Dynamic reorganization of the actin cytoskeleton is fundamental to a number of cell biological events. A variety of actin-regulatory proteins modulate polymerization and depolymerization of actin and contribute to actin cytoskeletal reorganization. Cyclase-associated protein (CAP) is a conserved actin-monomer-binding protein that has been studied for over 20 years. Early studies have shown that CAP sequesters actin monomers; recent studies, however, have revealed more active roles of CAP in actin filament dynamics. CAP enhances the recharging of actin monomers with ATP antagonistically to ADF/cofilin, and also promotes the severing of actin filaments in cooperation with ADF/cofilin. Self-oligomerization and binding to other proteins regulate activities and localization of CAP. CAP has crucial roles in cell signaling, development, vesicle trafficking, cell migration and muscle sarcomere assembly. This Commentary discusses the recent advances in our understanding of the functions of CAP and its implications as an important regulator of actin cytoskeletal dynamics, which are involved in various cellular activities.

The coordinated regulation of the assembly and disassembly of actin filaments enables spatially and temporally controlled dynamics of actin cytoskeletal structures within cells. The combined functions of a minimal set of actin-regulatory proteins for nucleation, depolymerization and capping can explain the basic mechanism of persistent actin filament turnover, for example, at the leading edge of motile cells or in the comet tails of bacterial pathogens (Bugyi and Carlier, 2010; Pollard and Cooper, 2009). However, the actin-cytoskeletal structures in different types of cell exhibit different patterns of dynamics that involve a number of additional actin-binding proteins. On the one hand, CAP has been described as an actin-monomer-binding protein that can sequester actin monomers and prevent them from polymerization in vitro (Hubberstey and Mottillo, 2002). On the other hand, genetic and cell biological studies of CAP in several organisms have suggested that CAP has important roles in the dynamic actin reorganization that cannot be simply explained by its actin-monomer-sequestering function (Balcer et al., 2003; Bertling et al., 2004; Freeman and Field, 2000). Recent studies have revealed new biochemical functions of CAP. It promotes actin filament dynamics (Box 1), closely cooperates with ADF/cofilin (Box 2) in vitro and in vivo (Moriyama and Yahara, 2002), and self-oligomerization of CAP enhances its activities (Quintero-Monzon et al., 2009). Furthermore, the conservation of CAPs among eukaryotes suggests that CAP is a fundamentally important actin regulator. This article covers biochemical and cell biological functions of CAPs with an emphasis on their roles in regulating the actin cytoskeleton.

Domain organization of CAP

Depending on the species, CAP consists of 450–550 amino acids, and has several distinct motifs and domains (Fig. 1A), which is discussed in this section. Most species have only a single CAP isoform, although vertebrates have two CAP isoforms, CAP1 and CAP2. Parasitic protozoa only have the C-terminal sequence of CAP, the so-called C-CAPs (Hliscs et al., 2010) (Fig. 1A). Isoforms and orthologs of CAP are discussed in the next section.

Fig. 1.

Structure of CAP. (A) Domain organization of CAP and C-CAP (CAP in protozoans). Representative interaction partners and actin-regulatory functions are shown below functional domains. (B,C) Ribbon diagram of the structures of the helical folded domain (HFD) (Protein Data Bank accession number 1S0P) (B) and β-sheet/β-helix (CARP) domain (Protein Data Bank accession number 1K8F) (C). Structures of monomers (top) and dimers (bottom) are shown. Locations of N- and C-termini are indicated. The molecular graphics were produced with PyMOL (Schrödinger, LLC).

Fig. 1.

Structure of CAP. (A) Domain organization of CAP and C-CAP (CAP in protozoans). Representative interaction partners and actin-regulatory functions are shown below functional domains. (B,C) Ribbon diagram of the structures of the helical folded domain (HFD) (Protein Data Bank accession number 1S0P) (B) and β-sheet/β-helix (CARP) domain (Protein Data Bank accession number 1K8F) (C). Structures of monomers (top) and dimers (bottom) are shown. Locations of N- and C-termini are indicated. The molecular graphics were produced with PyMOL (Schrödinger, LLC).

The most N-terminal amino acids (∼40 residues) of CAP are predicted to form a coiled-coil (Fig. 1A); this region of yeast Srv2/CAP is sufficient for binding to adenylyl cyclase, an enzyme that produces cAMP (Nishida et al., 1998) and is required for self-oligomerization of Srv2/CAP (Quintero-Monzon et al., 2009). However, this N-terminal region is unstructured in solution (Mavoungou et al., 2004) and coiled-coil formation has not been structurally demonstrated thus far. Therefore, a coiled-coil might form upon binding to adenylyl cyclase or self-oligomerization of CAP. Most of the N-terminal region is composed of a stable bundle of six antiparallel α-helices (Fig. 1B) that are termed helical folded domain (HFD). Structural analyses show that the HFD forms a laterally associated anti-parallel dimer (Ksiazek et al., 2003; Yusof et al., 2005) (Fig. 1B, bottom). The HFD binds to the ADF/cofilin-G-actin complex (Moriyama and Yahara, 2002; Quintero-Monzon et al., 2009) and also interacts with ADF/cofilin-bound actin filaments to promote their severing (Chaudhry et al., 2013) (Box 2).

In the central region of CAP, there are two proline-rich regions (P1 and P2) (Fig. 1A). P1 is highly conserved in CAPs from different species – with a stretch of ∼10 proline residues, whereas P2 is less conserved – with an only scattered appearance of proline residues. Several interaction partners for CAP selectively bind to either P1 or P2. For instance, profilin, an actin-monomer-binding protein, binds to P1 of yeast Srv2/CAP (Bertling et al., 2007) and of mouse CAP1 (Makkonen et al., 2013). The Src homology 3 (SH3) domain of the tyrosine kinase Abl binds to the P1 of human CAP1 (Freeman et al., 1996), that of the yeast actin-binding protein 1 (ABP1) binds to P2 of yeast Srv2/CAP (Freeman et al., 1996; Lila and Drubin, 1997). Also located in the central region of CAP is a Wiscott Aldrich Syndrome protein homology 2 (WH2) domain (Fig. 1A) (Paunola et al., 2002). WH2 of yeast Srv2/CAP binds to G-actin without exhibiting a strong preference for actin-bound nucleotides and has a crucial role in the nucleotide exchange of G-actin, as described below (Chaudhry et al., 2010) (see Box 1 for the significance of nucleotides in actin dynamics).

The C-terminal half of CAP is mostly composed of β-strands, which are arranged into six coils of right-handed β-helixes (Dodatko et al., 2004) (Fig. 1C). The most C-terminal ∼35 amino acids contain a dimerization motif, in which β-strands participate in a β-sheet of a second CAP molecule, resulting in a strand-exchanged dimer (Fig. 1C, bottom). In addition, the C-terminal half of CAP contains a sequence motif that is present in a number of proteins, including tubulin-binding cofactor C and X-linked retinitis pigmentosa 2 protein (RP2) (Dodatko et al., 2004); the homologous domains are currently termed CAPs and RP2 (CARP) domains in the Simple Modular Architecture Research Tool (SMART) database (Schultz et al., 1998). The CARP domain of CAP binds to G-actin independently of the WH2 domain (Chaudhry et al., 2010; Makkonen et al., 2013; Mattila et al., 2004; Nomura and Ono, 2013).

CAP isoforms, orthologs and nomenclature

CAPs have been identified in almost all representative eukaryotic species and are most probably present in all eukaryotes (Fig. 2; supplementary material Table S1). The domain structure of CAP is conserved among species. However, C-CAPs in protozoa only have the CARP domain and the dimerization motif (Hliscs et al., 2010) (Fig. 1A). Leishmania and Trypanosoma C-CAPs have a proline-rich sequence at the N-terminus, but other C-CAPs do not (Hliscs et al., 2010). Leishmania major has another long form of a CAP-related protein that contains the CARP domain (Fig. 2, see the atypical CAP in the phylogenetic tree) and contains an N-terminal region that is not homologous to the HFD domain of CAPs (Hliscs et al., 2010). Fungi, protists and many invertebrates only have one CAP gene (Fig. 2). However, multiple CAP isoforms are present in vertebrates, nematodes and green plants (Fig. 2). Vertebrates have two CAP isoforms, CAP1 and CAP2 (Swiston et al., 1995; Yu et al., 1994), which are ∼60% identical in their amino acid sequence and exhibit different expression patterns. In mammals, CAP1 is widely expressed – except in skeletal muscle, whereas CAP2 is predominantly expressed in brain, heart, skeletal muscle and testis (Bertling et al., 2004; Swiston et al., 1995). In particular, CAP2 has been shown to be a crucial regulator of sarcomere assembly in striated muscle (heart and skeletal muscle), as discussed below. However, it is currently unknown whether CAP1 and CAP2 have different biochemical properties. Similarly, the nematode C. elegans has both a muscle-specific CAP (CAS-1) (Nomura et al., 2012) and a nonmuscle CAP (CAS-2) (Nomura and Ono, 2013). In addition, several CAP isoforms have been identified in several green plants (Fig. 2; supplementary material Table S1), but their functional differences have not been determined. Thus, other than the tissue-specific expression of CAP isoforms, their functional differences are not well understood. Multiple isoforms of other actin regulators, including ADF/cofilin (Ono, 2007; Poukkula et al., 2011) and profilin (Jockusch et al., 2007; Kovar et al., 2000; Polet et al., 2006), are also expressed in these multicellular organisms, suggesting that specific isoforms of these actin regulators cooperate to modulate actin filament dynamics in a tissue-specific manner.

Fig. 2.

Phylogenetic tree of CAPs in eukaryotes. Amino acid sequences of CAP-related proteins from representative organisms were analyzed by Clustal W (Thompson et al., 1994) and classified as shown on the left. See supplementary material Table S1 for database accession numbers sequences.

Fig. 2.

Phylogenetic tree of CAPs in eukaryotes. Amino acid sequences of CAP-related proteins from representative organisms were analyzed by Clustal W (Thompson et al., 1994) and classified as shown on the left. See supplementary material Table S1 for database accession numbers sequences.

