sikorski, A. F., W. Swat, M. Brzezinska, Z. Wroblewski, and B. Bisikirska. A protein cross‐reacting with anti‐spectrin antibodies is present in higher plant cells. Z Naturforsch [C] 48: 580–583, 1993.
 463. Sikorski, A. F., G. Terlecki, I. S. Zagon, and S. R. Goodman. Synapsin I‐mediated interaction of brain spectrin with synaptic vesicles. J. Cell Biol. 114: 313–318, 1991.
 464. Sinard, J. H., G. W. Stewart, A. C. Argent, D. M. Gilligan, and J. S. Morrow. A novel isoform of beta adducin utilizes an alternatively spliced exon near the C‐terminus. Mol. Biol. Cell 6: 269a, 1995.
 465. Sinard, J. H., G. W. Stewart, A. C. Argent, and J. S. Morrow. Stomatin binding to adducin: a novel link between transmembrane ion transport and the cytoskeleton. Mol. Biol. Cell 5 (suppl): 421a, 1994.
 466. Sinard, J. H., G. W. Stewart, A. C. Argent, P. R. Stabach, and J. S. Morrow. Characterization of a novel alternative transcript of beta adducin. Genbank #U43959, 1995.
 467. Skolnik, E. Y., A. Batzer, N. Li, C.‐H. Lee, E. Lowenstein, M. Mohammadi, B. Margolis, and J. Schlessinger. The function of GRB2 in linking the insulin receptor to ras signaling pathways. Science 260: 1953–1955, 1993.
 468. Smith, P. R., A. L. Bradford, E. H. Joe, K. J. Angelides, D. J. Benos, and G. Saccomani. Gastric parietal cell H(+)‐K(+)‐ATPase microsomes are associated with isoforms of ankyrin and spectrin. Am. J. Physiol. 264: C63–70, 1993.
 469. Smith, P. R., G. Saccomani, E. H. Joe, K. J. Angelides, and D. J. Benos. Amiloride‐sensitive sodium channel is linked to the cytoskeleton in renal epithelial cells. Proc. Natl. Acad. Sci. U.S.A. 88: 6971–6975, 1991.
 470. Sontag, J.‐M., E. M. Fykse, Y. Ushkaryov, J.‐P. Liu, P. J. Robinson, and T. C. Sudhof. Differential expression and regulation of multiple dynamins. J. Biol. Chem. 269: 4547–4554, 1994.
 471. Speicher, D. W., T. M. DeSilva, K. D. Speicher, J. A. Ursitti, P. Hembach, and L. Weglarz. Location of the human red cell spectrin tetramer binding site and detection of a related “closed” hairpin loop dimer using proteolytic footprinting. J. Biol. Chem. 268: 4227–4235, 1993.
 472. Speicher, D. W., and V. T. Marchesi. Erythrocyte spectrin is comprised of many homologous triple helical segments. Nature 311: 177–180, 1984.
 473. Speicher, D. W., J. S. Morrow, W. J. Knowles, and V. T. Marchesi. Identification of proteolytically resistant domains of human erythrocyte spectrin. Proc. Natl. Acad. Sci. U.S.A. 77: 5673–5677, 1980.
 474. Speicher, D. W., J. S. Morrow, W. J. Knowles, and V. T. Marchesi. A structural model of human erythrocyte spectrin: alignment of chemical and functional domains. J. Biol. Chem. 257: 9093–9101, 1982.
 475. Speicher, D. W., L. Weglarz, and T. M. DeSilva. Properties of human red cell spectrin heterodimer (side‐to‐side) assembly and identification of an essential nucleation site. J. Biol. Chem. 267: 14775–14782, 1992.
 476. Srinivasan, Y., M. Lewallen, and K. J. Angelides. Mapping the binding site on ankyrin for the voltage‐dependent sodium channel from brain. J. Biol. Chem. 267: 7483–7489, 1992.
 477. Srinivasan, Y. L., E. J. Davis, V. Bennett, and K. Angelides. Ankyrin and spectrin associate with voltage‐dependent Na channels in brain. Nature 333: 177–180, 1988.
