EFA6A encodes two isoforms with distinct biological activities in neuronal cells

11
2108 Research Article Introduction The regulation of GTPases forms the basis of a wide variety of cellular processes, such as differentiation, migration and proliferation. The ADP-ribosylation factor (ARF) family of GTPases comprises six polypeptides (ARF1-ARF6) that are involved in membrane trafficking and maintenance of organelle structure (D’Souza-Schorey and Chavrier, 2006). The ARF6-signaling pathway has a role in endosomal-plasma membrane recycling, and in cortical actin cytoskeleton remodeling (D’Souza-Schorey and Chavrier, 2006), suggesting that its activity could coordinate membrane and cytoskeleton dynamics. ARF6 has also been demonstrated to be involved in regulated exocytosis (Caumont et al., 1998; Yang and Mueckler, 1999), in clathrin-dependent and - independent endocytosis (Naslavsky et al., 2003), in phagocytosis (Niedergang et al., 2003), in cytokinesis (Schweitzer and D’Souza- Schorey, 2002), in coordinating the transition of epithelial cells from a stationary to a motile state (Santy and Casanova, 2001; Turner and Brown, 2001) and in tumor cell invasion (Hashimoto et al., 2004; Tague et al., 2004). ARF6 is a ubiquitously expressed GTPase (Yang et al., 1998). Several molecules that appear to act in vitro as GEFs or GAPs for ARF6 have been identified, indicating that the regulation of ARF6 activity is finely tuned (Randazzo and Hirsch, 2004; Shin and Nakayama, 2004). Distinct GEFs and GAPs could regulate the diverse activities of ARF6 by localizing at specific subcellular compartments and mediating ARF6 regulation at distinct sites (D’Souza-Schorey and Chavrier, 2006; Shin and Nakayama, 2004). Localization of the ARF6 GEFs and GAPs could be under the control of diverse signals (upon binding of different phosphoinositides via the PH domains, or via protein-protein interactions) and in turn might regulate different biological processes via the recruitment of specific effectors. ARF6 GEFs include the EFA6 family and the ARNO/cytohesin family (Cox et al., 2004; D’Souza-Schorey and Chavrier, 2006); these molecules are characterized by the presence of a Sec7 domain, mediating the GEF activity, a PH domain and a coiled- coil region. The EFA6 gene family includes four members: EFA6A, EFA6B, EFA6C and EFA6D (Derrien et al., 2002; Franco et al., 1999; Matsuya et al., 2005; Sakagami et al., 2006; Sakagami, 2008). Among the putative ARF6 GEFs, some, such as ARNO (Kim et al., 1998), EFA6B (Derrien et al., 2002) and EFA6D (Sakagami et al., 2006), are widely expressed, whereas EFA6A and EFA6C appear to be tissue specific, and are both highly expressed in the brain (Matsuya et al., 2005; Perletti et al., 1997; Suzuki et al., 2002), underlying a fundamental role for ARF6 regulation in neural tissue (Jaworski, 2007). The process of neurite extension and remodeling requires a close coordination between the cytoskeleton and the cell membranes, to allow polarized growth of neuronal processes (Bretscher and Aguado-Velasco, 1998; da Silva and Dotti, 2002). Cytoskeletal remodeling during neuronal development underlies axonal and dendritic extension and branching, apical dendrite development, spine formation and synaptic plasticity. ARF6 is expressed in developing and adult rat brain (Choi et al., 2006; Suzuki et al., 2002). The ARF6 GEF ARNO has been shown The processes of neurite extension and remodeling require a close coordination between the cytoskeleton and the cell membranes. The small GTPase ARF6 (ADP-ribosylation factor 6) has a central role in regulating membrane traffic and actin dynamics, and its activity has been demonstrated to be involved in neurite elaboration. EFA6A has been shown to act as a guanine nucleotide exchange factor (GEF) for ARF6. Here, we report that two distinct isoforms of the EFA6A gene are expressed in murine neural tissue: a long isoform of 1025 amino acids (EFA6A), and a short isoform of 393 amino acids (EFA6As). EFA6A encompasses proline-rich regions, a Sec7 domain (mediating GEF activity on ARF6), a PH domain, and a C-terminal region with coiled-coil motifs. EFA6As lacks the Sec7 domain, and it comprises the PH domain and the C- terminal region. The transcript encoding EFA6As is the result of alternative promoter usage. EFA6A and EFA6As have distinct biological activities: upon overexpression in HeLa cells, EFA6A induces membrane ruffles, whereas EFA6As gives rise to cell elongation; in primary cortical neurons EFA6A promotes neurite extension, whereas EFA6As induces dendrite branching. Our findings suggest that EFA6A could participate in neuronal morphogenesis through the regulated expression of two functionally distinct isoforms. Supplementary material available online at http://jcs.biologists.org/cgi/content/full/122/12/2108/DC1 Key words: EFA6A, ARF6 GEF, Cytoskeletal remodeling, Neurons, Neurite extension, Dendrite branching, Rho-GTPases Summary EFA6A encodes two isoforms with distinct biological activities in neuronal cells Cristina Sironi 1 , Tambet Teesalu 2 , Anna Muggia 1 , Gabriele Fontana 1 , Fortunata Marino 1 , Sara Savaresi 3 and Daniela Talarico 1, * 1 Division of Genetics and Cell Biology, San Raffaele Scientific Institute, Milan, Italy 2 Vascular Mapping Center, Burnham Institute for Medical Research at University of California Santa Barbara, Santa Barbara, CA 93106, USA 3 Department of Biochemistry, University of Zürich, Winterthurer Strasse 190, 8057 Zürich, CH, Switzerland *Author for correspondence (e-mail: [email protected]) Accepted 3 March 2009 Journal of Cell Science 122, 2108-2118 Published by The Company of Biologists 2009 doi:10.1242/jcs.042325 Journal of Cell Science

Transcript of EFA6A encodes two isoforms with distinct biological activities in neuronal cells

2108 Research Article

IntroductionThe regulation of GTPases forms the basis of a wide variety ofcellular processes, such as differentiation, migration andproliferation. The ADP-ribosylation factor (ARF) family of GTPasescomprises six polypeptides (ARF1-ARF6) that are involved inmembrane trafficking and maintenance of organelle structure(D’Souza-Schorey and Chavrier, 2006). The ARF6-signalingpathway has a role in endosomal-plasma membrane recycling, andin cortical actin cytoskeleton remodeling (D’Souza-Schorey andChavrier, 2006), suggesting that its activity could coordinatemembrane and cytoskeleton dynamics. ARF6 has also beendemonstrated to be involved in regulated exocytosis (Caumont etal., 1998; Yang and Mueckler, 1999), in clathrin-dependent and -independent endocytosis (Naslavsky et al., 2003), in phagocytosis(Niedergang et al., 2003), in cytokinesis (Schweitzer and D’Souza-Schorey, 2002), in coordinating the transition of epithelial cells froma stationary to a motile state (Santy and Casanova, 2001; Turnerand Brown, 2001) and in tumor cell invasion (Hashimoto et al.,2004; Tague et al., 2004).

ARF6 is a ubiquitously expressed GTPase (Yang et al., 1998).Several molecules that appear to act in vitro as GEFs or GAPs forARF6 have been identified, indicating that the regulation of ARF6activity is finely tuned (Randazzo and Hirsch, 2004; Shin andNakayama, 2004). Distinct GEFs and GAPs could regulate the diverseactivities of ARF6 by localizing at specific subcellular compartmentsand mediating ARF6 regulation at distinct sites (D’Souza-Schoreyand Chavrier, 2006; Shin and Nakayama, 2004). Localization of the

ARF6 GEFs and GAPs could be under the control of diverse signals(upon binding of different phosphoinositides via the PH domains, orvia protein-protein interactions) and in turn might regulate differentbiological processes via the recruitment of specific effectors.

ARF6 GEFs include the EFA6 family and the ARNO/cytohesinfamily (Cox et al., 2004; D’Souza-Schorey and Chavrier, 2006);these molecules are characterized by the presence of a Sec7domain, mediating the GEF activity, a PH domain and a coiled-coil region. The EFA6 gene family includes four members: EFA6A,EFA6B, EFA6C and EFA6D (Derrien et al., 2002; Franco et al., 1999;Matsuya et al., 2005; Sakagami et al., 2006; Sakagami, 2008).Among the putative ARF6 GEFs, some, such as ARNO (Kim etal., 1998), EFA6B (Derrien et al., 2002) and EFA6D (Sakagami etal., 2006), are widely expressed, whereas EFA6A and EFA6C appearto be tissue specific, and are both highly expressed in the brain(Matsuya et al., 2005; Perletti et al., 1997; Suzuki et al., 2002),underlying a fundamental role for ARF6 regulation in neural tissue(Jaworski, 2007).

The process of neurite extension and remodeling requires a closecoordination between the cytoskeleton and the cell membranes, toallow polarized growth of neuronal processes (Bretscher andAguado-Velasco, 1998; da Silva and Dotti, 2002). Cytoskeletalremodeling during neuronal development underlies axonal anddendritic extension and branching, apical dendrite development,spine formation and synaptic plasticity.

