Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2009 Jun 27.
Published in final edited form as: J Biol Chem. 2008 Jan 16;283(12):7733–7744. doi: 10.1074/jbc.M707142200

Amyloidogenic Processing but not AICD Production Requires a Precisely Oriented APP Dimer Assembled by Transmembrane GXXXG Motifs

Pascal Kienlen-Campard *, Bernadette Tasiaux *, Joanne Van Hees #, Mingli Li #, Sandra Huysseune *, Takeshi Sato %, Jeffrey Z Fei &, Saburo Aimoto %, Pierre J Courtoy §, Steven O Smith &, Stefan N Constantinescu #,£, Jean-Noël Octave *,£
PMCID: PMC2702479  NIHMSID: NIHMS115586  PMID: 18201969

Abstract

The β-amyloid peptide (Aβ) is the major constituent of the amyloid core of senile plaques found in the brain of patients with Alzheimer's disease (AD). Aβ is produced by the sequential cleavage of the Amyloid Precursor Protein (APP) by β- and γ-secretases. Cleavage of APP by γ-secretase also generates the APP Intracellular C-terminal Domain (AICD) peptide, which might be involved in regulation of gene transcription. APP contains three glycine-xxx-glycine (GxxxG) motifs in its juxtamembrane and transmembrane (TM) regions. Such motifs are known to promote dimerization via close apposition of TM sequences. We demonstrate that pairwise replacement of glycines by leucines or isoleucines, but not alanines, in a GxxxG motif led to a drastic reduction of Aβ40 and Aβ42 secretion. β-Cleavage of mutant APP was not inhibited, and reduction of Aβ secretion resulted from inhibition of γ-cleavage. It was anticipated that decreased γ-cleavage of mutant APP would result from inhibition of its dimerization. Surprisingly, mutations of the GxxxG motif actually enhanced dimerization of the APP C-terminal fragments, possibly via a different TM α-helical interface. Increased dimerization of the TM APP C-terminal domain did not affect AICD production. These results clearly demonstrate that both orientation and dimerization of the APP TM domain differently affect Aβ and AICD production.


The progressive deposition of β-amyloid peptide (Aβ)1 leading to the formation of senile plaques is an invariant feature of Alzheimer's disease (AD). Aβ is a 39 to 43 amino acid peptide, with two major isoforms of 40 and 42 amino acids (1;2). Aβ is produced by the amyloidogenic cleavage of its precursor, the Amyloid Precursor Protein or APP (3).

The amyloidogenic processing of APP is initiated by β-cleavage within the lumenal/extracellular domain of the protein. The β-cleavage of APP is performed by the BACE proteins (BACE1 and 2) that are integral membrane proteins belonging to the aspartyl protease family (4-8). β-cleavage produces a 99 amino acid, membrane-anchored APP C-terminal fragment (βCTF), which is further cleaved by the γ-secretase activity to generate Aβ. The γ-secretase activity is contained in a high molecular weight multiprotein complex formed at least by the following proteins: a Presenilin (PS1 or PS2), Nicastrin (Nct), Pen-2 and Aph-1 (9). The activity of the γ-secretase complex is also required for the generation of the intracellular fragment named AICD (APP Intracellular C-terminal Domain). AICD was shown to translocate to the nucleus (10;11), and there is growing experimental evidence suggesting a role for AICD in the regulation of gene transcription (12-17), even if the identity of APP target genes remains a matter of debate (18). The γ-secretase complex therefore plays a central role in the onset and progression of AD, not only because proteolysis of βCTF controls the production of Aβ, but it also controls the intracellular signaling associated with APP, which in turn might regulate the expression of genes involved in the disease.

The physical interaction between APP and the secretases or other partners is crucial for its processing, and yet it is very poorly understood. APP contains several glycine-xxx-glycine (GxxxG) motifs at the junction between the juxtamembrane and transmembrane (TM) regions (19-22). GxxxG motifs are known from the sequence of the glycophorin A (GpA) protein to mediate sequence-specific dimerization and very close apposition of TM helices (23). In glycophorin A, the sequence LIxxGVxxGVxxT mediates tight dimerization between TM helices (24) by direct glycine-glycine contacts (25). It has later been recognized that GxxxG motifs can mediate more generic oligomerization of TM domains (26). More importantly, it has been shown that glycine is compatible with α-helical secondary structure in lipid bilayers and that, due to its small size, this residue allows the close association of interacting helices (27;28). GxxxG motifs have been shown to play a role in the assembly, trafficking and activity of several proteins of the γ-secretase complex (29;30).