 478. Stabach, P. R., C. D. Cianci, S. B. Glantz, Z. Zhang, and J. S. Morrow. Site‐directed mutagenesis of αll spectrin (fodrin) reveals determinants of its μ‐calpain susceptibility. Biochemistry. In press, 1996.
 479. Stabach, P. R., S. P. Kennedy, and J. S. Morrow. Polarized assembly of the spectrin skeleton is ankyrin‐independent in MDCK cells. Mol. Biol. Cell 4: 58a, 1993.
 480. Stappert, J., and R. Kemler. A short core region of E‐cadherin is essential for catenin binding and is highly phosphorylated. Cell Adhes. Commun. 2: 319–327, 1994.
 481. Steiner, J. P., and V. Bennett. Ankyrin‐independent membrane protein‐binding sites for brain and erythrocyte spectrin. J. Biol. Chem. 263: 14417–14425, 1988.
 482. Steiner, J. P., H. T. J. Walke, and V. Bennett. Calcium/calmodulin inhibits direct binding of spectrin to synaptosomal membranes. J. Biol. Chem. 264: 2783–2791, 1989.
 483. Stewart, G. W., and A. C. Argent. The integral band 7 membrane protein of the human erythrocyte membrane. Biochem. Soc. Trans. 20: 785–790, 1992.
 484. Stewart, G. W., A. C. Argent, and B. C. Dash. Stomatin: a putative cation transport regulator in the red cell membrane. Biochim. Biophys. Acta 1225: 15–25, 1993.
 485. Stewart, G. W., B. E. Hepworth‐Jones, J. N. Keen, B.C.J. Dash, A. C. Argent, and C. M. Casimir. Isolation of cDNA coding for a ubiquitóus membrane protein deficient in high Na, low K stomatocytic erythrocytes. Blood 79: 1593–1601, 1992.
 486. Stokke, B. T., A. Mikkelsen, and A. Elgsaeter. The human erythrocyte membrane skeleton may be an ionic gel. I. Membrane mechanochemical properties. Eur. Biophys. J. 13: 203–218, 1986.
 487. Stokke, B. T., A. Mikkelsen, and A. Elgsaeter. The human erythrocyte membrane skeleton may be an ionic gel. II. Numerical analyses of cell shapes and shape transformations. Eur. Biophys. J. 13: 219–233, 1986.
 488. Stokke, B. T., A. Mikkelsen, and A. Elgsaeter. The human erythrocyte membrane skeleton may be an ionic gel. III. Micropipette aspiration of unswollen erythrocytes. J. Theor. Biol. 123: 205–211, 1986.
 489. Straub, V., R. E. Bittner, J. J. Leger, and T. Voit. Direct visualization of the dystrophin network on skeletal muscle fiber membrane. J. Cell Biol. 119: 1183–1191, 1992.
 490. Su, L. K., B. Vogelstein, and K. W. Kinzler. Association of the APC tumor suppressor protein with catenins. Science 262: 1734–1737, 1993.
 491. Sung, L. A., V. M. Fowler, K. Lambert, M. A. Sussman, D. Karr, and S. Chien. Molecular cloning and characterization of human fetal liver tropomodulin. A tropomyosin‐binding protein. J. Biol. Chem. 267: 2616–2621, 1992.
 492. Sussman, M. A., and V. M. Fowler. Tropomodulin binding to tropomyosins. Isoform‐specific differences in affinity and stoichiometry. Eur. J. Biochem. 205: 355–362, 1992.
 493. Svetina, S., A. Iglic, V. Kralj‐Iglic, and B. Zeks. Cytoskeleton and red cell shape. Cell Mol. Biol. Lett. 1: 67–75, 1996.
 494. Svetina, S., V. Kralj‐Iglic, and B. Zeks. Cell shape and lateral distribution of mobile membrane constituents. In: Proceedings X. School on Biophysics of Membrane Transport, Part II, edited by J. Kuczera and S. Przestalski, Wroclaw: Agricultural University of Wroclaw, 1990, p. 139–155.