ARF6 is expressed in developing and adult rat brain (Choi et al.,2006; Suzuki et al., 2002). The ARF6 GEF ARNO has been shown

The processes of neurite extension and remodeling require aclose coordination between the cytoskeleton and the cellmembranes. The small GTPase ARF6 (ADP-ribosylation factor6) has a central role in regulating membrane traffic and actindynamics, and its activity has been demonstrated to be involvedin neurite elaboration. EFA6A has been shown to act as aguanine nucleotide exchange factor (GEF) for ARF6. Here, wereport that two distinct isoforms of the EFA6A gene areexpressed in murine neural tissue: a long isoform of 1025 aminoacids (EFA6A), and a short isoform of 393 amino acids(EFA6As). EFA6A encompasses proline-rich regions, a Sec7domain (mediating GEF activity on ARF6), a PH domain, anda C-terminal region with coiled-coil motifs. EFA6As lacks theSec7 domain, and it comprises the PH domain and the C-terminal region. The transcript encoding EFA6As is the result

of alternative promoter usage. EFA6A and EFA6As have distinctbiological activities: upon overexpression in HeLa cells, EFA6Ainduces membrane ruffles, whereas EFA6As gives rise to cellelongation; in primary cortical neurons EFA6A promotesneurite extension, whereas EFA6As induces dendrite branching.Our findings suggest that EFA6A could participate in neuronalmorphogenesis through the regulated expression of twofunctionally distinct isoforms.

Supplementary material available online athttp://jcs.biologists.org/cgi/content/full/122/12/2108/DC1

Key words: EFA6A, ARF6 GEF, Cytoskeletal remodeling, Neurons,Neurite extension, Dendrite branching, Rho-GTPases

Summary

EFA6A encodes two isoforms with distinct biologicalactivities in neuronal cellsCristina Sironi1, Tambet Teesalu2, Anna Muggia1, Gabriele Fontana1, Fortunata Marino1, Sara Savaresi3 andDaniela Talarico1,*1Division of Genetics and Cell Biology, San Raffaele Scientific Institute, Milan, Italy2Vascular Mapping Center, Burnham Institute for Medical Research at University of California Santa Barbara, Santa Barbara, CA 93106, USA3Department of Biochemistry, University of Zürich, Winterthurer Strasse 190, 8057 Zürich, CH, Switzerland*Author for correspondence (e-mail: [email protected])

Accepted 3 March 2009Journal of Cell Science 122, 2108-2118 Published by The Company of Biologists 2009doi:10.1242/jcs.042325

Jour

nal o

f Cel

l Sci

ence

2109Activity of EFA6A isoforms in neurons

to regulate dendritic and axonal development in hippocampalneurons (Hernandez-Deviez et al., 2002; Hernandez-Deviez et al.,2004). Also, ARF6 has been shown to be involved in neuriteextension in chicken retinal neurons (Albertinazzi et al., 2003), andin regulating spine formation (Choi et al., 2006; Miyazaki et al.,2005).

A role for EFA6A in neuronal morphogenesis is suggested bythe finding that a GEF-defective EFA6A mutant enhances dendriteformation (Sakagami et al., 2004), and the observation that EFA6Aoverexpression promotes spine formation in hippocampal neurons(Choi et al., 2006). EFA6A is expressed in neural tissue duringdevelopment, and in adult brain EFA6A mRNAs are detected in thecerebral cortex, the hippocampus and the dentate gyrus (Sakagamiet al., 2004; Suzuki et al., 2002). Northern analysis has shown twodistinct transcripts, of about 4 kb and 2 kb, respectively, both inhuman (Perletti et al., 1997) and rat (Suzuki et al., 2002) brain RNA.Both transcripts are expressed in neural tissue during embryogenesisand in postnatal life (Suzuki et al., 2002).

In the current study, we isolated and characterized the murinecDNAs corresponding to the two EFA6A mRNA transcripts, andanalyzed their functional activity in HeLa cells and in neuronal cells.Our findings indicate that the two isoforms encoded by EFA6A havedistinct effects on cell morphogenesis, and suggest that theirregulated expression in neuronal cells might have a role in theprocess of neurite elaboration.

ResultsCloning of EFA6A and EFA6AsTo isolate murine EFA6A, we screened mouse cDNA libraries usingthe 3� region of human EFA6A (Perletti et al., 1997) as a probe,followed by the RACE technique to extend 5� the cloned sequences.Through this approach, a 3785 bp cDNA was isolated. The murineEFA6A mRNA has the potential to encode a 1025 amino acidpolypeptide, encompassing two proline-rich sequences, a Sec7 anda PH domain, and two putative coiled-coil motifs in the C-terminalregion (identified by the COILS algorithm) (Lupas, 1996) (Fig. 1A;supplementary material Fig. S1A). The initiation methionine codonis preceded by an in-frame stop codon. A murine cDNA clone of3950 bp (Accession no. NM_028627) is almost identical to themurine EFA6A cDNA that we cloned (no mismatches in 3782 bpoverlap). Our cDNA only diverged from this sequence for a 3 bpinsertion in position 1645, which results in the insertion of a serinein position 519 of the encoded protein. As this occurs at a site ofsplicing, and the three-nucleotide insertion is also found in severalESTs, it is likely to be a splicing variant. A murine EFA6A cDNAidentical to the NM_028627 GenBank sequence, encoding a 1024amino acid polypeptide has been recently reported by Sakagami etal. (Sakagami et al., 2007).

By screening a day 11.5 mouse embryo cDNA library we alsoisolated a different 1850 bp cDNA, which appeared to be identicalto the EFA6A cDNA in the 3� region, up to nucleotide 2185, andthen diverged. This cDNA has the potential to encode a 393 aminoacid polypeptide. The putative methionine initiation codon is notpreceded by a stop codon; we have been unable to isolate clonesextending 5� by RACE experiments. The nucleotide sequencesurrounding the predicted initiating methionine conforms to theKozak consensus for efficient translation (Kozak, 1987). Theencoded polypeptide has an N-terminal unique stretch of 66 aminoacids (supplementary material Fig. S1B), followed by 327 aminoacids identical to EFA6A, encompassing the PH domain and thecoiled-coil regions (Fig. 1A; supplementary material Fig. S1A). We

named this polypeptide EFA6As (for short EFA6A). EFA6As lacksthe Sec7 domain, endowed with the GEF activity on ARF6; it couldact as a scaffolding protein, tethering interacting proteins via itscoiled-coil motifs.

The sequence encoding EFA6As appears to be the result oftranscription from an alternative internal promoter. The EFA6AcDNA is encoded by 16 exons; the first exon of EFA6As, encodingthe 66 N-terminal amino acids, is contained in a large 3.4 kb intron,

Fig. 1. Constructs used for transfection, expression of EFA6A transcripts inadult murine tissues and at different stages of brain development, andexpression of EFA6A and EFA6As polypeptides in mouse brain.(A) Schematic representation of EFA6A, ΔC, GEF– and 375 EFA6A mutants,and EFA6As. P, proline-rich regions; SEC7, Sec7 domain; PH, PH domain;CC, coiled-coil motifs. (B) Northern analysis of total RNA from tissues of 6-week-old CD-1 mice. (C) Northern analysis of RNA from embryonic (E14.5and E18.5), postnatal (P1, P4 and P8) and adult brain. Kidney, RNA from adultkidney. Samples (15 μg of total RNA per lane) were fractionated on 1%agarose-formaldehyde gel, and transferred to Hybond N. After transfer, thefilter was hybridized with a 32P-labeled 520 bp fragment (bp 2722-3242 ofEFA6A cDNA) (upper panels). Lower panels, ethidium bromide staining of thegels. (D) Protein extracts (50 μg/lane) were loaded on 8% SDS-polyacrylamidegels, transferred to PVDF membrane, and incubated with a polyclonal anti-EFA6A antibody, raised against a 13 amino acid sequence of the C-terminalregion, identical in EFA6A and EFA6As. Liver, kidney, brain: protein extractsfrom 6-week-old CD-1 mouse tissues; EGFP, EGFP-transfected HeLa cells;EFA6A, EFA6A-transfected HeLa cells; EFA6As, EFA6As-transfected HeLacells. Arrow indicates EFA6A and arrowhead EFA6As. The high molecularmass bands detected in kidney and in liver disappear when the antibody is pre-incubated with the peptide used for the immunization. As the transcriptencoding EFA6A is not expressed liver or kidney by Northern or PCR analysis(Sakagami et al., 2006) (and this study), they probably represent proteins thatshare homology with the immunizing peptide.

Jour

nal o

f Cel

l Sci

ence

2110

between exons 9 and 10 of EFA6A (supplementary material Fig.S1C). The region 5� to the start of the EFA6As sequence containsa CpG island; in this region we have identified a 200 bp fragmentthat displays promoter activity in neuronal cells (G.F., unpublished).

The identification of two distinct EFA6A cDNA sequences,identical in the 3� region, is in line with the two distinct transcripts,

Journal of Cell Science 122 (12)

of about 4 kb and 2 kb, detected by Northern analysis, uponhybridization of human (Perletti et al., 1997), rat (Suzuki et al.,2002) or murine (Fig. 1B,C) brain RNA with probes encompassingthe 3� region of EFA6A.