Cyclase-associated protein is the original name for CAP. However, several alternative names have also been used. CAP was originally identified in budding yeast as an adenylyl-cyclase-associated protein (Field et al., 1990) and also as a suppressor of hyperactive RAS2(V19), thus explaining the yeast name Srv2 (Fedor-Chaiken et al., 1990). CAP and Srv2 are identical, and often designated as Srv2/CAP. ASP-56 has been isolated from pig platelets as an actin-monomer-binding protein, and partial peptide sequences have shown that ASP-56 is porcine CAP1 (Gieselmann and Mann, 1992). Rat CAP1 was cloned as a mammalian CAP homolog 1 and originally named MCH1 (Zelicof et al., 1993). A Drosophila CAP gene was identified following the screening of mutant flies and is allelic to capulet (Baum et al., 2000; Wills et al., 2002) and act up (Benlali et al., 2000). C. elegans CAP genes were namend cas-1 and cas-2 (Nomura et al., 2012), because ‘cap’ had already been used for genes that encode subunits of actin capping proteins (Waddle et al., 1993). Apart from these exceptions, cyclase-associated protein or CAP is used as a common name for these proteins in most of the literature.

Conserved functions of CAPs as regulators of the actin cytoskeleton

Although CAP was originally discovered as a component of the Ras–cAMP signaling pathway, studies of CAPs in various species suggest that this signaling function is specific to only a limited number of organisms; but, instead, their cytoskeletal function appears to be conserved among eukaryotes (Hubberstey and Mottillo, 2002). Two unrelated phenotypes can be detected in Srv2/CAP-deficient yeast cells: defective Ras–cAMP signaling and morphological abnormalities (Fedor-Chaiken et al., 1990; Field et al., 1990). These could be independently suppressed by the expression of an N-terminal half and a C-terminal half of Srv2/CAP, respectively (Gerst et al., 1991). The morphological abnormalities in Srv2/CAP-deficient cells are associated with a lack of polarized actin distribution, the loss of normal actin cables and the formation of abnormal actin bundles (Vojtek et al., 1991). Furthermore, the expression of a heterologous CAP from other species, e.g. fission yeast, in Srv2/CAP-deficient budding yeast can suppress the actin-related phenotype but not the Ras-dependent phenotype, suggesting that the actin-regulatory function is evolutionarily conserved among CAPs (Kawamukai et al., 1992). Similar genetic studies have demonstrated a conserved actin-regulatory function of CAPs from Lentinus edodes (shiitake mushroom) (Zhou et al., 1998), rats (Zelicof et al., 1993) and humans (Matviw et al., 1992). Thus, these genetic assays have provided solid evidence that CAPs are, indeed, conserved regulators of the actin cytoskeleton.

Effects of CAP on G-actin – monomer sequestration and enhancement of nucleotide exchange

Early biochemical studies on CAPs suggested that CAPs are actin-monomer sequestering proteins that prevent actin nucleation and elongation from filament ends (Freeman et al., 1995). However, recent studies have revealed that CAPs fulfill more active roles in promoting the dynamics of actin filaments. Therefore, the interpretations of some early genetic and cell biology studies may need to be revised on the basis of these recently identified functions for CAPs in actin dynamics.

CAP binds to G-actin at a molar ratio of 1∶1 and inhibits the spontaneous actin polymerization of G-actin. This activity has been demonstrated for CAPs from budding yeast (Freeman et al., 1995), Dictyostelium (Gottwald et al., 1996), Arabidopsis (Chaudhry et al., 2007), C. elegans (CAS-1 and CAS-2) (Nomura et al., 2012; Nomura and Ono, 2013), pig platelets (ASP-56/CAP1) (Gieselmann and Mann, 1992), humans (CAP2) (Peche et al., 2013) and Cryptosporidium (C-CAP) (Hliscs et al., 2010). The effect of CAP on actin elongation is somewhat controversial; Freeman et al. (Freeman et al., 1995) reported that yeast Srv2/CAP inhibits the incorporation of G-actin to either pointed or barbed ends of actin filaments, whereas Mattila et al. (Mattila et al., 2004) reported that yeast Srv2/CAP specifically inhibits the incorporation of G-actin to the barbed end. This discrepancy might be owing to different ratios of Srv2/CAP to G-actin used in these studies, because it was shown that human CAP1 accelerates the addition of G-actin to barbed ends when it is present in sub-stoichiometric amounts compared with G-actin, but inhibits monomer addition when CAP1 and G-actin are used at stoichiometric amounts (Moriyama and Yahara, 2002). At the steady state, CAP sequesters G-actin when it is present in a concentration that is equimolar to that of G-actin. As the cellular concentration of CAPs has not been determined, it is unclear whether CAP is present in cells at a concentration that is sufficient to contribute to sequestration of G-actin, which is generally maintained at a high level of 30∼50% of total actin in non-muscle cells (Heacock and Bamburg, 1983). CAPs localize to actin patches in yeast (Lila and Drubin, 1997) and to the leading edges of motile Dictyostelium (Gottwald et al., 1996) and mammalian cells (Freeman and Field, 2000; Zelicof et al., 1996). Therefore, the local concentration of CAP could be sufficiently high to sequester actin monomers at these locations. However, it remains unclear whether the actin-monomer sequestering function of CAP is indeed important in vivo.

CAP also enhances the rate of exchange of G-actin-bound nucleotides (Fig. 3), which might be its most important actin-regulatory effect. During the ‘actin treadmilling’ cycle, the constant recharging of G-actin with ATP and ATP hydrolysis by F-actin drive the biased depolymerization of ADP–G-actin from the pointed end of the actin filament and the polymerization of ATP–G-actin from the barbed end (Bugyi and Carlier, 2010; De La Cruz and Pollard, 2001; Kinosian et al., 1993; Pollard, 1986; Pollard et al., 1992) (Box 1). However, several G-actin-binding proteins, including thymosin β4 (Goldschmidt-Clermont et al., 1992) and ADF/cofilin (Nishida, 1985) inhibit this nucleotide exchange and, therefore, prevent this ‘recharging’ process. Profilin catalytically promotes the nucleotide exchange at G-actin (Goldschmidt-Clermont et al., 1991; Mockrin and Korn, 1980; Nishida, 1985) (Fig. 3) and, therefore, enhances actin turnover in the presence of thymosin β4 (Carlier et al., 1993; Pantaloni and Carlier, 1993) or ADF/cofilin (Blanchoin and Pollard, 1998; Didry et al., 1998). Similarly, CAP is capable of enhancing nucleotide exchange of G-actin and, thus, actin turnover in the presence of ADF/cofilin (Balcer et al., 2003; Chaudhry et al., 2007; Moriyama and Yahara, 2002; Nomura et al., 2012; Quintero-Monzon et al., 2009) (Fig. 3). Such a functional similarity between CAP and profilin was suggested by an early study that showed that the cytoskeletal phenotypes of Srv2/CAP-deficient yeast cells can be suppressed by overexpression of profilin (Vojtek et al., 1991). However, in plants and algae, profilins lack this nucleotide–exchange activity (Kovar et al., 2000; Kovar et al., 2001; Perelroizen et al., 1996) and CAP is the only known nucleotide exchanger for G-actin in plants (Chaudhry et al., 2007). However, unlike profilin, which preferentially binds to ATP–G-actin (Pantaloni and Carlier, 1993), CAP binds to ADP–G-actin with a relatively high affinity and effectively competes with ADF/cofilin for binding to ADP–G-actin (Mattila et al., 2004; Moriyama and Yahara, 2002). Yeast profilin has a much weaker nucleotide-exchange activity than mammalian profilin (Eads et al., 1998; Wen et al., 2008) and fails to enhance nucleotide exchange when ADF/cofilin is bound to ADP–G-actin (Chaudhry et al., 2010). Thus, a sequential processing mechanism of actin monomers by CAP (for ADP/ATP exchange) and profilin (for the incorporation of ATP–G-actin to actin barbed-ends) has been proposed in yeast (Mattila et al., 2004). In addition, profilin binds to P1 of Srv2/CAP and collaborates in polarized actin localization in yeast (Bertling et al., 2007) (Fig. 3). However, the functional relationship between CAP and profilin is not well understood in other organisms. For instance, mammalian profilin enhances the nucleotide exchange of G-actin to much faster rates than yeast profilin (Wen et al., 2008), but it is unclear whether profilin and CAP function in a redundant or hierarchical manner to regulate actin turnover in mammalian cells.

Fig. 3.

Regulation of actin filament dynamics by CAP, ADF/cofilin, profilin and AIP1. CAP competes with ADF/cofilin for binding to G-actin (right) and promotes its nucleotide exchange (top). CAP also promotes severing of ADF/cofilin-bound actin filaments (bottom). These activities are similar to those exerted by profilin and AIP1. ADF/cofilin cooperatively binds to actin filaments and severs them at the boundary between ADF/cofilin-bound and the bare segments (Elam et al., 2013). Note that the molecular organization of the CAP oligomer is hypothetical. Also note that CAP binds to actin monomers in an oligomeric form but, for simplicity, only a monomeric CAP is shown as a G-actin-bound form.

Fig. 3.

Regulation of actin filament dynamics by CAP, ADF/cofilin, profilin and AIP1. CAP competes with ADF/cofilin for binding to G-actin (right) and promotes its nucleotide exchange (top). CAP also promotes severing of ADF/cofilin-bound actin filaments (bottom). These activities are similar to those exerted by profilin and AIP1. ADF/cofilin cooperatively binds to actin filaments and severs them at the boundary between ADF/cofilin-bound and the bare segments (Elam et al., 2013). Note that the molecular organization of the CAP oligomer is hypothetical. Also note that CAP binds to actin monomers in an oligomeric form but, for simplicity, only a monomeric CAP is shown as a G-actin-bound form.