The presence of three GxxxG motifs in APP suggests that the glycine face of the APP TM helix may be involved in interactions with other proteins or with itself, and offers a molecular basis for APP homo- and hetero-oligomerization. Strikingly, one genetic mutation that leads to early-onset Alzheimer's disease, the Flemish mutation, is represented by the APP A617G mutation, which creates a fourth in-register GxxxG motif preceding the TM domain of APP (31). Moreover, we have reported that the GxxxG motifs play a major role in fibrillization of Aβ40 and Aβ42 (21). The mechanisms by which homo- or heterodimerization of APP (32) act on its processing are far from being understood.

Here we show that APP processing via the amyloidogenic pathway to both Aβ40 and Aβ42 depends on the presence of a small residue, either glycine or alanine at the position of the GxxxG motifs. Pairwise replacement of glycine by leucine within these motifs in human APP695 leads to significantly less Aβ production. One APP mutant in particular, where glycines of the middle GxxxG motif (G625 and G629) were mutated to leucine, exhibited altered amyloidogenic processing with a drastic reduction of Aβ formation. Similar results were observed when glycines (G625 and G629) were mutated to isoleucines. The mutation of GxxxG motifs to LxxxL did not affect β-cleavage or the APP-BACE1 interaction, but decreased the γ-cleavage of APP, without impairing the APP-PS1 interaction. Strikingly, the γ-secretase-mediated release of AICD was not altered by the GxxxG to LxxxL mutation. Unexpectedly, the mutation of the middle GxxxG motif (G625xxxG629) to LxxxL, rather than weakening dimerization, enhanced the formation of homodimers of the APP C-terminal fragments. Since the leucine substitutions are not compatible with the close helix packing allowed by the GxxxG motifs, they must alter the helix orientation within the homodimer. Taken together our results show that the GxxxG motifs in the TM domain of APP are required for Aβ, but not for AICD production, and that dimerization per se does not suffice to promote the amyloidogenic processing of APP. Both orientation and dimerization of the APP TM domain differently affect the cleavage of APP by γ-secretase, a process critically involved in AD.

Experimental Procedures

Antibodies and Reagents

The polyclonal anti-human Aβ used for IP was described previously (33). The human specific WO-2 and 6E10 antibodies were from The Genetics Company (Schlieren, Switzerland) and Signet Laboratories (Dedham, MA), respectively. The polyclonal antibody directed against the APP C-terminus (34) was kindly provided by N. Sergeant (INSERM U422, Lille, France). The anti-actin polyclonal antibody was from Sigma (St Louis, MO). The polyclonal anti BACE1 and monoclonal MAB1563 antibodies directed against human PS1 were from Oncogene Research (Cambridge, MA) and Chemicon (Temecula, CA), respectively. The monoclonal anti-HA and polyclonal anti-Myc were purchased from Santa Cruz (Santa Cruz, CA) and Roche (Basel, Switzerland), respectively. Secondary antibodies were from Amersham Bioscience (Uppsala, Sweden). All cell culture media and antibiotics were from Invitrogen (Carlsbad, CA).

Plasmids

A cDNA encoding the full-length sequence of human APP695 was inserted into the SmaI/SalI restriction sites of pSVK3 vector (pSVK3-APP695). All the mutated APP695, C99 HA- and Myc-tagged C99 constructs were derived from the pSVK3-APP695 parental vector. Mutated human APP695Gal4 constructs were derived from the pMst-APP (APPGal4) parental vector (15). cDNAs coding for the 6 double glycine mutations in the APP juxtamembrane and transmembrane domains were generated by overlap extension PCR with synthetic oligonucleotides (see Supplemental data). All constructs were verified by full sequencing.