Using BLAST analysis, we identified a human fetal brain cDNAthat has the potential to encode the EFA6A short form (Accessionno. CR616163.1), as well as several human ESTs encompassingthe 5� region of EFA6As cDNA. In addition, a cDNA clone of 1832bp from murine visual cortex (Accession no. AK158851.1), identicalto the sequence that we isolated, is present in GenBank. Interestingly,searching GenBank with the N-terminal amino acid sequence ofEFA6As, several chicken, fish and Xenopus EST sequencesencoding highly homologous polypeptides were identified,indicating that the transcript encoding EFA6As appears to be highlyconserved in vertebrates.

mRNAs encoding EFA6A and EFA6As are expressed inmouse brain and show a distinct temporal regulation duringdevelopmentWe performed a Northern analysis of total RNA from adult murinetissues, using as a probe a 520 bp fragment, common to cDNAsencoding EFA6A and EFA6As (encompassing the sequenceencoding 118 amino acids of the C-terminal region and 166nucleotides of the 3� untranslated region). Two distinct transcripts,of about 4 kb and 2 kb, respectively, were detected in brain RNA,whereas stomach, intestine, uterus and ovary mainly expressed the2 kb transcript (Fig. 1B). No specific signal was detectable in tongue,thymus, spleen, lung, heart, liver and kidney. These results are inline with those observed in the analysis of RNA from adult humantissues using a probe from human EFA6A (Perletti et al., 1997).Both transcripts were detected in embryonic, early postnatal andadult mouse brain (Fig. 1C). The mRNA transcript encodingEFA6A peaked at postnatal day 4-8, and it was downregulated inthe adult, whereas the highest level of the mRNA encoding EFA6Aswas detected in adult brain. These data closely parallel what wasobserved in rat brain (Sakagami et al., 2004; Suzuki et al., 2002).

The distinct temporal regulation of the levels of mRNAs encodingEFA6A and EFA6As suggests that the two polypeptides mightperform different functions in neuronal cells.

Both EFA6A and EFA6As proteins are expressed in mousebrainTo analyze the expression of endogenous EFA6A and EFA6As inmurine tissues, we raised two distinct polyclonal antibodies, by

Fig. 2. Phenotype of HeLa cells transfected with EFA6A, 375, ΔC andEFA6As. HeLa cells were plated on collagen-coated coverslips, andtransfected by pEi with expression plasmids encoding farnesylated-EGFP(A,B), EFA6A (C-F), 375 (G,H), ΔC (I,J), EFA6As (K,L). After 20-24 hours,cells were fixed and processed for immunofluorescence with anti-EFA6A(C,E,G,K), or anti-FLAG (I) polyclonal antibodies followed by an Alexa Fluor594-conjugated goat anti-rabbit secondary antibody, and stained with AlexaFluor 488-conjugated phalloidin (B,D,F,H,J,L). Cells overexpressing EFA6Adisplay membrane ruffles, where EFA6A colocalizes with F-actin, and loss ofstress fibers (C-F); a similar phenotype is observed in cells transfected with the375 construct (G,H); cells transfected with the ΔC mutant do not showsignificant morphological changes (I,J). Overexpression of EFA6As inducescell elongation (K,L). Scale bar: 20 μm. (M) Quantification of the fraction oftransfected cells displaying a substantial (over 90%) loss of stress fibers.***P<0.001, one-way ANOVA. (N) Quantification of transfected cellsshowing an elongated morphology (cells were scored as elongated when theirlength was three times or more the width). Results are means ± s.d.***P<0.001, one-way ANOVA.

Jour

nal o

f Cel

l Sci

ence

2111Activity of EFA6A isoforms in neurons

immunizing rabbits against a histidine-tagged fusion proteinencompassing the C-terminal 155 amino acids, or against a 13 aminoacid peptide corresponding to amino acids 988-1000 of murineEFA6A. By western blot analysis, both antibodies appeared to detectspecific bands in extracts from murine brain tissues (Fig. 1D), withapparent molecular masses of about 120 kDa and 45 kDa,comigrating with EFA6A and EFA6As overexpressed in transfectedHeLa cells. These same bands were not present in extracts fromtissues that do not express EFA6A, such as kidney or liver (Fig. 1D).

We also attempted to generate anti-EFA6As polyclonalantibodies, directed against the N-terminal 66 amino acids uniqueto EFA6As; however, the antibodies obtained detected theoverexpressed EFA6As in extracts of transfected cells, but not theendogenously expressed protein. The lack of success in generatinghigh-titer specific antibodies is probably the result of the remarkablesequence conservation of the 66 N-terminal amino acids acrossspecies (i.e. 100% identity between the murine and the humansequence; 86% identity between the murine and the chickensequence).

Murine EFA6A and EFA6As induce cytoskeletal remodeling inHeLa cellsWe tested the biological activity of EFA6A and EFA6As in HeLacells, by transfection of the cDNAs cloned in the pcDNA3

expression vector, followed by immunofluorescence analysis. Cellstransfected with EFA6A displayed membrane ruffles at the cellperiphery, and a substantial loss of stress fibers (Fig. 2C-F,M),compared with untransfected or EGFP-transfected cells (Fig.2A,B,M); EFA6A was detected mostly at the cell membrane, andit colocalized with F-actin at membrane ruffles. The biologicalactivity of EFA6A appears to require the C-terminal region ofEFA6A, because a ΔC mutant (Fig. 1A), encompassing the first883 amino acids of EFA6A, but lacking the 142 C-terminalresidues, does not induce morphological changes or loss of stressfibers (Fig. 2I,J,M). This finding indicates that the C-terminal region,common to EFA6A and EFA6As, is essential for cytoskeletalremodeling in HeLa cells.

Transfection of HeLa cells with a plasmid encoding a polypeptidelacking the first 374 amino acids of murine EFA6A (construct 375)(Fig. 1A) induced a similar phenotype, indicating that the N-terminalregion appears to be dispensable for biological activity in transfectedHeLa cells (Fig. 2G,H,M).

Cells overexpressing EFA6As were characterized by an elongatedmorphology, and by an increase of the short microvilli-like actin-rich protrusions on the dorsal surface (Fig. 2K,L,N). The majorityof the protein appeared to be localized on the plasma membraneand on these protrusions; no substantial loss of stress fibers wasdetected in EFA6As-transfected cells (Fig. 2M). Deletion mutants

Fig. 3. EFA6A, but not EFA6As, induces neurite outgrowth in PC12 cells. PC12 cells were plated on collagen-coated coverslips, and transfected the next day withEGFP, EFA6A or EFA6As. After incubating for 3-4 hours with the DNA-pEi complexes, the medium was replaced with growth medium, or differentiationmedium, supplemented with 50 ng/ml NGF. 40 hours later, cells were fixed and processed for immunofluorescence. Cells were stained with polyclonal anti-FLAGantibodies, followed by Alexa Fluor 594-conjugated goat anti-rabbit secondary antibodies, and Alexa Fluor 488-conjugated phalloidin, to label filamentous actin.(A) Morphology of EGFP-transfected PC12 cells. Arrows indicate transfected cells. (B) Neurite outgrowth and actin-rich protrusions in EFA6A-transfected cells.NGF-treated EFA6A-overexpressing cells often display multiple neurites, branched and rich in filopodia. Arrows indicate transfected cells; in the F-actin column,the arrow at the top of the second image from the bottom highlights a neurite from a non-transfected cell, for comparison. (C) EFA6As overexpression does notinduce morphological changes in the absence of NGF; a fraction of NGF-treated EFA6As-overexpressing cells displays multiple short neurites. Arrows indicatetransfected cells. Scale bars: 20 μm.

Jour

nal o

f Cel

l Sci

ence

2112

of EFA6As lacking the C-terminal 142 amino acids did not induceelongation of transfected HeLa cells, indicating that biologicalactivity of EFA6As also requires the C-terminal region (data notshown).

Our results with HeLa cells indicate that both isoforms encodedby EFA6A induce cytoskeletal remodeling, but lead to distinctmorphological changes in transfected cells: EFA6A promotes theformation of membrane ruffles and the loss of stress fibers, whereasEFA6As induces cell elongation.

EFA6A induces neurite outgrowth in PC12 cellsTo assess the biological function of murine EFA6A and EFA6Asin neuronal cells, we transfected the EFA6A cDNA in ratpheochromocytoma PC12 cells. After 3-4 hours of incubation withthe DNA-pEi complexes, cells were incubated for 40 hours in theabsence or in the presence of NGF, fixed and analyzed byimmunofluorescence. Scoring of the morphology of transfectedcells, alongside with control EGFP-transfected cultures (Fig. 3A;Fig. 4A), showed that 13.6% of EFA6A-transfected cells displayedneurites, which were rich in filopodia (Fig. 3B; Fig. 4A). Moreover,the majority of EFA6A-transfected cells displayed numerous shortprotrusions, and increased spreading when compared with non-transfected cells (Fig. 3B).