In human CAP1 and yeast Srv2/CAP, HFD in the N-terminal half of CAP binds to the ADF/cofilin-actin complex (Fig. 3) and augments the ADF/cofilin-mediated turnover of actin filaments (Moriyama and Yahara, 2002; Quintero-Monzon et al., 2009). These results were originally interpreted as evidence for an active involvement of HFD in the dissociation of the ADF/cofilin-actin complex. However, a recent study shows that HFD is not required for the recycling of ADF/cofilin and G-actin but, instead, has a separate function in actin filament severing (Chaudhry et al., 2013), which is discussed below.

The sites necessary for monomer sequestering and nucleotide exchange activities of CAP reside in its C-terminal half that contains WH2 and CARP (Fig. 1A). The C-terminal half of CAP can bind to one molecule of G-actin (Freeman et al., 1995), and WH2 and CARP bind independently to G-actin (Chaudhry et al., 2010; Makkonen et al., 2013; Mattila et al., 2004; Peche et al., 2013), suggesting that they separately contact G-actin. An N-terminal part of WH2 from other proteins, such as WASP, was found to bind to G-actin at a cleft between its subdomains 1 and 3 (Dominguez, 2007), and WH2 of CAP might bind to G-actin in a similar manner. Although a CARP-binding site on G-actin has not been determined, mutations in its subdomains 2 and 4 – which are the candidate sites for CARP binding – disrupt its interaction with Srv2/CAP in yeast (Amberg et al., 1995). Although the CARP domain is sufficient to promote the nucleotide exchange of ADP–G-actin (Makkonen et al., 2013), WH2 is necessary to facilitate nucleotide exchange when ADF/cofilin is bound to ADP–G-actin (Chaudhry et al., 2010; Nomura and Ono, 2013). Since ADF/cofilin and WH2 share the same binding site on G-actin (Dominguez and Holmes, 2011), WH2 might be needed for CAP to be able to compete with ADF/cofilin for G-actin binding. Nonetheless, the structural basis for the acceleration of the nucleotide exchange of G-actin by CAP is largely unknown.

Effects of CAP on F-actin – enhancement of filament severing

CAP also directly interacts with actin filaments and enhances filament disassembly (Moriyama and Yahara, 2002). Recent microscopic observations have clearly demonstrated that CAP induces the severing of actin filaments (Fig. 3) (Chaudhry et al., 2013; Normoyle and Brieher, 2012). Mammalian CAP1 alone can sever actin filaments at an acidic pH but not at neutral pH (Normoyle and Brieher, 2012), whereas mammalian and avian ADF/cofilins efficiently sever actin filaments at basic pH but not at neutral and acidic pHs (Hawkins et al., 1993; Hayden et al., 1993; Yonezawa et al., 1985). However, when CAP1 and ADF/cofilin are combined, they promote severing of actin filaments within a wide pH range (Normoyle and Brieher, 2012). Thus, either CAP1 or ADF/cofilin alone might function as a pH-dependent F-actin disassembly factor, but CAP1 and ADF/cofilin might act together as pH-independent F-actin disassembly factors. Yeast Srv2/CAP also promotes the severing of ADF/cofilin-bound actin filaments and this activity is mediated by the N-terminal half of Srv2/CAP that contains the coiled-coil region and the HFD (Chaudhry et al., 2013). This activity of CAP resembles that of actin-interacting protein 1 (AIP1), which promotes the severing and disassembly of ADF/cofilin-bound actin filaments (Ono, 2003) (Fig. 3). Indeed, a genetic study in yeast suggests redundant functions of the N-terminal half of Srv2/CAP and AIP1 (Chaudhry et al., 2013). However, the mechanism of how actin filaments are severed by the HFD of CAP still remains unclear. Functionally important surface residues in HFD of yeast Srv2/CAP have been mapped to the surface of the HDF protein, opposite to the location of its N- and C-termini. </emph>(Quintero-Monzon et al., 2009) (Fig. 1B), but how these residues interact with ADF/cofilin-bound F-actin is unknown. In mammalian CAP2, the WH2 domain is sufficient to sever actin filaments in the absence of ADF/cofilin at a neutral pH (Peche et al., 2013), but its requirement for filament severing in the context of full-length CAP2 still needs to be determined.

Oligomerization of CAP and its effect on actin regulation

Both native and recombinant forms of CAP dimerize or oligomerize, and oligomerization of CAP has been implicated in its function to regulate actin. In yeast cells, Srv2/CAP (55 kDa) and actin (42 kDa) form a high-molecular-mass complex of ∼600 kDa (Balcer et al., 2003; Yang et al., 1999) that contains six Srv2/CAP and six actin molecules (Balcer et al., 2003; Quintero-Monzon et al., 2009). Mammalian CAP1 is also always tightly associated with actin monomers in tissues or cells that need to be dissociated using urea when CAP1 is to be purified (Gieselmann and Mann, 1992; Moriyama and Yahara, 2002; Normoyle and Brieher, 2012), which suggests that the formation of a high-molecular-mass CAP-actin complex is a conserved feature. CAP can also dimerize or oligomerize in the absence of actin and, as mentioned above, three separate regions are known to mediate self-association: the N-terminal coiled-coil region (Quintero-Monzon et al., 2009; Yu et al., 1999), the HFD (Ksiazek et al., 2003; Yusof et al., 2005; Yusof et al., 2006) (Fig. 1B) and the C-terminal dimerization motif (Dodatko et al., 2004) (Fig. 1C). Furthermore, a study, in which short CAP fragments where used, suggests that the N-terminal part of CAP can bind to both the N-terminal and C-terminal sections of a second CAP (Hubberstey et al., 1996). A mutational study of Dictyostelium CAP suggests that several configurations of a CAP dimer are possible (Yusof et al., 2006). Therefore, a high number of CAP oligomer arrangements are possible, but the actual molecular architecture of the physiological CAP oligomer and CAP-actin complex is currently unknown. Nonetheless, disruption of CAP oligomerization through mutation of its N-terminal coiled-coil region impairs the activity of CAP to promote the disassembly of ADF/cofilin-bound actin filaments (Chaudhry et al., 2013; Quintero-Monzon et al., 2009) and deletion of its C-terminal dimerization motif weakens its binding to actin (Zelicof et al., 1996). Thus, oligomerization of CAP enhances its function to regulate actin to some extent, but the structure-function relationship of CAP oligomers and their activities still remains be elucidated.

The Srv2/CAP-actin complex purified from yeast cells has distinct structural features that suggest a highly ordered assembly of CAP molecules within the complex. Its analysis by electron microscopy shows that there are at least four small lobes in the core and two large peripheral lobes (Balcer et al., 2003), but these structural features could not be correlated with the molecular domains of Srv2 and actin. The purified N-terminal section of yeast Srv2/CAP, which contains the coiled-coil region and the HFD that can sever ADF/cofilin-bound actin filaments, oligomerizes into a hexamer. The latter has a radial pattern composed of six blade-like structures that resemble a wheel-type ‘shuriken’, a throwing weapon used by Japanese Ninja (Chaudhry et al., 2013). However, when the coiled-coil region, which mediates its oligomerization, is absent, HFD alone only has a considerably weaker severing activity, suggesting that CAP oligomerization enhances its activity. Thus, additional structural and functional information for CAP oligomers are required before we fully understand how CAP oligomerization promotes its activities.

A link between cellular signaling and the actin cytoskeleton

As noted above, the N- and C-termini of Srv2/CAP independently regulate the Ras–cAMP signalling pathway and the actin cytoskeleton in yeast, respectively, and link these two mechanisms (Freeman et al., 1995; Gerst et al., 1991; Mintzer and Field, 1994). Srv2/CAP mediates the recruitment of Ras and adenylyl cyclase to actin filaments. Interestingly, actin-anchored Ras is an active GTP-bound form and stimulates the production of cAMP, which results in the production of reactive oxygen species and is followed by apoptosis (Gourlay and Ayscough, 2006). The function of CAP as a link between adenylyl cyclase and actin has also been shown in other fungi such as Candida albicans (Wang et al., 2010; Zou et al., 2010) and Magnaporthe oryzae (a rice pathogen) (Zhou et al., 2012). In Dictyostelium, although a direct binding between adenylyl cyclase and CAP has not been detected, CAP also regulates both intracellular cAMP signaling and chemotactic behaviors, and potentially links these two mechanisms (Noegel et al., 2004; Noegel et al., 1999; Sultana et al., 2009; Sultana et al., 2012). However, it is unknown whether CAP is involved in cAMP signaling in animal cells.

As mentioned above, CAP also binds to the SH3 domain of the Abl tyrosine kinase with its proline-rich P1 region (Freeman et al., 1996). In Drosophila neuronal development, CAP (Capulet/Act up) and Abl together restrict the midline crossing of axons, suggesting that CAP and Abl are part of a complex that links the growth-cone-repellent signal to changes in the actin filament dynamics (Wills et al., 2002). It has been shown that rat CAP1 is enriched in the neuronal growth cones (Lu et al., 2011; Nozumi et al., 2009), suggesting that CAP is a conserved regulator of neuronal growth. Mutations in Drosophila CAP also disturb oocyte polarity (Baum et al., 2000), epithelial cell polarity (Baum and Perrimon, 2001), Hedgehog signaling (Benlali et al., 2000), Hippo signaling (Fernández et al., 2011) and eye development (Benlali et al., 2000; Lin et al., 2012). However, it is not clear whether these phenotypic changes are owing to the direct role of CAP in certain signaling pathways or because of secondary effects of disorganized actin filaments. CAP-mutant Drosophila cells accumulate excessive actin filaments that arise owing of impaired actin monomer sequestration (Baum and Perrimon, 2001; Benlali et al., 2000). However, a study in Drosophila neuronal dendrites shows that cofilin (Twinstar) localizes to accumulated actin filaments in CAP-null cells (Medina et al., 2008) in a manner similar to that in mammalian cofilin within CAP1-depleted cells (Bertling et al., 2004), suggesting a role of CAP in the correct localization of ADF/cofilin. Therefore, revisiting some of previously characterized CAP-mutant phenotypes in Drosophila in the context of actin filament turnover might reveal in vivo functions of CAP that have not previously been described.