Cell Cultures and Transfection

The culture of CHO cells expressing human APP and /or human PS1 has been previously described (35). For transient transfection, cells were seeded at the density of 3.105 cells/cm2 24 h prior to transfection (2μg pSVK3APP-695/well) with Lipofectamine™ according to manufacturer's instruction (Invitrogen, Carlsbad, CA). Stable cell lines expressing human APP mutants were established by co-transfecting pSVK3-APP expression vector and PSV2Neo vector at a 10:1 ratio (2 μg DNA/well). Whole populations expressing human APP695 were selected in the presence of 500 μg/ml G418 and further subcloned.

APPGal4 Transactivation Assays

CHO cells were transfected with the following plasmids: pMst-APPGal4 expression vectors (0.4 μg/well), pG5E1B-luc (0.4 μg/well), pCMV5-Fe65 (0.2 μg/well), phRG-TK (0.02 μg/well). Cells were harvested 48 h post-transfection in 0.1 ml/well reporter lysis buffer and the firefly and renilla luciferase activities were measured with the Dual Glo™ luciferase assay system (Promega). The firefly luciferase activity was standardized by the Renilla luciferase activity to control for transfection efficiency.

Immunoprecipitation and Co-immunoprecipitation

Immunoprecipitation of Aβ from the culture media with a polyclonal anti-Aβ antibody (30 μl/ml) has been previously described (36). Cells were grown in 6-well dishes washed in ice cold PBS, solubilized in IP buffer (1 ml/well of 25 mM Tris pH 7.6, 0.5 % Triton X-100®, 0.5 % NP40 containing a protease inhibitor mix). Cellular extracts were cleared by centrifugation (20,000 g, 5 min, 4°C). Immunoprecipitation of APP was carried out on the supernatants with 2.5 μg/ml of anti-human APP (WO-2) antibody. Co-immunoprecipitations were carried out in the same conditions with 2.5 μl/ml of anti BACE1, 2 μl/ml of anti human PS1 (MAB1563), 2.5 μg/ml of anti-human APP (WO-2), 5 μl/ml of anti-HA or 5 μl/ml of anti-Myc antibodies. Immunoprecipitates were analyzed by Western blotting as described below.

Blue Native Gel Electrophoresis and Western Blotting

Blue native (BN)-PAGE was performed on membrane protein extracts following the procedure of Schagger and von Jagow (37). Membrane proteins were prepared by homogenizing cells in 10% Glycerol, 150 mM NaCl, 20 mM Hepes pH 7.4 in the presence protease inhibitors (Roche Molecular Biochemicals, Basel, Switzerland). Homogenates were centrifuged (1,000 g, 5 min, 4°C) and supernatants were recovered and centrifuged for 1h at 100,000 g (4°C). Pellets were resuspended in solubilization buffer containing 100 μl of 1% CHAPS, 500 mM ε-amino-n-caproic acid (EACA), 50 mM Bis-Tris pH 7.0 and protease inhibitors. After centrifugation (100,000 g, 20 min, 4°C), supernatants corresponding to 150 μg of proteins were added to 0.5 vol. of loading buffer (0.75 % Serva Blue, 500 mM EACA, 30 % Glycerol, 1 % CHAPS, 50 mM Bis-Tris pH 7.0). The protein complexes were separated on a 6–10% acrylamide gel using 50 mM Bis-Tris pH 7.0 as an anode buffer and 15 mM Bis-Tris/50 mM tricine containing 0.02% Serva Blue G-250 as a cathode buffer. Ferritine, thyroglobulin and β-globin were used as molecular weight markers. Before blotting, the gels were soaked in Tris–glycine transfer buffer (25 mM Tris, 192 mM glycine, 20 % methanol, 0.01 %SDS) containing 0.02% SDS for 10 min, and the proteins were transferred to nitrocellulose Hybond-C membranes (Amersham Bioscience, Uppsala, Sweden) at 250 mA for 2 h and probed with specific antibodies. Western blotting was performed on cell lysates (10 μg protein) or culture media (15 μl) separated into NuPage™ 4-12 % Bis-Tris gels (Invitrogen, Carlsbad, CA) and further transferred to nitrocellulose membrane (Amersham Bioscience, Uppsala, Sweden). Membranes were incubated overnight at 4°C with the primary antibody (0.5 μg/ml WO-2, 1:2,500 anti-APP C-ter, 1:1000 anti-BACE1, 1:5000 MAB1563, 1:1,000 anti-HA, 1:1,000 anti-Myc, 1:2,000 anti-actin), washed and incubated with 1:10,000 secondary antibody (anti-mouse or anti-rabbit Ig) conjugated to horseradish peroxidase, followed by ECL revelation. Chemiluminescence films (Amersham, Uppsala, Sweden) were digitalized with a Gel Doc 2000 imaging device coupled to the Quantity One™ analysis software (Bio-Rad, Hercules, CA) for quantification. The cellular APP levels were normalized against actin for quantification.