In EFA6A-transfected cells exposed to NGF, the fraction ofneurite-bearing cells was about 25% (approximately equal to thesum of EFA6A-induced and NGF-elicited neurite outgrowth),suggesting that the NGF effect was additive, but not synergistic,with that of EFA6A (Fig. 4A). Neurites displayed a high densityof filopodia (Fig. 3B). NGF-treated EFA6A-overexpressing cellsoften showed multiple neurites (Fig. 3B); this phenotype wasobserved in 65% of the cells, whereas only 30% of the EGFP-transfected cells had more than one neurite. EFA6A activates ARF6,leading to activation of phosphatidylinositol(4)P 5-kinase andphospholipase D, and of the Rac GTPase (D’Souza-Schorey andChavrier, 2006). In addition, in melanoma cells, ARF6 has been

Journal of Cell Science 122 (12)

shown to increase Erk phosphorylation (Tague et al., 2004).Activation of the Ras-MAPK pathway and Rac activation have amajor role in NGF-dependent neurite extension in PC12 cells(Reichardt, 2006). Thus, a crosstalk between the NGF- and EFA6A-signaling pathways in neurite outgrowth in PC12 cells might takeplace.

The 375 mutant (deleted in the N-terminal region), and the ΔCmutant (lacking the 142 C-terminal amino acids) displayed reducedneurite-promoting activity (Fig. 4A).

To evaluate the role of the ARF6 pathway in neurite inductionby EFA6A, we generated a putative GEF-defective mutant, byreplacing the highly conserved Glu622 in motif 1 with Lys (EFA6A-E622K, represented as GEF–) (Fig. 1A). A mutation in this positionof the Sec7 domain of human EFA6A has been shown to impairGEF activity on ARF6 (Franco et al., 1999). Transfection of GEF–

EFA6A led to a loss of neurite induction by 60.2% (Fig. 4B). Tofurther assess the role of ARF6 in the biological activity of EFA6Ain neuronal cells, we cotransfected PC12 cells with a dominant-negative ARF6 mutant (ARF6T27N), and observed a decrease inneurite outgrowth similar to that observed with the GEF– mutant(Fig. 4B).

Therefore, the ARF6 pathway appeared to have a role in neuriteoutgrowth induced by EFA6A; however, interfering with the ARF6pathway did not completely block the biological activity of EFA6Ain neuronal cells. Similarly, a constitutively active ARF6 mutant(ARF6Q67L) was weak in inducing neurite outgrowth in PC12 cells,suggesting that further molecular events elicited by EFA6A have arole in the extension of neurites (Fig. 4B). The results observedwith the ΔC deletion mutant suggest that, besides the Sec7 domain(endowed with the GEF activity on ARF6), the C-terminal regionof the molecule might be engaged in molecular interactionscontributing to EFA6A-induced neurite outgrowth.

To explore the function of the Rho family GTPases, which areknown to have a fundamental role in the growth and remodelingof neuronal processes (Govek et al., 2005), we cotransfected

Fig. 4. Biological activity of EFA6A and mutants in PC12 cells, and role of the ARF6 and Rho-GTPase pathway on EFA6A activity. (A) EFA6A induces neuriteoutgrowth in transfected PC12 cells, whereas EFA6As does not promote neurite extension. Cells were transfected with EGFP, EFA6A, 375 and ΔC EFA6A deletionmutants, or EFA6As, and after 2 days of growth with or without NGF, cells were fixed and processed for immunofluorescence. Neurites were scored when theirlength was equal to or greater than twice the diameter of the cell body. *P<0.05; **P<0.01; ***P<0.001, one-way ANOVA. Data represent averages of severalindependent experiments ± s.d.; at least 50 cells per experiment were scored for each transfected construct. (B) Role of the exchange activity on ARF6 in EFA6A-induced neurite outgrowth in PC12 cells. Cells were transfected with EGFP, dominant-negative ARF6 (ARF6T27N), constitutively active ARF6 (ARFQ67L),EFA6A, EFA6A GEF– mutant, or cotransfected with EFA6A and ARF6T27N; after 2 days of growth with or without NGF, cells were processed forimmunofluorescence. Neurites were scored as described in A. ***P<0.001, one-way ANOVA. (C) Role of Rho-GTPases on EFA6A-induced neurite outgrowth inPC12 cells. Cells were transfected with EGFP, dominant-negative Cdc42 or Rac1, constitutively active RhoA, EFA6A, or cotransfected with EFA6A and dominant-negative Cdc42 or Rac1, or constitutively active RhoA. Neurites were scored as described in A. **P<0.01; ***P<0.001, one-way ANOVA.

Jour

nal o

f Cel

l Sci

ence

2113Activity of EFA6A isoforms in neurons

EGFP-tagged dominant-negative mutants of Cdc42 and Rac1 orconstitutively active RhoA, along with EFA6A. Dominant-negativeCdc42N17 or constitutively active RhoAV14 were able tosubstantially inhibit neurite outgrowth by EFA6A in PC12 cells,whereas coexpression of dominant-negative Rac1N17 resulted ina partial inhibition (64.4%) of EFA6A-induced biological activity(Fig. 4C). Transfection of EFA6As did not induce neurite outgrowth(Fig. 3C, Fig. 4A); in cells treated with NGF, multiple protrusions,resembling short neurites, were observed in about 30% of EFA6As-transfected PC12 cells (Fig. 3C).

Therefore, in PC12 cells, EFA6A overexpression promotesneurite extension, which is inhibited by interfering both with theARF6 as well as with the Rho-GTPase pathway, whereas EFA6Ashas a modest biological activity, inducing the formation ofprotrusions upon exposure to NGF.

Activity of EFA6A and EFA6As in embryonic rat corticalneuronsTo assess the role of EFA6A and EFA6As in primary neurons, E18.5rat cortices were dissociated, and cells were transfected after 2 daysin culture. Cultures were fixed 20-24 hours later, and themorphology of transfected neurons was evaluated afterimmunofluorescence staining. The majority of EGFP-transfectedcontrol cells displayed a pyramidal morphology, with a long apicaldendrite, and 4-5 short basal dendrites (Fig. 5A,B), whereas onlyabout 7% of the neurons had a nonpyramidal phenotype,characterized by 5-6 extended processes (Fig. 6A). By contrast, inEFA6A-transfected cells, the fraction of nonpyramidal neurons washighly increased, to 53% (Fig. 5C,D; Fig. 6A). The N-terminaldeletion 375 mutant also promoted a substantial increase ofnonpyramidal neurons, whereas only 17.8% of the neuronsexpressing the ΔC mutant, lacking the C-terminal 142 amino acids,displayed multiple extended processes.

To assess the involvement of ARF6 in EFA6A activity in primaryneurons, cells were transfected with the GEF– mutant orcotransfected with EFA6A and dominant-negative ARF6. Thefraction of nonpyramidal neurons was significantly reduced bothin cells transfected with the GEF– mutant, and upon cotransfectionwith EFA6A and dominant-negative ARF6, indicating a role forthe ARF6 pathway in the extension of processes (Fig. 6B).Moreover, in cells transfected with the GEF– mutant, orcotransfected with EFA6A and dominant-negative ARF6, increaseddendrite branching was observed (Fig. 6D). This is in line with thestudies of Hernandez-Davies and colleagues, who found increaseddendritic branching upon transfection of hippocampal neurons witha dominant-negative mutant of the ARF6-GEF ARNO (Hernandez-Deviez et al., 2004).

As Rho family GTPases have been shown to have a fundamentalrole in growth and remodeling of neuronal processes in primarycortical cultures (Threadgill et al., 1997), we assayed the effect ofcotransfecting dominant-negative Cdc42, Rac1 or constitutivelyactive RhoA on the phenotype induced by EFA6A. In cellscotransfected with EFA6A and RhoAV14, no increase in thefraction of nonpyramidal neurons was observed (Fig. 6C), indicatingthat low RhoA activity is required for processes extension inducedby EFA6A. A significant decrease in the fraction of nonpyramidalneurons was also observed upon cotransfection of Cdc42N17 orRac1N17 (Fig. 6C).

Transfection of EFA6As did not result in changes in the fractionof nonpyramidal neurons (Fig. 6A), whereas a significant increasein the branching of the basal dendrites was observed in transfected

cells (Fig. 5E,F; Fig. 6E). Cotransfection of Cdc42N17 or RhoAV14resulted in a drastic reduction of dendrite branching (Fig. 6E). Incells cotransfected with Rac1N17, dendrite branching induced byEFA6As was also significantly reduced (Fig. 6E).

Therefore, in primary neurons the two isoforms have distinctbiological activities: EFA6A promotes neurite extension, whichrequires both ARF6 and Rho-GTPase activity, whereas EFA6Asinduces dendrite branching that is significantly reduced byinterfering with the Rho-GTPase pathway. Representative imagesof cortical neurons transfected with the different constructs areshown in supplementary material Figs S2 and S3.

In line with the results of the overexpression studies, siRNAsdirected against EFA6A led to a significant increment in thefraction of transfected neurons displaying increased neuritebranching (Fig. 7A; supplementary material Fig. S4). Uponinterfering with the EFA6As transcript, an increase in the fractionof neurons with a nonpyramidal multipolar morphology wasobserved (Fig. 7B; supplementary material Fig. S4).