CAP in cell migration, cytokinesis and cancer

CAP has been implicated in both cell migration and cytokinesis, but its precise role in these processes is not understood. As noted above, CAP is enriched in actin-rich lamellipodia of migrating cells, both in Dictyostelium (Gottwald et al., 1996; Noegel et al., 1999) and cultured mammalian cells (Bertling et al., 2004; Freeman and Field, 2000; Moriyama and Yahara, 2002; Zelicof et al., 1996). The localization of CAP to the leading edge of the cell is mediated through its N-terminal part in an unknown mechanism (Freeman and Field, 2000; Moriyama and Yahara, 2002; Noegel et al., 1999). Knockdown of CAP strongly inhibits the formation of lamellipodia in Drosophila S2 cells (Rogers et al., 2003). Knockdown of CAP also reduces the rate of cell migration in Dictyostelium (Noegel et al., 1999) and cultured mammalian cells (Bertling et al., 2004; Yamazaki et al., 2009). These reports suggest that CAP enhances lamellipodial protrusion by promoting actin filament dynamics during cell migration. However, CAP also localizes to the posterior cortex of migrating Dictyostelium cells (Noegel et al., 1999) and promotes the formation of stress fibers in mammalian fibroblasts (Freeman and Field, 2000), suggesting that it has roles in different aspects of actin dynamics within motile cells; but how exactly it regulates cell migration is still not fully understood.

Reduction of CAP protein levels in Dictyostelium leads to an increased number of multinucleated cells, suggestive of a cytokinesis defect (Noegel et al., 1999). However, localization of CAP to cleavage furrows or other cytokinetic structures has not been reported. Deletion of the CAP gene in budding yeast or fission yeast causes growth defects at a high temperature or under certain nutrient conditions (Fedor-Chaiken et al., 1990; Field et al., 1990; Kawamukai et al., 1992), but it is unknown whether these are caused by cytokinesis defects.

Cell migration and cytokinesis are aberrantly regulated in cancer cells. Mammalian CAP isoforms are overexpressed in several types of cancer, suggesting a correlation between increased levels of CAP and occurrence of cancer phenotypes. For instance, CAP1 is overexpressed in pancreatic cancer (Yamazaki et al., 2009) and breast cancer (Patsialou et al., 2012), and CAP2 has been found to be overexpressed in hepatocellular carcinoma (Effendi et al., 2013; Shibata et al., 2006). Therefore, if a functional correlation between CAP levels and cancer phenotypes can be established, CAP might serve as a diagnostic marker or therapeutic target for certain types of cancer. In addition, a recent study has shown that CAP1 interacts with the tumor suppressor BRCA1-associated RING domain 1 (BARD1) (Woods et al., 2012), further suggesting an involvement of CAP1 in tumorigenesis. However, the functional significance of this interaction is unknown.

Sarcomere organization in striated muscle

In nematodes and vertebrates, one of the two CAP isoforms is predominantly expressed in striated muscle and has an essential role in the assembly of sarcomeres – the basic contractile units. For example, vertebrate CAP2 is expressed at high levels in certain tissues, including skeletal and cardiac muscles of mammals (Bertling et al., 2004; Peche et al., 2007; Swiston et al., 1995), frogs (Xenopus) (Wolanski et al., 2009), and zebrafish (Effendi et al., 2013). Interestingly, the expression of mouse CAP2 is prominently induced during in vitro differentiation of embryonic stem cells into cardiomyocytes, indicating that CAP2 is a marker for cardiac development (Christoforou et al., 2008; Terami et al., 2007). In mammalian skeletal and cardiac muscles, CAP2 is localized in the nucleus and diffused in the cytoplasm in skeletal myoblasts and embryonic cardiomyocytes, whereas, in differentiated skeletal myotubes and cardiomyocytes, it localizes to the M-lines, which are at the center of the thick filaments (Peche et al., 2007). CAP2-null mice die earlier than wild-type mice and show a phenotype that is typically associated with dilated cardiomyopathy, including enlarged heart with dilated ventricle and atrium, increased fibrosis, and impaired heart function (Peche et al., 2013; Jeffrey Field, University of Pennsylvania, personal communication). Furthermore, the myofibrils in skeletal and cardiac muscles from CAP2-null mice are disorganized with disturbed M-lines, thin filaments and Z-discs (Peche et al., 2013). Thus, although it is clear that mammalian CAP2 is important for sarcomere organization in striated muscle, the mechanism by which it localizes to the M-line, and its significance, as well as its role in the regulation of sarcomeric actin filaments are not understood. The human CAP2 gene (CAP2) is located at the chromosome region 6p22.3, and deletions at this region cause several clinical phenotypes including heart defects (Bremer et al., 2009). Therefore, molecular genetics approaches that target human CAP2 might reveal specific genetic lesions that are associated with heart diseases.

In the nematode C. elegans, the CAP isoform CAS-1 is predominantly expressed in striated muscle and localizes to the M-lines, where it is required for sarcomeric actin organization (Nomura et al., 2012). CAS-1-null worms are homozygous lethal at late embryonic to early larval stages and exhibit severe disorganization of muscle actin filaments with formation of actin aggregates (Nomura et al., 2012). In CAS-1-null worms, the muscle-specific ADF/cofilin UNC-60B is mislocalized to these actin aggregates. UNC-60B is also a crucial regulator of sarcomeric actin organization (Ono et al., 2003; Ono et al., 1999), and its mislocalization in CAS-1-null muscle strongly suggests that CAS-1 is required for correct function of UNC-60B. Thus, the genetic studies in C. elegans provide evidence that CAS-1 cooperates with UNC-60B to promote sarcomere organization in striated muscle.

Thus, CAP is a conserved regulator of sarcomere organization. Sarcomeric actin filaments in striated muscle requires a dynamic regulation of actin assembly and disassembly to maintain myofibril integrity (Ono, 2010). The presence of a set of muscle-specific regulators of actin dynamics suggests that the assembly and maintenance of sarcomeres require a specific actin-regulatory mechanism. Mammalian cofilin-2 is a muscle-specific ADF/cofilin (Ono et al., 1994; Vartiainen et al., 2002). Mutations in human cofilin-2 (CFL2) cause nemaline myopathy (Agrawal et al., 2007), and cofilin-2-null mice show impaired maintenance of myofibrils in skeletal muscle (Agrawal et al., 2012). Therefore, it will be interesting to determine whether cofilin-2 and CAP2 specifically cooperate to regulate sarcomere assembly in mammalian muscle.

Plant development

In plants, CAP is the only known enhancer of G-actin nucleotide exchange, because plant profilins do not possess this functionality (Chaudhry et al., 2007). Overexpression of Arabidopsis CAP1 causes a reduction in the size and number of cells, resulting in smaller leaves and petioles (Barrero et al., 2002). However, overexpression of Arabidopsis CAP1 in tobacco plants leads to a reduced organ size that is primarily due to reduced size but not number of cells (Barrero et al., 2003). In addition, CAP1-null mutations in Arabidopsis cause a number of developmental defects, including reduced statue and reduced elongation of pollen tubes and root hairs (Deeks et al., 2007), and actin bundle organization is disrupted in root hairs and epidermis. The CAP1-null mutation and mutations in the Arp2/3 complex pathway synergistically enhance morphological defects, suggesting that CAP1 is actively involved in the Arp2/3-dependent actin regulation, or that CAP1 and Arp2/3 are involved in the same process (Deeks et al., 2007). Plants have multiple CAP isoforms, but their functional differences have not been elucidated.

Other biological functions of CAPs

CAP also has other functions, such as in endocytosis or Drosophila development, as discussed in earlier reviews (Hubberstey and Mottillo, 2002; Stevenson and Theurkauf, 2000). More recently, it has been shown that CAP is essential for the correct functioning of the endo-lysosomal system in Dictyostelium (Sultana et al., 2005), suggesting that CAP is widely involved in membrane trafficking. In addition, upon induction of apoptosis, mammalian CAP1 translocates to the mitochondria where it further promotes the execution of apoptosis (Wang et al., 2008). It is worth noting that cofilin also translocates to the mitochondria and promotes apoptosis (Chua et al., 2003), suggesting that CAP1 and cofilin are pro-apoptotic proteins, although it remains unclear how exactly they promote apoptosis. Intriguingly, CAP1 has been identified as an autoantigen in patients with systemic lupus erythematosus (Frampton et al., 2000) and rheumatoid arthritis (Kinloch et al., 2005), suggesting that CAP1 also has an extracellular function.

Recent studies have revealed that CAP exerts active roles in promoting actin filament dynamics. The wide distribution of CAPs in different cell types and organisms suggest that CAPs are common regulators of actin dynamics; thus, they are expected to be involved in many diverse processes that require actin reorganization (Fig. 4). Some high-throughput studies have provided hints for these, so far not described, functions of CAPs. For example, upregulation of CAP1 has been reported in the brains of schizophrenic patients (Martins-de-Souza et al., 2010) and in macrophages treated with the anti-tumor agent lapachol (Oliveira et al., 2012), suggesting that CAP1-mediated actin reorganization is induced under these conditions. Despite the fact that several proteins are known to bind to CAP (Fig. 1), the mechanisms, by which CAP activity is regulated, are poorly understood. Mouse CAP1 is phosphorylated (Guo-Lei Zhou, Arkansas State University, and Jeffrey Field, University of Pennsylvania, personal communication), suggesting that activity and localization of CAP can be regulated through kinase signaling pathways. CAP1 is also an intracellular substrate of a matrix metalloproteinase (Cauwe et al., 2008), indicating that the protein level of CAP is controlled by proteolysis. It is also important to better understand the functional relationship between CAP and other actin-regulatory proteins. CAP can participate in both assembly and disassembly of actin filaments, and other actin-regulatory proteins may affect how CAP controls actin dynamics. Thus, further studies on CAPs are most likely to reveal a greater complexity of the regulatory mechanism that governs actin filament dynamics.

Fig. 4.

Biological functions of CAP.

Fig. 4.

Biological functions of CAP.