 495. Svetina, S., S. Vrhoec, M. Gros, and B. Zeks. Red blood cell membrane as a possible model system for studying mechanisms of chlorpromazine action. Acta Pharmacologica 42: 25–35, 1992.
 496. Svetina, S., and B. Zeks. Bilayer couple as a possible mechanism of biological shape formation. Biomed. Biochim. Acta 44: 979–986, 1985.
 497. Svetina, S., and B. Zeks. Elastic properties of layered membranes and their role in transformations of cellular shapes. In: Biomechanics of Active Movement and Division of Cells, edited by N. Akkas, Berlin: Springer‐Verlag, 1994, p. 479–486.
 498. Svoboda, K., C. F. Schmidt, D. Branton, and S. M. Block. Conformation and elasticity of the isolated red blood cell membrane skeleton. Biophys. J. 63: 784–793, 1992.
 499. Takaoka, Y., H. Ideguchi, M. Matsuda, N. Sakamoto, T. Takeuchi, and Y. Fukumaki. A novel mutation in the erythrocyte protein 4.2 gene of Japanese patients with hereditary spherocytosis (protein 4.2 Fukuoka). Br. J. Haematol. 88: 527–533, 1994.
 500. Takeichi, M. Cadherin cell adhesion receptors as a morphogenetic regulator. Science 251: 1451–1455, 1991.
 501. Takeichi, M. Morphogenetic roles of classic cadherins. Curr. Opin. Cell Biol. 7: 619–627, 1995.
 502. Tanaka, T., K. Kadowaki, E. Lazarides, and K. Sobue. Ca2+‐dependent regulation of the spectrin/actin interaction by calmodulin and protein 4.1. J. Biol. Chem. 266: 1134–1140, 1991.
 503. Tang, T. K., T. L. Leto, V. T. Marchesi, and E. J. Benz. Expression of specific isoforms of protein 4.1 in erythroid and non‐erythroid tissues. Adv. Exp. Med. Biol. 241: 81–95, 1988.
 504. Tang, T. K., Z. Qin, T. Leto, V. T. Marchesi, and E. J. Benz, Jr.. Heterogeneity of mRNA and protein products arising from the protein 4.1 gene in erythroid and nonerythroid tissues. J. Cell Biol. 110: 617–624, 1990.
 505. Tanner, M. J. The major integral proteins of the human red cell. Baillieres Clin. Haematol. 6: 333–356, 1993.
 506. Tao, Y.‐S., R. Edwards, J. Bryan, and P. D. McCrea. The actin bundling protein fascin associates with beta‐catenin. Mol. Biol. Cell 6: 300a, 1995.
 507. Taylor, S. A., R. G. Snell, A. Buckler, C. Ambrose, M. Duyao, D. Church, C. S. Lin, M. Altherr, G. P. Bates, N. Groot, et al. Cloning of the alpha‐adducin gene from the Huntington's disease candidate region of chromosome 4 by exon amplification. Nat. Genet. 2: 223–227, 1992.
 508. Tillack, T. W., S. L. Marchesi, V. T. Marchesi, and E. Steers, Jr.. A comparative study of spectrin: a protein isolated from red blood cell membranes. Biochim. Biophys. Acta 200: 125–131, 1970.
 509. Tilley, L., R. A. McPherson, G. L. Jones, and W. H. Sawyer. Structural organisation of band 3 in Melanesian ovalocytes. Biochim. Biophys. Acta 1181: 83–89, 1993.
 510. Torres, M., C. Yost, D. Kimelman, and R. T. Moon. Phosphorylation of beta‐catenin by glycogen synthase kinase‐3 increases its interaction with cadherin and decreases its axis inducing activity in xenopus. Mol. Biol. Cell 6: 117a, 1995.
 511. Touhara, K., J. Inglese, J. A. Pitcher, G. Shaw, and R. J. Lefkowitz. Binding of G protein beta gamma‐subunits to pleckstrin homology domains. J. Biol. Chem. 269 ‐: 10217–10220, 1994.
 512. Trave, G., A. Pastore, M. Hyvonen, and M. Saraste. The C‐terminal domain of alpha‐spectrin is structurally related to calmodulin. Eur. J. Biochem. 227: 35–42, 1995.