Biological effects of co-overexpression of EFA6A and EFA6Asin primary cortical neuronsWe addressed the biological effect of co-overexpression of EFA6Aand EFA6As in primary cortical neurons (supplementary material

Fig. 5. EFA6A induces neurite extension in transfected primary corticalneurons, whereas EFA6As promotes dendrite branching. Primary corticalneurons were plated on coverslips coated with poly-L-lysine and laminin,transfected the next day by pEi with farnesylated-EGFP (f-EGFP, to visualizemembranes), EFA6A or EFA6As, cultured for 20-24 hours, fixed andprocessed for immunofluorescence. Cells were stained with polyclonal anti-GFP (A), polyclonal anti-EFA6A (C,E) and monoclonal anti-MAP2 antibodies(B,D,F), followed by Alexa Fluor 488-conjugated goat anti-rabbit, and AlexaFluor 594-conjugated goat anti-mouse secondary antibodies. Arrows indicatetransfected cells. The insets in D and F show a high power view of thetransfected cells. Scale bar: 20 μm.

Jour

nal o

f Cel

l Sci

ence

2114

Fig. S5). The fraction of neurons cotransfected with both EFA6Aand EFA6As displaying a nonpyramidal morphology was lower thanthat observed in cells transfected with EFA6A (Fig. 8A); thisreduction appeared more substantial in cells transfected with a higherEFA6As:EFA6A ratio. These data indicate that, when expressed athigh levels, EFA6As could inhibit EFA6A biological activity.

However, in primary cortical neurons cotransfected with EFA6Aand EFA6As, a significant reduction of EFA6As-induced dendritebranching was observed, suggesting that coexpression of EFA6Acould interfere with the biological activity of the short isoform (Fig.8B). Together, these data suggest a possible role of the balancebetween EFA6A and EFA6As expression levels in the regulationof neuronal morphogenesis.

EFA6As does not modulate the GEF activity of ARF6To investigate whether the short isoform could have a role inmodulating the EFA6A GEF activity toward ARF6, we performedARF6 activation assays in transfected HeLa cells. Cells were

Journal of Cell Science 122 (12)

transfected with wild-type ARF6 and the N-terminal deletion 375construct (Fig. 1A) or EFA6A, and the effect of cotransfection ofEFA6As on the levels of ARF6-GTP was tested. As shown in Fig.8C, the levels of activated ARF6 were not significantly changedby coexpression of EFA6As, indicating that the short isoform doesnot modulate ARF6 activation elicited by the EFA6A Sec7 domain.

DiscussionIn this study we report that EFA6A encodes two polypeptides,EFA6A and EFA6As, which each have distinct biological activities.In HeLa cells, EFA6A overexpression gives rise to membrane rufflesand to loss of stress fibers, whereas overexpression of EFA6As, theshort isoform that we identified, induces cell elongation and actin-rich protrusions.

Induction of membrane ruffles and loss of stress fibers uponhuman EFA6A overexpression have been reported in previousstudies in epithelial CHO cells (Franco et al., 1999), as well as infibroblastic BHK cells (Derrien et al., 2002). The human construct

Fig. 6. Biological activity of EFA6A and mutants, role of the ARF6 and Rho-GTPase pathways, and analysis of dendritic branching in transfected cortical neurons.(A) Expression of EFA6A induces neurite outgrowth in transfected cortical neurons. Primary cortical rat neurons were transfected with EGFP, EFA6A, 375 and ΔCEFA6A deletion mutants, or EFA6As, and 1 day later fixed and processed for immunofluorescence. Transfected neurons were identified by costaining with anti-GFP or anti-EFA6A polyclonal antibodies and anti-MAP2 monoclonal antibody. Nonpyramidal neurons were scored when they displayed three or more processeswith length equal to or greater than five times the diameter of the cell body. Data are average of several independent experiments ± s.d.; 50 cells or more werescored for each experiment **P<0.01; ***P<0.001, one-way ANOVA. (B) Role of the exchange activity on ARF6 on EFA6A-induced neurite outgrowth intransfected cortical neurons. Primary cortical neurons were transfected with EGFP, constitutively active ARF6 (ARFQ67L), dominant-negative ARF6(ARF6T27N), EFA6A or EFA6A GEF– mutant, or cotransfected with EFA6A and dominant-negative ARF6. Nonpyramidal neurons were scored as described in A.***P<0.001, one-way ANOVA. (C) Role of Rho-GTPases on EFA6A-induced neurite outgrowth in transfected cortical neurons. Cells were transfected with EGFP,dominant-negative Cdc42 or Rac1, constitutively active RhoA, EFA6A, or cotransfected with EFA6A and dominant-negative Cdc42 or Rac1, or constitutivelyactive RhoA. Nonpyramidal neurons were scored as described in A. ***P<0.001, one-way ANOVA. (D) Expression of EFA6A GEF–, or cotransfection of EFA6Awith dominant-negative ARF6 (ARF6T27N) induces increased dendrite branching in cortical neurons. Primary cortical neurons were transfected with EGFP,EFA6A, GEF–, ARF6T27N or cotransfected with EFA6A and ARF6T27N, and 1 day later fixed and processed for immunofluorescence. Dendrite complexity wasevaluated by counting the number of dendritic tips; collaterals longer than half-cell diameter were counted as branches. **P<0.01, one-way ANOVA.(E) Expression of EFA6As induces increased dendrite branching in cortical neurons, which is reduced by cotransfection of dominant-negative Cdc42 or Rac1, orconstitutively active RhoA. ***P<0.001, one-way ANOVA.

Jour

nal o

f Cel

l Sci

ence

2115Activity of EFA6A isoforms in neurons

used in those studies lacks the 374 amino acids at the N-terminus,similarly to our 375 mutant, which also induces ruffles and loss ofstress fibers: these data together indicate that the N-terminal regiondoes not appear to significantly contribute to cytoskeletalremodeling induced by EFA6A.

In neuronal cells, the cell type that physiologically expresses theEFA6A-encoded polypeptides, overexpression of EFA6A inducesneurite extension, whereas EFA6As leads to increased dendritebranching in primary cortical neurons. This is the first reportdescribing the biological activity of wild-type EFA6A in neuronalcell lines and in primary cortical neurons at early stages in culture.We have observed that neurite extension by EFA6A is significantlydecreased by cotransfection of dominant-negative ARF6, or byimpairing the EFA6A GEF activity upon specific mutation of theSec7 domain.

Extension and elaboration of neuronal processes requirecoordination of directed membrane growth, modulation of adhesion,and dynamic changes in the cytoskeleton, biological activities thatthe ARF6-GTPase appears to modulate in a variety of cell systems(Donaldson, 2003). Several studies have indicated a role for ARF6in neurite development: ARF6 is expressed both in developing andin adult brain (Suzuki et al., 2001), and overexpression of dominant-negative mutants in neuronal cells affects neurite development(Albertinazzi et al., 2003; Hernandez-Deviez et al., 2002;Hernandez-Deviez et al., 2004). Moreover, recent studies indicatea role for ARF6 also in the regulation of dendritic spine formation(Miyazaki et al., 2005; Choi et al., 2006).

The GEF activity of EFA6A on ARF6 is not enough to efficientlyinduce extension of neuronal processes, in line with the observationthat constitutively active ARF6 does not promote neurite extensionin hippocampal or in retinal neurons (Hernandez-Deviez et al., 2002;Albertinazzi et al., 2003). Neurite extension by EFA6A requiresadditional biochemical interactions mediated by the C-terminalregion. The C-terminal region of EFA6A has been shown in vitroto interact with the K+ channel TWIK1 (Decressac et al., 2004),and in the recent study by Sakagami and co-workers, it has been

reported to interact with the actin-binding protein α-actinin(Sakagami et al., 2007).

The short isoform EFA6As that we cloned and characterizedpromotes increased branching of the basal dendrites, similarly tothat observed upon transfection of the EFA6A GEF– mutant(Sakagami et al., 2004) (and this study), suggesting that indeed themain functional difference between the two isoforms of EFA6Alies in the GEF activity displayed by the Sec7 domain. EFA6As islikely to act as a scaffold protein, and elicit its effects on cytoskeletaldynamics via the recruitment of interacting molecules.

The Rho-GTPase pathway, which has a fundamental role inregulating the morphogenesis of neuronal processes (Govek et al.,2005), appears to be involved in the biological activities of bothEFA6A and EFA6As, which are blocked by overexpression ofdominant-negative Cdc42 or constitutively active RhoA, andsignificantly reduced upon cotransfection of dominant-negativeRac1. The C-terminal region, common to EFA6A and EFA6As, isessential for their biological activity in transfected cells, and it isprobably involved in protein-protein interactions, which couldimpinge on the Rho-GTPase pathway.

Neurotrophins and other secreted factors, neuronal activity,adhesion molecules, changes in the actin and microtubulecytoskeleton all have an important role in the formation and/orstabilization of dendritic branches, a process that is essential forthe establishment of the neuronal circuitry (Jan and Jan, 2003).Increased dendritic (Hernandez-Deviez et al., 2002) and axonal(Hernandez-Deviez et al., 2004) branching has been observed uponexpression of catalytically inactive ARNO, or dominant-negativeARF6, suggesting a role for ARNO and ARF6 signaling innegatively regulating neurite branching. As EFA6As overexpressiondoes not appear to modulate the levels of active ARF6, EFA6As-promoted dendrite branching in primary cortical neurons couldinvolve negative modulation the ARF6-signaling pathway byinterfering with its effectors.