Box 1. Role of adenine nucleotides (ATP and ADP) in actin filament dynamics

One actin molecule binds to one molecule of ATP or ADP, and the nucleotide status of actin strongly affects its properties in both polymerization and depolymerization. ATP-bound monomeric globular (G-) actin (ATP–G-actin) is more competent for polymerization than ADP–G-actin. Polymerized filamentous (F-) actin has two ends with different properties, the pointed (minus) and barbed (plus) ends. ATP–G-actin binds to the barbed end of F-actin with higher affinity than to the pointed end. When a concentration of ATP–G-actin falls below the dissociation constant for ATP–G-actin with the pointed end of F-actin, actin constantly depolymerizes from the pointed end and polymerizes at the barbed end without an apparent change in the ratio of G- and F-actin, which is called treadmilling. However, ADP–G-actin binds to pointed and barbed ends with nearly equal affinity, and therefore it does not undergo treadmilling. Furthermore, exchange of actin-bound nucleotide occurs primarily at G-actin, whereas hydrolysis of ATP into ADP occurs primarily at F-actin. The newly depolymerized actin monomers are predominantly bound by ADP. When free ATP is present, ADP–G-actin can ‘recharge’ itself by rapidly exchanging ADP for ATP and, thus, maintaining the treadmilling cycle. However, when no free ATP is available, ADP–G-actin cannot recharge itself, which prevents the treadmilling cycle.

Box 2. Actin depolymerizing factor (ADF)/cofilin

ADF/cofilin is an actin-regulatory protein that promotes the turnover of actin filaments. It binds to G-actin and F-actin in a molar ratio of 1∶1. ADF/cofilin severs F-actin and promotes dissociation of G-actin from filament ends. Because F-actin is stable, enhancement of F-actin disassembly by ADF/cofilin can supply a pool of G-actin for new rounds of actin polymerization and can promote persistent turnover of actin filaments. ADF/cofilin preferentially binds to an actin molecule bound by ADP and inhibits nucleotide exchange of G-actin. Several other actin regulatory proteins including CAP, profilin and AIP1 cooperate with ADF/cofilin in order to regulate the dynamics of actin filaments as shown in Fig. 3.

The author thanks Kazumi Nomura for critical reading of the manuscript, and Guo-Lei Zhou and Jeffrey Field for communication of unpublished results.

Funding

S.O. is supported by a grant from the National Institutes of Health (R01 AR48615). Deposited in PMC for release after 12 months.