 513. Tripodi, G., A. Piscone, G. Borsani, S. Tisminetzky, S. Salardi, A. Sidoli, P. James, S. Pongor, G. Bianchi, and F. E. Baralle. Molecular cloning of an adducin‐like protein: evidence of a polymorphism in the normotensive and hypertensive rats of the milan strain. Biochem. Biophys. Res. Commun. 177: 939–947, 1991.
 514. Troyanovsky, S. M., R. B. Troyanovsky, L. G. Eshkind, V. A. Krutovskikh, R. E. Leube, and W. W. Franke. Identification of the plakoglobin‐binding domain in desmoglein and its role in plaque assembly and intermediate filament anchorage. J. Cell. Biol. 127: 151–160, 1994.
 515. Tse, W. T., M. C. Lecomte, F. F. Costa, M. Garbarz, C. Feo, P. Boivin, D. Dhermy, and B. G. Forget. Point mutation in the beta‐spectrin gene associated with alpha 1/74 hereditary elliptocytosis. Implications for the mechanism of spectrin dimer self‐association. J. Clin. Invest. 86: 909–916, 1990.
 516. Ungewickell, E., P. M. Bennett, R. Calvert, V. Ohanian, and W. B. Gratzer. In vitro formation of a complex between cytoskeletal proteins of the human erythrocyte. Nature 280: 811–814, 1979.
 517. Ursitti, J. A., and V. M. Fowler. Immunolocalization of tropomodulin, tropomyosin and actin in spread human erythrocyte skeletons. J. Cell Sci. 107: 1633–1639, 1994.
 518. Ursitti, J. A., D. W. Pumplin, J. B. Wade, and R. J. Bloch. Ultrastructure of the human erythrocyte cytoskeleton and its attachment to the membrane. Cell Motil. Cytoskeleton 19: 227–243, 1991.
 519. Ursitti, J. A., and J. B. Wade. Ultrastructure and immunocyto‐chemistry of the isolated human erythrocyte membrane skeleton. Cell Motil. Cytoskeleton 25: 30–42, 1993.
 520. Viel, A., and D. Branton. Interchain binding at the tail end of the Drosophila spectrin molecule. Proc. Natl. Acad. Sci. U.S.A. 91: 10839–10843, 1994.
 521. Viel, A., M. Gee, and D. Branton. Molecular analysis of interchain binding at the tail end of Drosophila spectrin. Mol. Biol. Cell 6: 270a (abstract), 1995.
 522. Vybiral, T., J. C. Winkelmann, R. Roberts, E. H. Joe, D. L. Casey, J. K. Williams, and H. F. Epstein. Human cardiac and skeletal muscle spectrins: differential expression and localization. Cell Mot. Cytoskeleton 21: 293–304, 1992.
 523. Wallis, C., E. Wenegieme, and J. Babitch. Characterization of calcium binding to brain spectrin. J. Biol. Chem. 267: 4333–4337, 1992.
 524. Wang, C. C., R. Moriyama, C. R. Lombardo, and P. S. Low. Partial characterization of the cytoplasmic domain of human kidney band 3. J. Biol. Chem. 270: 17892–17897, 1995.
 525. Wang, D., W. C. Mentzer, T. Cameron, and R. M. Johnson. Purification of band 7.2b, a 31‐kDa integral phosphoprotein absent in hereditary stomatocytosis. J. Biol. Chem. 266: 17826–17831, 1991.
 526. Wang, D. S., and G. Shaw. The association of the C‐terminal region of beta I sigma II spectrin to brain membranes is mediated by a PH domain, does not require membrane proteins, and coincides with a inositol‐1,4,5 triphosphate binding site. Biochem. Biophys. Res. Commun. 217: 608–615, 1995.
 527. Wang, K., and F. M. Richards. An approach to nearest neighbor analysis of membrane proteins. J. Biol. Chem. 249: 8005–8018, 1974.