What is the functional significance of the two different isoformsencoded by EFA6A? The two isoforms of EFA6A appear to be

Fig. 7. Effect of siRNAs against EFA6A or EFA6As on neuronal morphology. (A) Primary cortical neurons were cotransfected with EGFP and control luciferasesiRNA (siLuc2) or siRNAs specific for EFA6A (siEFA6A1 and siEFA6A2) or EFA6As (siEFA6AsA and siEFA6AsB). 24 hours after transfection, cells were fixedand processed for immunofluorescence. Transfected neurons were identified by staining with an anti-GFP polyclonal antibody and an anti-MAP2 monoclonalantibody. Dendrite complexity was evaluated by counting the number of dendritic tips; collaterals longer than half-cell diameter were counted as branches. Data areaverage of several independent experiments ± s.d.; 50 cells or more were scored for each experiment. *P<0.05; **P<0.01, one-way ANOVA. (B) Nonpyramidalneurons were scored when they displayed three or more processes with length equal to or greater than five times the diameter of the cell body. **P<0.01, one-wayANOVA. (C) siRNAs reduce EFA6A and EFA6As protein levels in transfected HeLa cells. HeLa cells (7�105 cells per six-well plate) cotransfected withLipofectamine 2000 with the EFA6A expression vector and siRNAs siLuc2, siEFA6A1 or siEFA6A2, and with EFA6As and siRNAs siLuc2, siEFA6AsA, orsiEFA6AsB. 24 hours later, cells were lysed, and 40 μg of each sample were loaded on 8% SDS-polyacrylamide gel, transferred to PVDF membrane, and incubatedwith a polyclonal anti-EFA6A antibody. The histograms show the densitometric analysis of the reduction of EFA6A and EFA6As polypeptides in two independentexperiments.

Jour

nal o

f Cel

l Sci

ence

2116

conserved in evolution, because, in the NCBI database, ESTsencompassing the 5� region of EFA6As are also detected inchicken, fish and Xenopus. The transcripts encoding EFA6A andEFA6As are detected in developing and adult brain, with a distincttemporal regulation: mRNA encoding the EFA6A isoform peaksat postnatal day 4-8, whereas expression of the 2 kB transcript,encoding EFA6As, is highest in mature neural tissue (Sakagamiet al., 2004) (and our data). The expression data, along with thebiological activity of the two isoforms in neuronal cells, couldsuggest distinct functions for these proteins in the formation andremodeling of the neuronal processes. The role of EFA6A couldbe that of promoting processes elongation at earlier stages ofdevelopment, whereas EFA6As activity could contribute toelaboration and remodeling of the arbor at later embryonic andpostnatal stages of neural development. In adult life, EFA6Asmight be involved in formation and remodeling of dendritic spines.The expression of EFA6A and EFA6As appears to be driven bydifferent promoters (G.F., unpublished), and thereby the transcriptsencoding the two isoforms could be subjected to a distinctregulation by external stimuli.

We have observed that when EFA6A and EFA6As are co-overexpressed in transfected cortical neurons, at high levels ofEFA6As expression, neurite extension, and thereby increase in thefraction of nonpyramidal neurons promoted by EFA6A, is inhibited.Coexpression of EFA6As does not modulate the GEF activity ofthe EFA6A Sec7 domain on ARF6, suggesting that EFA6As doesnot act as a dominant-negative inhibitor of EFA6A by interferingwith its ability to activate the ARF6 GTPases. However, EFA6Asmight compete with EFA6A for binding to the same interactor/svia the C-terminal region, which is identical in the two isoforms.Indeed, dendrite branching promoted by EFA6As is significantlyreduced by coexpression of EFA6A, supporting the hypothesis thatthe two isoforms could compete for common effectors. The temporal

Journal of Cell Science 122 (12)

regulation of the expression levels of EFA6A and EFA6As inneuronal cells in vivo could contribute to the regulation ofneuromorphogenesis during development and differentiation. Owingto the significant sequence conservation of the C-terminal regionof the EFA6 family members (Derrien et al., 2002), it is also possiblethat EFA6As could modulate the biological activity of othercoexpressed EFA6 proteins.

The finding that the polypeptides encoded by the EFA6A mRNAtranscripts, which are regulated during neural development, areendowed with distinct biological activities in neuronal cells, supporta role for EFA6A in the regulation of neuronal morphogenesis.Further studies will clarify the molecular mechanisms underlyingthe function of EFA6A and EFA6As in neuronal cell arborization.

Materials and MethodsCloning and mutagenesisTo isolate murine EFA6A, we screened mouse embryo cDNA libraries using the 3�region of human EFA6A (Perletti et al., 1997) as a probe. Several positive cloneswere isolated and characterized. The longest clone isolated was of 2.3 kb, and lackedan initiation methionine codon and an in frame stop codon. For the isolation of thecomplete sequence of EFA6A cDNA (3785 bp), 5� RACE was performed usingspecific oligonucleotide primers and murine brain polyA+ RNA. Murine EFA6AscDNA was isolated from an E11.5 mouse embryo cDNA library upon hybridizationwith the 3� region of human EFA6A.

EFA6A and EFA6As cDNAs were tagged at the C-terminus with a FLAG epitopeby PCR, sequenced and cloned in the pcDNA3 expression vector; 375 and ΔC FLAG-tagged mutants were generated by PCR using specific primers, sequenced and clonedin pcDNA3; EFA6A-E622K (GEF–) was obtained mutagenizing the FLAG-taggedEFA6A cDNA by PCR using specific primers and the QuikChange site-directedmutagenesis kit from Stratagene.

Murine ARF6 was isolated by PCR from murine brain cDNA, Myc tagged at theC-terminus, and cloned in the pcDNA3 expression vector; to generate the ARF6T27Ndominant negative and ARF6Q67L constitutively active mutants, mutagenesis wasperformed with specific oligos using the QuikChange site-directed mutagenesis kitfrom Stratagene.

EGFP-tagged dominant-negative Rac1 and Cdc42 and constitutively active RhoAin the pCB6 expression vector were a kind gift from Michael Way (Cancer ResearchUK London Research Institute, London, UK).

Fig. 8. Biological effect of co-overexpression of EFA6A and EFA6As in primary cortical neurons, and lack of modulation of the GEF activity on ARF6 byEFA6As. (A) Coexpression of EFA6As significantly reduces the activity of EFA6A in promoting neurite extension. Primary cortical neurons were transfected withEGFP, EFA6A, EFA6As or cotransfected with EFA6A and EFA6As (EFA6A1/EFA6As0,5, respectively 1 and 0.5 μg of transfected plasmid; EFA6A1/EFA6As2,respectively 1 and 2 μg of transfected plasmid), and 1 day later they were fixed and processed for immunofluorescence. Nonpyramidal neurons were scored whenthey displayed three or more neurites with length equal to or greater than five times the diameter of the cell body. Data are average of several independentexperiments ± s.d.; 50 cells or more were scored for each experiment. *P<0.05; ***P<0.001, one-way ANOVA. (B) Coexpression of EFA6A significantly reducesthe activity of EFA6As in promoting dendrite branching. Primary cortical neurons were transfected with EGFP, EFA6A, EFA6As, or cotransfected with EFA6A andEFA6As (EFA6A1/EFA6As0,5, respectively 1 and 0.5 μg of transfected plasmid), and 1 day later they were fixed and processed for immunofluorescence. *P<0.05,one-way ANOVA. (C) HeLa cells transfected with the indicated constructs; 18-24 hours later, cells were lysed, and 600 μg of protein lysates were incubated with30 μg of the GST-ARHGAP10 fusion protein. Eluted proteins, as well as aliquots of the lysates (30 μg), were run on 10% polyacrylamide gels, transferred onPVDF, and incubated with monoclonal anti-Myc or polyclonal anti-EFA6A (anti-peptide) antibodies.

Jour

nal o

f Cel

l Sci

ence

2117Activity of EFA6A isoforms in neurons

Northern blot analysisTotal RNA from mouse tissues was purified using a standard guanidinium thiocyanate-phenol-chloroform extraction procedure. RNA was separated on 1% agarose gels andtransferred to Hybond N nylon membranes by capillary transfer; hybridization wasperformed using as a probe a 520 bp fragment of murine EFA6A (bp 2722-3242)labeled with 32P by random priming.

AntibodiesRabbit polyclonal anti-EFA6A antibodies were generated upon immunization with a13 amino acid peptide corresponding to amino acids 988-1000 of murine EFA6A orwith a histidine-tagged fusion protein encompassing the C-terminal 155 amino acids,followed by purification of IgGs by affinity chromatography on protein-A-Sepharosebeads.

Anti-FLAG polyclonal antibody and anti-MAP2 monoclonal antibody HM-2 werefrom Sigma, anti-Myc monoclonal antibody 9E10 was from Santa CruzBiotechnology, anti-GFP monoclonal antibody was from Roche. Alexa-Fluor-594 and-488 goat anti-rabbit and goat anti-mouse IgGs, Alexa-Fluor-488 and -594 phalloidinwere from Molecular Probes.