Surface Biotinylation of APP

Cell surface biotinylation was performed as previously described (38). CHO cells expressing human APP (∼ 5.106 cells) were washed with PBS containing 2 mM Ca2+ and 1.2 mM Mg2+ and incubated with 0.8 mL of sulfo-NHS-biotin (Pierce) at 1.5 mg/mL in PBS for 30 min at 4°C with mild shaking. Cells were then washed twice with cold PBS containing 100 mM glycine, incubated with the same solution for 45 min at 4°C to quench the unbound biotin reagent and further solubilized in lysis buffer (25 mM Tris-HCl, pH 6.8, 0.5% (v/v) Triton X-100, 0.5% (v/v) Nonidet P-40]) supplemented with protease inhibitors (1 μg/mL pepstatin, 10 μg/mL leupeptin, 1 mM PMSF) for 1 h at 4°C with vigorous shaking. After centrifugation at 15,800 g at 4°C for 20 min, 150 μL of supernatant fluid were incubated with an equal volume of streptavidin bead suspension (Pierce) for 1 h at room temperature. After centrifugation (15,800 g, 15 min, 4°C), supernatant fluids were collected for analysis of the non-biotinylated intracellular fraction. Biotinylated cell surface proteins contained in the pellet were washed four times with 0.8 mL lysis buffer and resuspended in 50 μL Western blotting sample buffer containing 50 mM dithiothreitol (DTT). Samples were analyzed by Western blotting.

Analytical Subcellular Fractionation

CHO cells were recovered in 0.25 M sucrose containing 1 mM EDTA, 3 mM imidazole buffered at pH 7.4 and complete protease inhibitors (Roche, Basel, Swizerland). The cellular suspension was homogenized in a tight Dounce homogenizer. A low-speed nuclear fraction was pelleted at 104 g for 10 min and washed three times by resuspension and sedimentation. Pooled postnuclear supernatants were further sedimented at 105 g for 60 min in a Ti50 rotor (Beckman, Palo Alto, CA). A high-speed particular pellet was resuspended in 0.3 ml of homogenization buffer, mixed with 2.3 M sucrose to reach 1.28 g/ml in density and layered at the bottom of a linear sucrose gradient (from 1.10 to 1.24 g/ml in density). After floatation by centrifugation at 2.105 g for 22 h in a Sw40 rotor (Beckman), 12 fractions were collected and analyzed for protein content. Western blotting and quantifications were performed as described below. Results were represented as normalized histograms (39).

ELISA

Two-days after plating or transfection, cells were reincubated for 8 h in fresh culture medium. Culture media were recovered and incubated with protease inhibitors (1μg/ml pepstatin, 10 μg/ml leupeptin, 1mM PMSF). Samples were cleared by centrifugation (16,000 g, 5 min, 4°C). IP Buffer (1:10 vol/vol) was added to supernatants to a final concentration of 25 mM Tris pH 7.6, 0.5 % Triton X-100®, 0.5% NP-40. Aβ40 and Aβ42 were quantified using fluorescent ELISA assays from BioSource (Camarillo, CA). Soluble αAPP and soluble βAPP were quantified in multiplex from 25 μl of cellular medium using the electrochemiluminescence detection method of Meso Scale Discovery (Gaithersburg, Maryland, USA).