 528. Waseem, A., and H. C. Palfrey. Erythrocyte adducin. Comparison of the alpha‐ and beta‐subunits and multiple‐site phosphorylation by protein kinase C and cAMP‐dependent protein kinase. Eur. J. Biochem. 178: 563–573, 1988.
 529. Wasenius, V. M., O. Narvanen, V. P. Lehto, and M. Saraste. Alpha‐actinin and spectrin have common structural domains. FEBS Lett. 221: 73–76, 1987.
 530. Wasenius, V. M., M. Saraste, P. Salven, M. Eramaa, L. Holm, and V. P. Lehto. Primary structure of the brain alpha‐spectrin. J. Cell Biol. 108: 79–93, 1989.
 531. Way, M., B. Pope, and A. Weeds. Molecular biology of actin binding proteins: evidence for a common structural domain in the F‐actin binding sites of gelsolin and alpha‐actinin. J. Cell Sci. Suppl. 14: 91–94, 1991.
 532. Weaver, D. C., G. R. Pasternack, and V. T. Marchesi. The structural basis of ankyrin function II. Identification of two functional domains. J. Biol. Chem. 259: 6170–6175, 1984.
 533. Weber, A., C. R. Pennise, G. G. Babcock, and V. M. Fowler. Tropomodulin caps the pointed ends of actin filaments [see comments]. J. Cell Biol. 127: 1627–1635, 1994.
 534. Weed, S. A., P. R. Stabach, and J. S. Morrow. Developmental expression of βç2 spectrin in developing rat brain. in preparation, 1996.
 535. Weed, S. A., P. R. Stabach, and J. S. Morrow. The MAD2 (pleckstrin homology) domain of βIç2 spectrin is not required for cortical membrane assembly or myotube formation in skeletal muscle. in preparation, 1996.
 536. Weinstein, R. S., H. D. Tazelaar, and J. M. Loew. Red cell comets: ultrastructure of axial elongation of the membrane skeleton. Blood Cells 11: 343–366, 1986.
 537. Welsh, C. F., D. Zhu, and L. Y. Bourguignon. Interaction of CD44 variant isoforms with hyaluronic acid and the cytoskeleton in human prostate cancer cells. J. Cell Physiol. 164: 605–612, 1995.
 538. Westberg, J. A., B. Entler, R. Prohaska, and J. P. Schroder. The gene coding for erythrocyte protein band 7.2b (EPB72) is located in band q34.1 of human chromosome 9. Cytogenet. Cell Genet. 63: 241–243, 1993.
 539. White, R. A., L. L. Peters, L. R. Adkison, C. Korsgren, C. M. Cohen, and S. E. Lux. The murine pallid mutation is a platelet storage pool disease associated with the protein 4.2 (pallidin) gene. Nat Genet. 2: 80–3, 1992.
 540. Willott, E., M. S. Balda, A. S. Fanning, B. Jameson, C. Van Itallie, and J. M. Anderson. The tight junction protein ZO‐1 is homologous to the Drosophila discs‐large tumor suppressor protein of septate junctions. Proc. Natl. Acad. Sci. U.S.A. 90: 7845–7838, 1993.
 541. Winkelmann, J. C., J. G. Chang, W. T. Tse, A. L. Scarpa, V. T. Marchesi, and B. G. Forget. Full‐length sequence of the cDNA for human erythroid beta‐spectrin. J. Biol. Chem. 265: 11827–11832, 1990.
 542. Winkelmann, J. C., F. F. Costa, B. L. Linzie, and B. G. Forget. Beta spectrin in human skeletal muscle. Tissue‐specific differential processing of 3' beta spectrin pre‐mRNA generates a beta spectrin isoform with a unique carboxyl terminus. J. Biol. Chem. 264: 20449–54, 1990.
 543. Winkelmann, J. C., and B. G. Forget. Erythroid and nonerythroid spectrins. Blood 81: 3173–3185, 1993.
 544. Winkelmann, J. C., T. L. Leto, P. C. Watkins, R. Eddy, T. B. Shows, A. J. Linnenbach, K. E. Sahr, N. Kathuria, V. T. Marchesi, and B. G. Forget. Molecular cloning of the cDNA for human erythrocyte beta‐spectrin. Blood 72: 328–334, 1988.