Western analysisTissues were dissected, snap frozen in liquid nitrogen, pulverized, lysed in RIPAbuffer [10 mM Tris-HCl (pH 7.4), 0.15 M NaCl, 1% sodium deoxycholate, 1% NP40,0.1% SDS, 1 mM EDTA, 10 mM KCl, protease inhibitors] or 50 mM Tris-HCl(pH 7.4)-2% SDS, centrifuged at 10,000 � g, and the protein concentration of thesupernatant was estimated. HeLa cells (4�105 cells/60 mm plate) were transfectedby pEi (pEi; average Mr ~25,000, Aldrich), with 10 μg EGFP, EFA6A or EFA6Asexpression plasmids, and lysed the next day in RIPA buffer. 25-50 μg proteinlysate/sample were adjusted into gel loading buffer (50 mM Tris-HCl, pH 6.8, 100 mMDTT, 2% sodium dodecyl sulfate, 10% glycerol, 0.1% bromophenol blue), heated at90°C for 5 minutes, separated on 8% SDS-polyacrylamide gels, transferred topolyvinylidene difluoride (PVDF). Membranes were stained with Ponceau red tocontrol for even loading, destained and processed according to standard procedures.The anti-EFA6A antibodies were diluted in 5% non-fat dry milk at 1 μg/ml.

Cells and transfectionHeLa cells were grown in DMEM supplemented with 10% fetal calf serum. Cells(2�104/well in 24-well plates) were plated on collagen-coated glass coverslips andtransfected the next day using polyethylenimine (pEi; average Mr ~25,000, Aldrich),with 1 μg plasmid DNA per well. After 3-4 hours of incubation with the DNA-pEicomplexes, fresh medium was added, and 20-24 hours later cells were fixed with 3%paraformaldehyde and processed for immunofluorescence. Loss of stress fibers andcell elongation were evaluated in several independent experiments; 50 transfectedcells per construct were scored for each experiment. Statistical analysis was performedby one-way ANOVA.

PC12 cells were grown in DMEM supplemented with 10% horse serum and 5%fetal calf serum. For differentiation, cells were grown in DMEM supplemented with2% horse serum, 1% fetal calf serum and 50 ng/ml NGF (Sigma). Cells (7�104/wellin 24-well plates) were plated on collagen-coated glass coverslips and transfectedthe next day by pEi, with 1-2 μg plasmid DNA per well. After 3-4 hours of incubationwith the DNA-pEi complexes, fresh medium (growth or differentiation medium) wasadded, and 40 hours later cells were fixed and processed for immunofluorescence.Neurites were scored when their length was at least double the diameter of the cellbody. Data were collected from several independent transfection experiments; at least50 transfected cells per construct were evaluated for each experiment; the results arepresented in percent ratios ± s.d. Statistical analysis was performed by one-wayANOVA followed by Tukey’s post-hoc test for multiple comparisons. Significancelevel was taken as P<0.05.

To establish cultures of primary cortical neurons, cortices from six to eight E18.5CD rat embryos were dissected, fragmented and processed as described (Treadgillet al., 1999). Cells were counted, and 1�106 cells/well in 24-well plates were platedon glass coverslips coated with laminin and poly-L-lysine in Eagle’s basal medium(BME) supplemented with 5% fetal calf serum and 1% N2 supplement (Gibco). Twodays later, cells were transfected by pEi (1-2 μg plasmid DNA per well, 2-3 hoursof incubation with the DNA-pEi complexes). Cells were fixed and processed forimmunofluorescence 20-24 hours later. Neuronal cells were identified by stainingwith anti-MAP2 monoclonal antibody. Processes were scored when their length wasfive times or more the diameter of the cell body. Data were collected from severalindependent transfection experiments; at least 50 transfected neurons per constructwere evaluated for each experiment; the results are presented in percent ratios ± s.d.Collaterals were scored as branches when their length was equal or more than halfthe diameter of the cell body. Statistical analysis was performed by one-way ANOVAfollowed by Tukey’s post-hoc test. Significance level was taken as P<0.05.

ImmunofluorescenceCells were fixed in 3% paraformaldehyde and 2% sucrose in PBS for 10 minutes atroom temperature, permeabilized in 0.2% Triton X-100, incubated with primaryantibodies diluted in PBS with 0.2% BSA for 1 hour at 37°C, washed three timeswith PBS-0.2% BSA and incubated in 2% BSA at 37°C for 15 minutes. Cells were

then incubated with fluorescently labeled secondary antibodies and/or fluorescentphalloidin diluted in PBS-0.2% BSA for 1 hour at 37°C, washed three times withPBS-0.2% BSA, stained with Hoechst 33258, and mounted on glass slides withMowiol 4-88 (Calbiochemical). Images were acquired using a Zeiss Axiophotmicroscope and a Hamamatsu C4742-95 digital camera, using the Hipic32 software.

RNA interferenceChemically synthesized, duplex siRNAs were purchased from Eurofins MWG. Thefollowing RNA interference sequences were used: siEFA6A1, 5�-ATTGGTGG -CTGGCGAGTAT-3� and siEFA6A2, 5�-GAACA ATG ACTTCAGC AAA-3� of ratEFA6A mRNA (GenBank accession no. XM_001066749); siEFA6AsA, 5�-GCA T -GATCGGCGTCAACAG-3� and siEFA6AsB, 5�-GCCGCCTGCAGAGCCGCAA-3� of rat EFA6As (AC 096363.8). As a control, the Luciferase GL2 siRNA, 5�-CGTACGCGGAATACTTCGA-3� (siLuc2) was used. Complementary smallinterfering RNA oligonucleotides were annealed, and 100 nM of the siRNAoligonucleotides were cotransfected with EFA6A or EFA6As expression plasmids inHeLa cells or in primary neurons by Lipofectamine 2000, to test the efficiency ofthe siRNAs in silencing EFA6A or EFA6s expression.

Primary cortical neurons (1.2�106 cells/well) were cotransfected with siRNAs andpEGFP (100 nM siRNA oligonucleotides, 50 ng pEGFP plasmid per well in 24-wellplates) by Lipofectamine 2000. Cells were processed for immunofluorescence withanti-GFP and anti-MAP2 antibodies 24 hours later. GFP-positive neurons (50 cellsper experiment) were scored for morphology.

ARF6 pull-down assayThe assay was performed following the protocol as described (Klein et al., 2006).HeLa cells (1.4�106 per 60 mm plate) were transfected with Lipofectamine 2000(Invitrogen), and 20 hours later, cells were lysed on ice with 500 μl of lysis buffer[1% Triton X-100, 50 mM Tris-HCl (pH 8), 100 mM NaCl, 10 mM MgCl2, 0.05%sodium deoxycholate, 0.005% SDS, 10% glycerol, 2 mM dithiothreitol, and proteaseinhibitors]. Lysates were clarified by centrifugation at 13,000 � g for 10 minutesand incubated with 30 μg GST-ARHGAP10 Arf-binding domain (Dubois et al., 2005)bound to glutathione-Sepharose beads (Amersham Biosciences) supplemented with0.5% BSA for 40 minutes at 4°C. The beads were washed three times with lysisbuffer without sodium deoxycholate and SDS, and bound proteins were eluted in30 μl of SDS sample buffer. The presence of Arf6-GTP was detected byimmunoblotting using an anti-Myc monoclonal antibody (9E10, Santa CruzBiotechnology).

We are grateful to Michael Way for providing the dominant-negativeand constitutively active Rac1, Cdc42 and RhoA GFP fusion constructs,and to Michael Franco for the gift of GST-ARHGAP10 Arf-bindingdomain. We are indebted to Francesco Blasi for support, helpfuldiscussions and critical review of the manuscript. We thank GiuseppeRotondo for helpful suggestions, and Vincenzo Zimarino for criticalreading of the manuscript. This work was supported by a Telethon-Italy grant (E. 0857) to D.T.

ReferencesAlbertinazzi, C., Za, L., Paris, S. and de Curtis, I. (2003). ADP-ribosylation factor 6

and a functional PIX/p95-APP1 complex are required for Rac1B-mediated neuriteoutgrowth. Mol. Biol. Cell 14, 1295-1307.

Bretscher, M. S. and Aguado-Velasco, C. (1998). Membrane traffic during cell locomotion.Curr. Opin. Cell Biol. 10, 537-541.

Caumont, A. S., Galas, M. C., Vitale, N., Aunis, D. and Bader, M. F. (1998). Regulatedexocytosis in chromaffin cells: translocation of ARF6 stimulates a plasma membrane-associated phospholipase D. J. Biol. Chem. 273, 1373-1379.

Choi, S., Ko, J., Lee, J. R., Lee, H. W., Kim, K., Chung, H. S., Kim, H. and Kim, E.(2006). ARF6 and EFA6A regulate the development and maintenance of dendritic spines.J. Neurosci. 26, 4811-4819.

Cox, R., Mason-Gamer, R. J., Jackson, C. L. and Segev, N. (2004). Phylogenetic analysisof Sec7-domain-containing Arf nucleotide exchangers. Mol. Biol. Cell 15, 1487-1505.

da Silva, J. S. and Dotti, C. G. (2002). Breaking the neuronal sphere: regulation of theactin cytoskeleton in neuritogenesis. Nat. Rev. Neurosci. 3, 694-704.