Computational Searches

Low-energy conformations of helix dimers were searched by rotating each helix through rotation angles φ1 and φ2 from 0° to 360° with a sampling step sizes of 25-45° and with an interhelical separation of 9.0 Å, 9.5 Å, 10.0 Å and 10.5 Å. The φ1 and φ2 angles are equivalent to the α and β angles originally described by Adams et al. (40). MD simulations were run with simulated annealing at each rotational orientation of the dimer using the program X-PLOR along with the united atom topology and parameters sets, TOPH19 and PARAM19, respectively. Calculations were carried out for helices with both left- and right-handed geometries with initial crossing angles of 25°. Four different runs were carried out for each starting geometry using torsional angle dynamics. The rotation and crossing angles were allowed to vary during the simulations. To maintain an α-helical conformation, distance restraints were applied between Oi and Ni+4 atoms along the backbone. For the APP homodimers, computational searches were carried out on residues 620-648. Position 624 (lysine) was neutral for the MD simulations and charged for the final energy minimization.

Results

GxxxG to LxxxL mutation reduces the amyloidogenic processing of APP

To investigate the role of the GxxxG motifs in the processing of human APP695, we simultaneously replaced by site-directed mutagenesis the two Gly (G) residues of each GxxxG motif to either Ala (A) or Leu (L). The amino acid substitutions at positions 621, 625, 629 and 633, that correspond to the positions 25, 29, 33 and 37 of Aβ are indicated for each mutant (Fig. 1A). The mutated positions are located at the interface between the extracellular and TM regions (621/625, mutants 1 and 3), or in the predicted TM domain (625/629 and 629/633, mutants 2-3 and 5-6). None of the mutations are directly located at a known APP cleavage.

Fig. 1.

Fig. 1

Expression and processing of human APP GxxxG mutants in CHO cells. A. Schematic representation of the transmembrane and juxtamembrane domains of human APP695. The positions of the three consecutive GxxxG motifs (621-625, 625-629, 629-633) are highlighted. The amino acid substitution (G to A or G to L) generated for each mutant appears in bold underlined. The cleavage sites of α-(α), β- (β and β′) and γ- (γ and ε) secretase activities are indicated by arrows. The epitopes of the human-specific WO-2 antibody are also shown along with the C-terminal position recognized by the Aβ40 and Aβ42 specific antibodies for ELISA. B. The expression of cellular APP695 or APP mutants was analyzed 48 after transfection by Western blotting revealed by the WO-2 antibody. The presence of APP is indicated by arrows. Actin was used as a protein loading control. Forty hours after transfection, CHO cells were conditioned in fresh culture medium for 8 h. The accumulation of soluble APPα (sαAPP) was analyzed by Western blotting revealed by the WO-2 antibody. Aβ was immunoprecipitated from the same culture medium and analyzed by Western blotting revealed by the WO-2 antibody. C. The Aβ to sαAPP ratio was calculated and represented as percentage of Aβ/sαAPP production in non-mutated controls (APP695). Values are means ± sem, n = 4; *p < 0.05, ** p < 0.01, *** p < 0.001, compared to control.

CHO cells were transfected with the plasmids encoding APP695 and the different APP695 mutants (1 to 6, see Fig. 1). Cells were harvested 48 h after transfection and the expression of full-length human APP was analyzed in cell lysates by Western blotting with the human specific WO-2 antibody. We detected the expression of the different mutants at the expected size (Fig. 1B). The release of soluble APP (sαAPP), which is generated by the α-secretase activity, was monitored in the corresponding extracellular culture medium. We detected production of the soluble APPα (sαAPP) by Western blotting of culture medium from transfected CHO cells with the same WO-2 antibody. All of the mutants produced comparable levels of sαAPP (Fig. 1B).

When Aβ production was analyzed by immunoprecipitation with polyclonal anti-human Aβ antibodies from the culture medium of transfected cells, we detected a significant decrease in Aβ production in mutants 4, 5 and 6, whereas when GxxxG motifs were replaced by alanine (mutant 1 to 3), there was no difference in Aβ production as compared to wild type APP (Fig. 1B). This decrease, especially for APP mut5, is also evident when the levels of extracellular Aβ were normalized to the levels of sαAPP and represented as the percentage of the Aβ/sαAPP ratio measured in the culture medium of cells expressing wild-type APP695 (Fig. 1C). Since the different APP mutations are located in the Aβ sequence, we verified that the WO-2 antibody was still able to recognize mutated Aβ. In addition, the decrease in Aβ production was confirmed for APP mut5 with another antibody (6E10) directed against the N-terminus of human Aβ, and this decrease did not result from the accumulation of Aβ in intracellular compartments or in cell membranes (not shown).