Agrawal
P. B.
,
Greenleaf
R. S.
,
Tomczak
K. K.
,
Lehtokari
V. L.
,
Wallgren-Pettersson
C.
,
Wallefeld
W.
,
Laing
N. G.
,
Darras
B. T.
,
Maciver
S. K.
,
Dormitzer
P. R.
 et al. (
2007
).
Nemaline myopathy with minicores caused by mutation of the CFL2 gene encoding the skeletal muscle actin-binding protein, cofilin-2.
Am. J. Hum. Genet.
80
,
162
167
.
Agrawal
P. B.
,
Joshi
M.
,
Savic
T.
,
Chen
Z.
,
Beggs
A. H.
(
2012
).
Normal myofibrillar development followed by progressive sarcomeric disruption with actin accumulations in a mouse Cfl2 knockout demonstrates requirement of cofilin-2 for muscle maintenance.
Hum. Mol. Genet.
21
,
2341
2356
.
Amberg
D. C.
,
Basart
E.
,
Botstein
D.
(
1995
).
Defining protein interactions with yeast actin in vivo.
Nat. Struct. Biol.
2
,
28
35
.
Balcer
H. I.
,
Goodman
A. L.
,
Rodal
A. A.
,
Smith
E.
,
Kugler
J.
,
Heuser
J. E.
,
Goode
B. L.
(
2003
).
Coordinated regulation of actin filament turnover by a high-molecular-weight Srv2/CAP complex, cofilin, profilin, and Aip1.
Curr. Biol.
13
,
2159
2169
.
Barrero
R. A.
,
Umeda
M.
,
Yamamura
S.
,
Uchimiya
H.
(
2002
).
Arabidopsis CAP regulates the actin cytoskeleton necessary for plant cell elongation and division.
Plant Cell
14
,
149
163
.
Barrero
R. A.
,
Umeda
M.
,
Yamamura
S.
,
Uchimiya
H.
(
2003
).
Over-expression of Arabidopsis CAP causes decreased cell expansion leading to organ size reduction in transgenic tobacco plants.
Ann. Bot. (Lond.)
91
,
599
603
.
Baum
B.
,
Perrimon
N.
(
2001
).
Spatial control of the actin cytoskeleton in Drosophila epithelial cells.
Nat. Cell Biol.
3
,
883
890
.
Baum
B.
,
Li
W.
,
Perrimon
N.
(
2000
).
A cyclase-associated protein regulates actin and cell polarity during Drosophila oogenesis and in yeast.
Curr. Biol.
10
,
964
973
.
Benlali
A.
,
Draskovic
I.
,
Hazelett
D. J.
,
Treisman
J. E.
(
2000
).
act up controls actin polymerization to alter cell shape and restrict Hedgehog signaling in the Drosophila eye disc.
Cell
101
,
271
281
.
Bertling
E.
,
Hotulainen
P.
,
Mattila
P. K.
,
Matilainen
T.
,
Salminen
M.
,
Lappalainen
P.
(
2004
).
Cyclase-associated protein 1 (CAP1) promotes cofilin-induced actin dynamics in mammalian nonmuscle cells.
Mol. Biol. Cell
15
,
2324
2334
.
Bertling
E.
,
Quintero-Monzon
O.
,
Mattila
P. K.
,
Goode
B. L.
,
Lappalainen
P.
(
2007
).
Mechanism and biological role of profilin-Srv2/CAP interaction.
J. Cell Sci.
120
,
1225
1234
.
Blanchoin
L.
,
Pollard
T. D.
(
1998
).
Interaction of actin monomers with Acanthamoeba actophorin (ADF/cofilin) and profilin.
J. Biol. Chem.
273
,
25106
25111
.
Bremer
A.
,
Schoumans
J.
,
Nordenskjöld
M.
,
Anderlid
B. M.
,
Giacobini
M.
(
2009
).
An interstitial deletion of 7.1Mb in chromosome band 6p22.3 associated with developmental delay and dysmorphic features including heart defects, short neck, and eye abnormalities.
Eur. J. Med. Genet.
52
,
358
362
.
Bugyi
B.
,
Carlier
M. F.
(
2010
).
Control of actin filament treadmilling in cell motility.
Annu. Rev. Biophys.
39
,
449
470
.
Carlier
M. F.
,
Jean
C.
,
Rieger
K. J.
,
Lenfant
M.
,
Pantaloni
D.
(
1993
).
Modulation of the interaction between G-actin and thymosin beta 4 by the ATP/ADP ratio: possible implication in the regulation of actin dynamics.
Proc. Natl. Acad. Sci. USA
90
,
5034
5038
.
Cauwe
B.
,
Martens
E.
,
Van den Steen
P. E.
,
Proost
P.
,
Van Aelst
I.
,
Blockmans
D.
,
Opdenakker
G.
(
2008
).
Adenylyl cyclase-associated protein-1/CAP1 as a biological target substrate of gelatinase B/MMP-9.
Exp. Cell Res.
314
,
2739
2749
.
Chaudhry
F.
,
Guérin
C.
,
von Witsch
M.
,
Blanchoin
L.
,
Staiger
C. J.
(
2007
).
Identification of Arabidopsis cyclase-associated protein 1 as the first nucleotide exchange factor for plant actin.
Mol. Biol. Cell
18
,
3002
3014
.
Chaudhry
F.
,
Little
K.
,
Talarico
L.
,
Quintero-Monzon
O.
,
Goode
B. L.
(
2010
).
A central role for the WH2 domain of Srv2/CAP in recharging actin monomers to drive actin turnover in vitro and in vivo.
Cytoskeleton (Hoboken)
67
,
120
133
.
Chaudhry
F.
,
Breitsprecher
D.
,
Little
K.
,
Sharov
G.
,
Sokolova
O.
,
Goode
B. L.
(
2013
).
Srv2/cyclase-associated protein forms hexameric shurikens that directly catalyze actin filament severing by cofilin.
Mol. Biol. Cell
24
,
31
41
.
Christoforou
N.
,
Miller
R. A.
,
Hill
C. M.
,
Jie
C. C.
,
McCallion
A. S.
,
Gearhart
J. D.
(
2008
).
Mouse ES cell-derived cardiac precursor cells are multipotent and facilitate identification of novel cardiac genes.
J. Clin. Invest.
118
,
894
903
.
Chua
B. T.
,
Volbracht
C.
,
Tan
K. O.
,
Li
R.
,
Yu
V. C.
,
Li
P.
(
2003
).
Mitochondrial translocation of cofilin is an early step in apoptosis induction.
Nat. Cell Biol.
5
,
1083
1089
.
De La Cruz
E. M.
,
Pollard
T. D.
(
2001
).
Structural biology. Actin’ up.
Science
293
,
616
618
.
Deeks
M. J.
,
Rodrigues
C.
,
Dimmock
S.
,
Ketelaar
T.
,
Maciver
S. K.
,
Malhó
R.
,
Hussey
P. J.
(
2007
).
Arabidopsis CAP1 – a key regulator of actin organisation and development.
J. Cell Sci.
120
,
2609
2618
.
Didry
D.
,
Carlier
M. F.
,
Pantaloni
D.
(
1998
).
Synergy between actin depolymerizing factor/cofilin and profilin in increasing actin filament turnover.
J. Biol. Chem.
273
,
25602
25611
.
Dodatko
T.
,
Fedorov
A. A.
,
Grynberg
M.
,
Patskovsky
Y.
,
Rozwarski
D. A.
,
Jaroszewski
L.
,
Aronoff-Spencer
E.
,
Kondraskina
E.
,
Irving
T.
,
Godzik
A.
 et al. (
2004
).
Crystal structure of the actin binding domain of the cyclase-associated protein.
Biochemistry
43
,
10628
10641
.
Dominguez
R.
(
2007
).
The beta-thymosin/WH2 fold: multifunctionality and structure.
Ann. New York Acad. Sci.
1112
,
86
94
.
Dominguez
R.
,
Holmes
K. C.
(
2011
).
Actin structure and function.
Annu. Rev. Biophys.
40
,
169
186
.
Eads
J. C.
,
Mahoney
N. M.
,
Vorobiev
S.
,
Bresnick
A. R.
,
Wen
K. K.
,
Rubenstein
P. A.
,
Haarer
B. K.
,
Almo
S. C.
(
1998
).
Structure determination and characterization of Saccharomyces cerevisiae profilin.
Biochemistry
37
,
11171
11181
.
Effendi
K.
,
Yamazaki
K.
,
Mori
T.
,
Masugi
Y.
,
Makino
S.
,
Sakamoto
M.
(
2013
).
Involvement of hepatocellular carcinoma biomarker, cyclase-associated protein 2 in zebrafish body development and cancer progression.
Exp. Cell Res.
319
,
35
44
.
Elam
W. A.
,
Kang
H.
,
De la Cruz
E. M.
(
2013
).
Biophysics of actin filament severing by cofilin.
FEBS Lett.
587
,
1215
1219
.
Fedor-Chaiken
M.
,
Deschenes
R. J.
,
Broach
J. R.
(
1990
).
SRV2, a gene required for RAS activation of adenylate cyclase in yeast.
Cell
61
,
329
340
.
Fernández
B. G.
,
Gaspar
P.
,
Brás-Pereira
C.
,
Jezowska
B.
,
Rebelo
S. R.
,
Janody
F.
(
2011
).
Actin-Capping Protein and the Hippo pathway regulate F-actin and tissue growth in Drosophila.
Development
138
,
2337
2346
.
Field
J.
,
Vojtek
A.
,
Ballester
R.
,
Bolger
G.
,
Colicelli
J.
,
Ferguson
K.
,
Gerst
J.
,
Kataoka
T.
,
Michaeli
T.
,
Powers
S.
 et al. (
1990
).
Cloning and characterization of CAP, the S. cerevisiae gene encoding the 70 kd adenylyl cyclase-associated protein.
Cell
61
,
319
327
.
Frampton
G.
,
Moriya
S.
,
Pearson
J. D.
,
Isenberg
D. A.
,
Ward
F. J.
,
Smith
T. A.
,
Panayiotou
A.
,
Staines
N. A.
,
Murphy
J. J.
(
2000
).
Identification of candidate endothelial cell autoantigens in systemic lupus erythematosus using a molecular cloning strategy: a role for ribosomal P protein P0 as an endothelial cell autoantigen.
Rheumatology (Oxford)
39
,
1114
1120
.
Freeman
N. L.
,
Field
J.
(
2000
).
Mammalian homolog of the yeast cyclase associated protein, CAP/Srv2p, regulates actin filament assembly.
Cell Motil. Cytoskeleton
45
,
106
120
.
Freeman
N. L.
,
Chen
Z.
,
Horenstein
J.
,
Weber
A.
,
Field
J.
(
1995
).
An actin monomer binding activity localizes to the carboxyl-terminal half of the Saccharomyces cerevisiae cyclase-associated protein.
J. Biol. Chem.
270
,
5680
5685
.
Freeman
N. L.
,
Lila
T.
,
Mintzer
K. A.
,
Chen
Z.
,
Pahk
A. J.
,
Ren
R.
,
Drubin
D. G.
,
Field
J.
(
1996
).
A conserved proline-rich region of the Saccharomyces cerevisiae cyclase-associated protein binds SH3 domains and modulates cytoskeletal localization.
Mol. Cell. Biol.
16
,
548
556
.
Gerst
J. E.
,
Ferguson
K.
,
Vojtek
A.
,
Wigler
M.
,
Field
J.
(
1991
).
CAP is a bifunctional component of the Saccharomyces cerevisiae adenylyl cyclase complex.
Mol. Cell. Biol.
11
,
1248
1257
.
Gieselmann
R.
,
Mann
K.
(
1992
).
ASP-56, a new actin sequestering protein from pig platelets with homology to CAP, an adenylate cyclase-associated protein from yeast.
FEBS Lett.
298
,
149
153
.
Goldschmidt-Clermont
P. J.
,
Machesky
L. M.
,
Doberstein
S. K.
,
Pollard
T. D.
(
1991
).
Mechanism of the interaction of human platelet profilin with actin.
J. Cell Biol.
113
,
1081
1089
.
Goldschmidt-Clermont
P. J.
,
Furman
M. I.
,
Wachsstock
D.
,
Safer
D.
,
Nachmias
V. T.
,
Pollard
T. D.
(
1992
).
The control of actin nucleotide exchange by thymosin β 4 and profilin. A potential regulatory mechanism for actin polymerization in cells.
Mol. Biol. Cell
3
,
1015
1024
.
Gottwald
U.
,
Brokamp
R.
,
Karakesisoglou