Decressac, S., Franco, M., Bendahhou, S., Warth, R., Knauer, S., Barhanin, J.,Lazdunski, M. and Lesage, F. (2004). ARF6-dependent interaction of the TWIK1 K+channel with EFA6, a GDP/GTP exchange factor for ARF6. EMBO Rep. 5, 1171-1175.

Derrien, V., Couillault, C., Franco, M., Martineau, S., Montcourrier, P., Houlgatte, R.and Chavrier, P. (2002). A conserved C-terminal domain of EFA6-family ARF6-guaninenucleotide exchange factors induces lengthening of microvilli-like membrane protrusions.J. Cell Sci. 115, 2867-2879.

Donaldson, J. G. (2003). Multiple roles for Arf6: sorting, structuring, and signaling at theplasma membrane. J. Biol. Chem. 278, 41573-41576.

D’Souza-Schorey, C. and Chavrier, P. (2006). ARF proteins: roles in membrane trafficand beyond. Nat. Rev. Mol. Cell. Biol. 7, 347-358.

Dubois, T., Paleotti, O., Mironov, A. A., Fraisier, V., Stradal, T. E., De Matteis, M. A.,Franco, M. and Chavrier, P. (2005). Golgi-localized GAP for Cdc42 functions

Jour

nal o

f Cel

l Sci

ence

downstream of ARF1 to control Arp2/3 complex and F-actin dynamics. Nat. Cell Biol.7, 353-364.

Franco, M., Peters, P. J., Boretto, J., van Donselaar, E., Neri, A., D’Souza-Schorey,C. and Chavrier, P. (1999). EFA6, a sec7 domain-containing exchange factor for ARF6,coordinates membrane recycling and actin cytoskeleton organization. EMBO J. 18, 1480-1491.

Govek, E. E., Newey, S. E. and Van Aelst, L. (2005). The role of the Rho GTPases inneuronal development. Genes Dev. 19, 1-49.

Hashimoto, S., Onodera, Y., Hashimoto, A., Tanaka, M., Hamaguchi, M., Yamada, A.and Sabe, H. (2004). Requirement for Arf6 in breast cancer invasive activities. Proc.Natl. Acad. Sci. USA 101, 6647-6652.

Hernandez-Deviez, D. J., Casanova, J. E. and Wilson, J. M. (2002). Regulation ofdendritic development by the ARF exchange factor ARNO. Nat. Neurosci. 5, 623-624.

Hernandez-Deviez, D. J., Roth, M. G., Casanova, J. E. and Wilson, J. M. (2004). ARNOand ARF6 regulate axonal elongation and branching through downstream activation ofphosphatidylinositol 4-phosphate 5-kinase alpha. Mol. Biol. Cell 15, 111-120.

Jan, Y. N. and Jan, L. Y. (2003). The control of dendrite development. Neuron 40, 229-242.

Jaworski, J. (2007). ARF6 in the nervous system. Eur. J. Cell Biol. 86, 513-524.Kim, H. S., Chen, Y. and Lonai, P. (1998). Complex regulation of multiple cytohesin-

like genes in murine tissues and cells. FEBS Lett. 433, 312-326.Klein, S., Franco, M., Chardin, P. and Luton, F. (2006). Role of the Arf6 GDP/GTP

cycle and Arf6 GTPase-activating proteins in actin remodeling and intracellular transport.J. Biol. Chem. 281, 12352-12361.

Kozak, M. (1987). An analysis of 5�-noncoding sequences from 699 vertebrate messengerRNAs. Nucleic Acids Res. 15, 8125-8148.

Lupas, A. (1996). Prediction and analysis of coiled-coil structures. Methods Enzymol. 266,513-525.

Matsuya, S., Sakagami, H., Tohgo, A., Owada, Y., Shin, H. W., Takeshima, H.,Nakayama, K., Kokubun, S. and Kondo, H. (2005). Cellular and subcellularlocalization of EFA6C, a third member of the EFA6 family, in adult mouse Purkinjecells. J. Neurochem. 93, 674-685.

Miyazaki, H., Yamazaki, M., Watanabe, H., Maehama, T., Yokozeki, T. and Kanaho,Y. (2005). The small GTPase ADP-ribosylation factor 6 negatively regulates dendriticspine formation. FEBS Lett. 579, 6834-6838.

Naslavsky, N., Weigert, R. and Donaldson, J. G. (2003). Convergence of non-clathrin-and clathrin-derived endosomes involves Arf6 inactivation and changes inphosphoinositides. Mol. Biol. Cell 14, 417-431.

Niedergang, F., Colucci-Guyon, E., Dubois, T., Raposo, G. and Chavrier, P. (2003).ADP ribosylation factor 6 is activated and controls membrane delivery duringphagocytosis in macrophages. J. Cell Biol. 161, 1143-1150.

Perletti, L., Talarico, D., Trecca, D., Ronchetti, D., Fracchiolla, N. S., Maiolo, A. T.and Neri, A. (1997). Identification of a novel gene, PSD, adjacent to NFKB2/lyt-10,which contains Sec7 and pleckstrin-homology domains. Genomics 46, 251-259.

Randazzo, P. A. and Hirsch, D. S. (2004). Arf GAPs: multifunctional proteins that regulatemembrane traffic and actin remodelling. Cell. Signal. 16, 401-413.

Reichardt, L. F. (2006). Neurotrophin-regulated signalling pathways. Philos. Trans. R.Soc. Lond. B Biol. Sci. 361, 1545-1564.

Sakagami, H. (2008). The EFA6 family: guanine nucleotide exchange factors for ADPribosylation factor 6 at neuronal synapses. Tohoku J. Exp. Med. 214, 191-198.

Sakagami, H., Matsuya, S., Nishimura, H., Suzuki, R. and Kondo, H. (2004).Somatodendritic localization of the mRNA for EFA6A, a guanine nucleotide exchangeprotein for ARF6, in rat hippocampus and its involvement in dendritic formation. Eur.J. Neurosci. 19, 863-870.

Sakagami, H., Suzuki, H., Kamata, A., Owada, Y., Fukunaga, K., Mayanagi, H. andKondo, H. (2006). Distinct spatiotemporal expression of EFA6D, a guanine nucleotideexchange factor for ARF6, among the EFA6 family in mouse brain. Brain Res. 1093,1-11.

Sakagami, H., Honma, T., Sukegawa, J., Owada, Y., Yanagisawa, T. and Kondo, H.(2007). Somatodendritic localization of EFA6A, a guanine nucleotide exchange factorfor ADP-ribosylation factor 6, and its possible interaction with alpha-actinin in dendriticspines. Eur. J. Neurosci. 25, 618-628.

Santy, L. C. and Casanova, J. E. (2001). Activation of ARF6 by ARNO stimulatesepithelial cell migration through downstream activation of both Rac1 and phospholipaseD. J. Cell Biol. 154, 599-610.

Schweitzer, J. K. and D’Souza-Schorey, C. (2002). Localization and activation of theARF6 GTPase during cleavage furrow ingression and cytokinesis. J. Biol. Chem. 277,27210-27216.

Shin, H. W. and Nakayama, K. (2004). Guanine nucleotide-exchange factors for arfGTPases: their diverse functions in membrane traffic. J. Biochem. (Tokyo) 136, 761-767.

Suzuki, I., Owada, Y., Suzuki, R., Yoshimoto, T. and Kondo, H. (2001). Localizationof mRNAs for six ARFs (ADP-ribosylation factors) in the brain of developing and adultrats and changes in the expression in the hypoglossal nucleus after its axotomy. BrainRes. Mol. Brain Res. 88, 124-134.

Suzuki, I., Owada, Y., Suzuki, R., Yoshimoto, T. and Kondo, H. (2002). Localizationof mRNAs for subfamily of guanine nucleotide-exchange proteins (GEP) for ARFs (ADP-ribosylation factors) in the brain of developing and mature rats under normal andpostaxotomy conditions. Brain Res. Mol. Brain Res. 98, 41-50.

Tague, S. E., Muralidharan, V. and D’Souza-Schorey, C. (2004). ADP-ribosylation factor6 regulates tumor cell invasion through the activation of the MEK/ERK signaling pathway.Proc. Natl. Acad. Sci. USA 101, 9671-9676.

Threadgill, R., Bobb, K. and Ghosh, A. (1997). Regulation of dendritic growth andremodeling by Rho, Rac, and Cdc42. Neuron 19, 625-634.

Turner, C. E. and Brown, M. C. (2001). Cell motility: ARNO and ARF6 at the cuttingedge. Curr. Biol. 11, R875-R877.

Yang, C. Z. and Mueckler, M. (1999). ADP-ribosylation factor 6 (ARF6) definestwo insulin-regulated secretory pathways in adipocytes. J. Biol. Chem. 274, 25297-25300.

Yang, C. Z., Heimberg, H., D’Souza-Schorey, C., Mueckler, M. M. and Stahl, P. D.(1998). Subcellular distribution and differential expression of endogenous ADP-ribosylation factor 6 in mammalian cells. J. Biol. Chem. 273, 4006-4011.

Journal of Cell Science 122 (12)2118

Jour

nal o

f Cel

l Sci

ence