 545. Winograd, E., D. Hume, and D. Branton. Phasing the conformational unit of spectrin. Proc. Natl. Acad. Sci. U.S.A. 88: 10788–10791, 1991.
 546. Wolfe, L. C., K. M. John, J. C. Falcone, A. M. Byrne, and S. E. Lux. A genetic defect in the binding of protein 4.1 to spectrin in a kindred with hereditary spherocytosis. N Engl. J. Med. 307: 1367–1374, 1982.
 547. Woods, D. F., and P. J. Bryant. The discs‐large tumor suppressor gene of Drosophila encodes a guanylate kinase homolog localized at septate junctions. Cell 66: 451–464, 1991.
 548. Woods, D. F., and P. J. Bryant. ZO‐1, DigA and PSD‐95/SAP‐90; homologous proteins in tight, septate and synaptic cell junctions. Mech. Devel. 44: 85–89, 1993.
 549. Yan, Y., E. Winograd, A. Viel, T. Cronin, S. C. Harrison, and D. Branton. Crystal structure of the repetitive segments of spectrin. Science 262: 2027–2030, 1993.
 550. Yawata, Y., M. Inaba, A. Yawata, A. Kanzaki, K. Ono, M. Takeuchi, K. Sato, Y. Maede, and Y. Takakuwa. Complete band 3 deficiency in cattle: a model for hereditary sherocytosis with striking instability of cytoskeletal network with marked exo‐ and endocytosis. Blood 86 (suppl): 468a (abstract), 1995.
 551. Yonemura, S., M. Itoh, A. Nagafuchi, and S. Tsukita. Cell‐to‐cell adherens junction formation and actin filament organization: similarities and differences between non‐polarized fibroblasts and polarized epithelial cells. J. Cell Sci. 108: 127–142, 1995.
 552. Yurchenco, P. D., D. W. Speicher, J. S. Morrow, W. J. Knowles, and V. T. Marchesi. Monoclonal antibodies as probes of domain structure of the spectrin alpha subunit. J. Biol. Chem. 257: 9102–9107, 1982.
 553. Zarkowsky, H. S., F. A. Oski, R. Shaafi, S. B. Shohet, and D. G. Nathan. Congenital hemolytic anemia with high sodium, low potassium red cells. I Studies of membrane permeability. N. Engl. J. Med. 278: 573–581, 1968.
 554. Zimmer, W. E., Y. Ma, I. S. Zagon, and S. R. Goodman. Developmental expression of brain β‐specrin isoform messenger RNAs. Brain Res. 594: 75–83, 1992.
 555. Zimmer, W. E., Y. P. Ma, and S. R. Goodman. Identification of a mouse brain beta‐spectrin cDNA and distribution of its mRNA in adult tissues. Brain Res. Bull. 27: 187–193, 1991.
 556. Zimmerman, U. J., D. W. Speicher, and A. B. Fisher. Secretagogue‐induced proteolysis of lung spectrin in alveolar epithelial type II cells. Biochim. Biophys. Acta 1137: 127–134, 1992.
 557. Ziparo, E., B. M. Zani, A. Filippini, M. Stefanini, and V. T. Marchesi. Proteins of the membrane skeleton in rat Sertoli cells. J. Cell Sci. 86: 145–154, 1986.
 558. Southgate, C. D., A. H. Chishti, B. Mitchell, S. J. Yi, and J. Palek. Targeted disruption of the murine erythroid band 3 gene results in spherocytosis and severe haemolytic anaemia despite a normal membrane skeleton. Nature Genet. 14: 227–230, 1996.

Contact Editor

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

* Required Field

How to Cite

Jon S. Morrow, David L. Rimm, Scott P. Kennedy, Carol D. Cianci, John H. Sinard, Scott A. Weed. Of Membrane Stability and Mosaics: The Spectrin Cytoskeleton. Compr Physiol 2011, Supplement 31: Handbook of Physiology, Cell Physiology: 485-540. First published in print 1997. doi: 10.1002/cphy.cp140111