I.
,
Schleicher
M.
,
Noegel
A. A.
(
1996
).
Identification of a cyclase-associated protein (CAP) homologue in Dictyostelium discoideum and characterization of its interaction with actin.
Mol. Biol. Cell
7
,
261
272
.
Gourlay
C. W.
,
Ayscough
K. R.
(
2006
).
Actin-induced hyperactivation of the Ras signaling pathway leads to apoptosis in Saccharomyces cerevisiae.
Mol. Cell. Biol.
26
,
6487
6501
.
Hawkins
M.
,
Pope
B.
,
Maciver
S. K.
,
Weeds
A. G.
(
1993
).
Human actin depolymerizing factor mediates a pH-sensitive destruction of actin filaments.
Biochemistry
32
,
9985
9993
.
Hayden
S. M.
,
Miller
P. S.
,
Brauweiler
A.
,
Bamburg
J. R.
(
1993
).
Analysis of the interactions of actin depolymerizing factor with G- and F-actin.
Biochemistry
32
,
9994
10004
.
Heacock
C. S.
,
Bamburg
J. R.
(
1983
).
The quantitation of G- and F-actin in cultured cells.
Anal. Biochem.
135
,
22
36
.
Hliscs
M.
,
Sattler
J. M.
,
Tempel
W.
,
Artz
J. D.
,
Dong
A.
,
Hui
R.
,
Matuschewski
K.
,
Schüler
H.
(
2010
).
Structure and function of a G-actin sequestering protein with a vital role in malaria oocyst development inside the mosquito vector.
J. Biol. Chem.
285
,
11572
11583
.
Hubberstey
A. V.
,
Mottillo
E. P.
(
2002
).
Cyclase-associated proteins: CAPacity for linking signal transduction and actin polymerization.
FASEB J.
16
,
487
499
.
Hubberstey
A.
,
Yu
G.
,
Loewith
R.
,
Lakusta
C.
,
Young
D.
(
1996
).
Mammalian CAP interacts with CAP, CAP2, and actin.
J. Cell. Biochem.
61
,
459
466
.
Jockusch
B. M.
,
Murk
K.
,
Rothkegel
M.
(
2007
).
The profile of profilins.
Rev. Physiol. Biochem. Pharmacol.
159
,
131
149
.
Kawamukai
M.
,
Gerst
J.
,
Field
J.
,
Riggs
M.
,
Rodgers
L.
,
Wigler
M.
,
Young
D.
(
1992
).
Genetic and biochemical analysis of the adenylyl cyclase-associated protein, cap, in Schizosaccharomyces pombe.
Mol. Biol. Cell
3
,
167
180
.
Kinloch
A.
,
Tatzer
V.
,
Wait
R.
,
Peston
D.
,
Lundberg
K.
,
Donatien
P.
,
Moyes
D.
,
Taylor
P. C.
,
Venables
P. J.
(
2005
).
Identification of citrullinated alpha-enolase as a candidate autoantigen in rheumatoid arthritis.
Arthritis Res. Ther.
7
,
R1421
R1429
.
Kinosian
H. J.
,
Selden
L. A.
,
Estes
J. E.
,
Gershman
L. C.
(
1993
).
Nucleotide binding to actin. Cation dependence of nucleotide dissociation and exchange rates.
J. Biol. Chem.
268
,
8683
8691
.
Kovar
D. R.
,
Drøbak
B. K.
,
Staiger
C. J.
(
2000
).
Maize profilin isoforms are functionally distinct.
Plant Cell
12
,
583
598
.
Kovar
D. R.
,
Yang
P.
,
Sale
W. S.
,
Drobak
B. K.
,
Staiger
C. J.
(
2001
).
Chlamydomonas reinhardtii produces a profilin with unusual biochemical properties.
J. Cell Sci.
114
,
4293
4305
.
Ksiazek
D.
,
Brandstetter
H.
,
Israel
L.
,
Bourenkov
G. P.
,
Katchalova
G.
,
Janssen
K. P.
,
Bartunik
H. D.
,
Noegel
A. A.
,
Schleicher
M.
,
Holak
T. A.
(
2003
).
Structure of the N-terminal domain of the adenylyl cyclase-associated protein (CAP) from Dictyostelium discoideum.
Structure
11
,
1171
1178
.
Lila
T.
,
Drubin
D. G.
(
1997
).
Evidence for physical and functional interactions among two Saccharomyces cerevisiae SH3 domain proteins, an adenylyl cyclase-associated protein and the actin cytoskeleton.
Mol. Biol. Cell
8
,
367
385
.
Lin
C. M.
,
Lin
P. Y.
,
Li
Y. C.
,
Hsu
J. C.
(
2012
).
Capulet and Slingshot share overlapping functions during Drosophila eye morphogenesis.
J. Biomed. Sci.
19
,
46
.
Lu
J.
,
Nozumi
M.
,
Takeuchi
K.
,
Abe
H.
,
Igarashi
M.
(
2011
).
Expression and function of neuronal growth-associated proteins (nGAPs) in PC12 cells.
Neurosci. Res.
70
,
85
90
.
Makkonen
M.
,
Bertling
E.
,
Chebotareva
N. A.
,
Baum
J.
,
Lappalainen
P.
(
2013
).
Mammalian and malaria parasite cyclase-associated proteins catalyze nucleotide exchange on G-actin through a conserved mechanism.
J. Biol. Chem.
288
,
984
994
.
Martins-de-Souza
D.
,
Maccarrone
G.
,
Wobrock
T.
,
Zerr
I.
,
Gormanns
P.
,
Reckow
S.
,
Falkai
P.
,
Schmitt
A.
,
Turck
C. W.
(
2010
).
Proteome analysis of the thalamus and cerebrospinal fluid reveals glycolysis dysfunction and potential biomarkers candidates for schizophrenia.
J. Psychiatr. Res.
44
,
1176
1189
.
Mattila
P. K.
,
Quintero-Monzon
O.
,
Kugler
J.
,
Moseley
J. B.
,
Almo
S. C.
,
Lappalainen
P.
,
Goode
B. L.
(
2004
).
A high-affinity interaction with ADP-actin monomers underlies the mechanism and in vivo function of Srv2/cyclase-associated protein.
Mol. Biol. Cell
15
,
5158
5171
.
Matviw
H.
,
Yu
G.
,
Young
D.
(
1992
).
Identification of a human cDNA encoding a protein that is structurally and functionally related to the yeast adenylyl cyclase-associated CAP proteins.
Mol. Cell. Biol.
12
,
5033
5040
.
Mavoungou
C.
,
Israel
L.
,
Rehm
T.
,
Ksiazek
D.
,
Krajewski
M.
,
Popowicz
G.
,
Noegel
A. A.
,
Schleicher
M.
,
Holak
T. A.
(
2004
).
NMR structural characterization of the N-terminal domain of the adenylyl cyclase-associated protein (CAP) from Dictyostelium discoideum.
J. Biomol. NMR
29
,
73
84
.
Medina
P. M.
,
Worthen
R. J.
,
Forsberg
L. J.
,
Brenman
J. E.
(
2008
).
The actin-binding protein capulet genetically interacts with the microtubule motor kinesin to maintain neuronal dendrite homeostasis.
PLoS ONE
3
,
e3054
.
Mintzer
K. A.
,
Field
J.
(
1994
).
Interactions between adenylyl cyclase, CAP and RAS from Saccharomyces cerevisiae.
Cell. Signal.
6
,
681
694
.
Mockrin
S. C.
,
Korn
E. D.
(
1980
).
Acanthamoeba profilin interacts with G-actin to increase the rate of exchange of actin-bound adenosine 5′-triphosphate.
Biochemistry
19
,
5359
5362
.
Moriyama
K.
,
Yahara
I.
(
2002
).
Human CAP1 is a key factor in the recycling of cofilin and actin for rapid actin turnover.
J. Cell Sci.
115
,
1591
1601
.
Nishida
E.
(
1985
).
Opposite effects of cofilin and profilin from porcine brain on rate of exchange of actin-bound adenosine 5′-triphosphate.
Biochemistry
24
,
1160
1164
.
Nishida
Y.
,
Shima
F.
,
Sen
H.
,
Tanaka
Y.
,
Yanagihara
C.
,
Yamawaki-Kataoka
Y.
,
Kariya
K.
,
Kataoka
T.
(
1998
).
Coiled-coil interaction of N-terminal 36 residues of cyclase-associated protein with adenylyl cyclase is sufficient for its function in Saccharomyces cerevisiae ras pathway.
J. Biol. Chem.
273
,
28019
28024
.
Noegel
A. A.
,
Rivero
F.
,
Albrecht
R.
,
Janssen
K. P.
,
Köhler
J.
,
Parent
C. A.
,
Schleicher
M.
(
1999
).
Assessing the role of the ASP56/CAP homologue of Dictyostelium discoideum and the requirements for subcellular localization.
J. Cell Sci.
112
,
3195
3203
.
Noegel
A. A.
,
Blau-Wasser
R.
,
Sultana
H.
,
Müller
R.
,
Israel
L.
,
Schleicher
M.
,
Patel
H.
,
Weijer
C. J.
(
2004
).
The cyclase-associated protein CAP as regulator of cell polarity and cAMP signaling in Dictyostelium.
Mol. Biol. Cell
15
,
934
945
.
Nomura
K.
,
Ono
S.
(
2013
).
ATP-dependent regulation of actin monomer-filament equilibrium by cyclase-associated protein and ADF/cofilin.
Biochem. J.
453
,
249
259
.
Nomura
K.
,
Ono
K.
,
Ono
S.
(
2012
).
CAS-1, a C. elegans cyclase-associated protein, is required for sarcomeric actin assembly in striated muscle.
J. Cell Sci.
125
,
4077
4089
.
Normoyle
K. P.
,
Brieher
W. M.
(
2012
).
Cyclase-associated protein (CAP) acts directly on F-actin to accelerate cofilin-mediated actin severing across the range of physiological pH.
J. Biol. Chem.
287
,
35722
35732
.
Nozumi
M.
,
Togano
T.
,
Takahashi-Niki
K.
,
Lu
J.
,
Honda
A.
,
Taoka
M.
,
Shinkawa
T.
,
Koga
H.
,
Takeuchi
K.
,
Isobe
T.
 et al. (
2009
).
Identification of functional marker proteins in the mammalian growth cone.
Proc. Natl. Acad. Sci. USA
106
,
17211
17216
.
Oliveira
R. A.
,
Correia-Oliveira
J.
,
Tang
L. J.
,
Garcia
R. C.
(
2012
).
A proteomic insight into the effects of the immunomodulatory hydroxynaphthoquinone lapachol on activated macrophages.
Int. Immunopharmacol.
14
,
54
65
.
Ono
S.
(
2003
).
Regulation of actin filament dynamics by actin depolymerizing factor/cofilin and actin-interacting protein 1: new blades for twisted filaments.
Biochemistry
42
,
13363
13370
.
Ono
S.
(
2007
).
Mechanism of depolymerization and severing of actin filaments and its significance in cytoskeletal dynamics.
Int. Rev. Cytol.
258
,
1
82
.
Ono
S.
(
2010
).
Dynamic regulation of sarcomeric actin filaments in striated muscle.
Cytoskeleton (Hoboken)
67
,
677
692
.
Ono
S.
,
Minami
N.
,
Abe
H.
,
Obinata
T.
(
1994
).
Characterization of a novel cofilin isoform that is predominantly expressed in mammalian skeletal muscle.
J. Biol. Chem.
269
,
15280
15286
.
Ono
S.
,
Baillie
D. L.
,
Benian
G. M.
(
1999
).
UNC-60B, an ADF/cofilin family protein, is required for proper assembly of actin into myofibrils in Caenorhabditis elegans body wall muscle.
J. Cell Biol.
145
,
491
502
.
Ono
K.
,
Parast
M.
,
Alberico
C.
,
Benian
G. M.
,
Ono
S.
(
2003
).
Specific requirement for two ADF/cofilin isoforms in distinct actin-dependent processes in Caenorhabditis elegans.
J. Cell Sci.
116
,
2073
2085
.
Pantaloni
D.
,
Carlier
M. F.
(
1993
).
How profilin promotes actin filament assembly in the presence of thymosin β 4.
Cell
75
,
1007
1014
.
Patsialou
A.
,
Wang
Y.
,
Lin
J.
,
Whitney
K.
,
Goswami
S.
,
Kenny
P. A.
,
Condeelis
J. S.
(
2012
).
Selective gene-expression profiling of migratory tumor cells in vivo predicts clinical outcome in breast cancer patients.
Breast Cancer Res.
14
,
R139
.
Paunola
E.
,
Mattila
P. K.
,
Lappalainen
P.
(
2002
).
WH2 domain: a small, versatile adapter for actin monomers.
FEBS Lett.
513
,
92
97
.
Peche
V.
,
Shekar
S.
,
Leichter
M.
,
Korte
H.
,
Schröder
R.
,
Schleicher
M.
,
Holak
T. A.
,
Clemen
C. S.
,
Ramanath-Y
B.
,
Pfitzer
G.
 et al. (
2007
).
CAP2, cyclase-associated protein 2, is a dual compartment protein.
Cell. Mol. Life Sci.
64
,
2702
2715
.
Peche
V. S.
,
Holak
T. A.
,
Burgute
B. D.
,
Kosmas
K.
,
Kale
S. P.
,
Wunderlich
F. T.
,
Elhamine
F.
,
Stehle
R.
,
Pfitzer
G.
,
Nohroudi
K.
 et al. (
2013
).
Ablation of cyclase-associated protein 2 (CAP2) leads to cardiomyopathy.
Cell. Mol. Life Sci.
70
,
527
543
.
Perelroizen
I.
,
Didry
D.
,
Christensen
H.
,
Chua
N. H.
,
Carlier
M. F.
(
1996
).
Role of nucleotide exchange and hydrolysis in the function of profilin in action assembly.
J. Biol. Chem.
271
,
12302
12309
.
Polet
D.
,
Lambrechts
A.
,
Ono
K.
,
Mah
A.
,
Peelman
F.
,
Vandekerckhove
J.
,
Baillie
D. L.
,
Ampe
C.
,
Ono
S.
(
2006
).
Caenorhabditis elegans expresses three functional profilins in a tissue-specific manner.
Cell Motil. Cytoskeleton
63
,
14
28
.
Pollard
T. D.
(
1986
).
Rate constants for the reactions of ATP- and ADP-actin with the ends of actin filaments.
J. Cell Biol.
103
,
2747
2754
.
Pollard
T. D.
,
Cooper
J. A.
(
2009
).
Actin, a central player in cell shape and movement.
Science
326
,
1208
1212
.
Pollard
T. D.
,
Goldberg
I.
,
Schwarz
W. H.
(
1992
).
Nucleotide exchange, structure, and mechanical properties of filaments assembled from ATP-actin and ADP-actin.
J. Biol. Chem.
267
,
20339
20345
.
Poukkula
M.
,
Kremneva
E.
,
Serlachius
M.
,
Lappalainen
P.
(
2011
).
Actin-depolymerizing factor homology domain: a conserved fold performing diverse roles in cytoskeletal dynamics.
Cytoskeleton (Hoboken)
68
,
471
490
.
Quintero-Monzon
O.
,
Jonasson
E. M.
,
Bertling
E.
,
Talarico
L.
,
Chaudhry
F.
,
Sihvo
M.
,
Lappalainen
P.
,
Goode
B. L.
(
2009
).
Reconstitution and dissection of the 600-kDa Srv2/CAP complex: roles for oligomerization and cofilin-actin binding in driving actin turnover.
J. Biol. Chem.
284
,
10923
10934
.
Rogers
S. L.
,
Wiedemann
U.
,
Stuurman
N.
,
Vale
R. D.
(
2003
).
Molecular requirements for actin-based lamella formation in Drosophila S2 cells.
J. Cell Biol.
162
,
1079
1088
.
Schultz
J.
,
Milpetz
F.
,
Bork
P.
,
Ponting
C. P.
(
1998
).
SMART, a simple modular architecture research tool: identification of signaling domains.
Proc. Natl. Acad. Sci. USA
95
,
5857
5864
.
Shibata
R.
,
Mori
T.
,
Du
W.
,
Chuma
M.
,
Gotoh
M.
,
Shimazu
M.
,
Ueda
M.
,
Hirohashi
S.
,
Sakamoto
M.
(
2006
).
Overexpression of cyclase-associated protein 2 in multistage hepatocarcinogenesis.
Clin. Cancer Res.
12
,
5363
5368
.
Stevenson
V. A.
,
Theurkauf
W. E.
(
2000
).
Actin cytoskeleton: putting a CAP on actin polymerization.
Curr. Biol.
10
,
R695
R697
.
Sultana
H.
,
Rivero
F.
,
Blau-Wasser
R.
,
Schwager
S.
,
Balbo
A.
,
Bozzaro
S.
,
Schleicher
M.
,
Noegel
A. A.
(
2005
).
Cyclase-associated protein is essential for the functioning of the endo-lysosomal system and provides a link to the actin cytoskeleton.
Traffic
6
,
930
946
.
Sultana
H.
,
Neelakanta
G.
,
Eichinger
L.
,
Rivero
F.
,
Noegel
A. A.
(
2009
).
Microarray phenotyping places cyclase associated protein CAP at the crossroad of signaling pathways reorganizing the actin cytoskeleton in Dictyostelium.
Exp. Cell Res.
315
,
127
140
.
Sultana
H.
,
Neelakanta
G.
,
Rivero
F.
,
Blau-Wasser
R.
,
Schleicher
M.
,
Noegel
A. A.
(
2012
).
Ectopic expression of cyclase associated protein CAP restores the streaming and aggregation defects of adenylyl cyclase a deficient Dictyostelium discoideum cells.
BMC Dev. Biol.
12
,
3
.
Swiston
J.
,
Hubberstey
A.
,
Yu
G.
,
Young
D.
(
1995
).
Differential expression of CAP and CAP2 in adult rat tissues.
Gene
165
,
273
277
.
Terami
H.
,
Hidaka
K.
,
Shirai
M.
,
Narumiya
H.
,
Kuroyanagi
T.
,
Arai
Y.
,
Aburatani
H.
,
Morisaki
T.
(
2007
).
Efficient capture of cardiogenesis-associated genes expressed in ES cells.
Biochem. Biophys. Res. Commun.
355
,
47
53
.
Thompson
J. D.
,
Higgins
D. G.
,
Gibson
T. J.
(
1994
).
CLUSTAL W: improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice.
Nucleic Acids Res.
22
,
4673
4680
.
Vartiainen
M. K.
,
Mustonen
T.
,
Mattila
P. K.
,
Ojala
P. J.
,
Thesleff
I.
,
Partanen
J.
,
Lappalainen
P.
(
2002
).
The three mouse actin-depolymerizing factor/cofilins evolved to fulfill cell-type-specific requirements for actin dynamics.
Mol. Biol. Cell
13
,
183
194
.
Vojtek
A.
,
Haarer
B.
,
Field
J.
,
Gerst
J.
,
Pollard
T. D.
,
Brown
S.
,
Wigler
M.
(
1991
).
Evidence for a functional link between profilin and CAP in the yeast S. cerevisiae.
Cell
66
,
497
505
.
Waddle
J. A.
,
Cooper
J. A.
,
Waterston
R. H.
(
1993
).
The alpha and beta subunits of nematode actin capping protein function in yeast.
Mol. Biol. Cell
4
,
907
917
.
Wang
C.
,
Zhou
G. L.
,
Vedantam
S.
,
Li
P.
,
Field
J.
(
2008
).
Mitochondrial shuttling of CAP1 promotes actin- and cofilin-dependent apoptosis.
J. Cell Sci.
121
,
2913
2920
.
Wang
Y.
,
Zou
H.
,
Fang
H. M.
,
Zhu
Y.
(
2010
).
Linking cellular actin status with cAMP signaling in Candida albicans.
Virulence
1
,
202
205
.
Wen
K. K.
,
McKane
M.
,
Houtman
J. C.
,
Rubenstein
P. A.
(
2008
).
Control of the ability of profilin to bind and facilitate nucleotide exchange from G-actin.
J. Biol. Chem.
283
,
9444
9453
.
Wills
Z.
,
Emerson
M.
,
Rusch
J.
,
Bikoff
J.
,
Baum
B.
,
Perrimon
N.
,
Van Vactor
D.
(
2002
).
A Drosophila homolog of cyclase-associated proteins collaborates with the Abl tyrosine kinase to control midline axon pathfinding.
Neuron
36
,
611
622
.
Wolanski
M.
,
Khosrowshahian
F.
,
Jerant
L.
,
Jap
I. S.
,
Brockman
J.
,
Crawford
M. J.
(
2009
).
Expression of CAP2 during early Xenopus embryogenesis.
Int. J. Dev. Biol.
53
,
1063
1067
.
Woods
N. T.
,
Mesquita
R. D.
,
Sweet
M.
,
Carvalho
M. A.
,
Li
X.
,
Liu
Y.
,
Nguyen
H.
,
Thomas
C. E.
,
Iversen
E. S.
 Jr
,
Marsillac
S.
 et al. (
2012
).
Charting the landscape of tandem BRCT domain-mediated protein interactions.
Sci. Signal.
5
,
rs6
.
Yamazaki
K.
,
Takamura
M.
,
Masugi
Y.
,
Mori
T.
,
Du
W.
,
Hibi
T.
,
Hiraoka
N.
,
Ohta
T.
,
Ohki
M.
,
Hirohashi
S.
 et al. (
2009
).
Adenylate cyclase-associated protein 1 overexpressed in pancreatic cancers is involved in cancer cell motility.
Lab. Invest.
89
,
425
432
.
Yang
S.
,
Cope
M. J.
,
Drubin
D. G.
(
1999
).
Sla2p is associated with the yeast cortical actin cytoskeleton via redundant localization signals.
Mol. Biol. Cell
10
,
2265
2283
.
Yonezawa
N.
,
Nishida
E.
,
Sakai
H.
(
1985
).
pH control of actin polymerization by cofilin.
J. Biol. Chem.
260
,
14410
14412
.
Yu
G.
,
Swiston
J.
,
Young
D.
(
1994
).
Comparison of human CAP and CAP2, homologs of the yeast adenylyl cyclase-associated proteins.
J. Cell Sci.
107
,
1671
1678
.
Yu
J.
,
Wang
C.
,
Palmieri
S. J.
,
Haarer
B. K.
,
Field
J.
(
1999
).
A cytoskeletal localizing domain in the cyclase-associated protein, CAP/Srv2p, regulates access to a distant SH3-binding site.
J. Biol. Chem.
274
,
19985
19991
.
Yusof
A. M.
,
Hu
N. J.
,
Wlodawer
A.
,
Hofmann
A.
(
2005
).
Structural evidence for variable oligomerization of the N-terminal domain of cyclase-associated protein (CAP).
Proteins
58
,
255
262
.
Yusof
A. M.
,
Jaenicke
E.
,
Pedersen
J. S.
,
Noegel
A. A.
,
Schleicher
M.
,
Hofmann
A.
(
2006
).
Mechanism of oligomerisation of cyclase-associated protein from Dictyostelium discoideum in solution.
J. Mol. Biol.
362
,
1072
1081
.
Zelicof
A.
,
Gatica
J.
,
Gerst
J. E.
(
1993
).
Molecular cloning and characterization of a rat homolog of CAP, the adenylyl cyclase-associated protein from Saccharomyces cerevisiae.
J. Biol. Chem.
268
,
13448
13453
.
Zelicof
A.
,
Protopopov
V.
,
David
D.
,
Lin
X. Y.
,
Lustgarten
V.
,
Gerst
J. E.
(
1996
).
Two separate functions are encoded by the carboxyl-terminal domains of the yeast cyclase-associated protein and its mammalian homologs. Dimerization and actin binding.
J. Biol. Chem.
271
,
18243
18252
.
Zhou
G. L.
,
Miyazaki
Y.
,
Nakagawa
T.
,
Tanaka
K.
,
Shishido
K.
,
Matsuda
H.
,
Kawamukai
M.
(
1998
).
Identification of a CAP (adenylyl-cyclase-associated protein) homologous gene in Lentinus edodes and its functional complementation of yeast CAP mutants.
Microbiology
144
,
1085
1093
.
Zhou
X.
,
Zhang
H.
,
Li
G.
,
Shaw
B.
,
Xu
J. R.
(
2012
).
The Cyclase-associated protein Cap1 is important for proper regulation of infection-related morphogenesis in Magnaporthe oryzae.
PLoS Pathog.
8
,
e1002911
.
Zou
H.
,
Fang
H. M.
,
Zhu
Y.
,
Wang
Y.
(
2010
).
Candida albicans Cyr1, Cap1 and G-actin form a sensor/effector apparatus for activating cAMP synthesis in hyphal growth.
Mol. Microbiol.
75
,
579
591
.

Supplementary information