The very strong decrease observed, particularly in mut5 (GG 625/629 LL) is consistent with the idea that the balance between amyloidogenic and non-amyloidogenic processing of APP was affected when glycines from GxxxG motifs were replaced by leucines, but not by alanines.

GG 625/629 LL mutation impairs neither the cellular localization nor the β-cleavage of APP

The consequences of the glycine to leucine mutations on APP amyloidogenic processing were analyzed in detail in CHO cell lines stably expressing APP695 or APP mut5 (GG 625/629 LL). A plausible explanation for these results was that the G to L mutation disrupted the interaction of APP with binding partners, leading to mislocalization of the protein. Consistently APP mutant5 exhibited an increase in the mature forms of full length APP (Fig. 1B). This may indicate a more efficient processing through the Golgi apparatus or a failure in endocytosis (41). Analytical subcellular fractionation and cell surface biotinylation experiments were carried out to study the subcellular distribution of APP and APP mut5. Using analytical subcellular fractionation, we did not observe any significant difference in the cellular distribution of APP and APP mut5 (Fig. 2A and 2B). As shown in Fig. 2C, similar levels of APP and APP mut5 were detected at the cell surface following cell surface biotinylation. This indicated that the decrease in Aβ did not result from a cellular redistribution of APP induced by the glycine to leucine mutation.

Fig. 2.

Fig. 2

GG 625/629 LL mutation does not modify the cellular distribution of APP. A. The particulate distribution of APP695 and APP mutant 5 (GG625/629LL) was analyzed in CHO cells by floatation in sucrose gradient and Western blotting revealed by the WO-2 antibody. Arrows indicate the expected positions of APP. B. Density distributions are represented as normalized histograms. C. The presence of APP at the plasma membrane was studied by cell surface biotinylation of CHO cells expressing APP and APP mutant 5. The presence of full length APP was detected by Western blotting using the human-specific WO-2 antibody in total (T), intracellular (IC) and cell surface (S) fractions. Arrows indicate the expected positions of APP. D. Quantification of surface biotinylation experiments. The cell surface APP (APPs) to total APP (APPt) ratio was calculated and given as percentage. Values are means ± sem, n = 5, ns = non-significant.

We investigated a possible role of the GxxxG motifs in the β-cleavage of APP. The β-cleavage of APP is performed by the BACE1 protein (5). Since the TM domain of BACE1 contains two AxxxA motifs that may also promote the interaction of TM helices (42), we analyzed the effects of the GG 625/629 LL mutation on APP-BACE1 interaction. The association of APP and BACE1 in cell membrane was analyzed by non-denaturating gel electrophoresis (Blue Native PAGE). The results shown in Fig. 3B strongly suggested that the formation of APP-BACE1 complexes was not impaired by the GG 625/629 LL mutation. This was confirmed by co-immunoprecipitation experiments (Fig. 3C). We further showed that soluble βAPP (sβAPP), produced by the cleavage of APP by β-secretase, was not affected in APP mut5 (Fig. 3D). The quantification of sαAPP in the same sample allowed us to conclude that the sβAPP/sαAPP ratio was the same for APP695 and APP mut5. In addition, the cellular APP metabolite produced by β-secretase (βCTF) did not decrease, but indeed accumulated in APP mut5 (Fig. 3E). This indicated that the GG 625/629 LL mutation did not impair the β-cleavage of APP, but seems to have blocked γ-cleavage.

Fig. 3.

Fig. 3

GG 625/629 LL mutation does not impair β-cleavage. A. Schematic representation of the APP fragments produced by β-secretase activity (sβAPP and CTFβ) along with the epitope recognized by the APP C-terminal antibody (C-ter). B. APP-BACE1 interaction was analyzed in CHO cell lines stably expressing APP695 and APP mut5 (GG625/629LL) or co-expressing APP and BACE1 by non-denaturating (Blue Native, BN) electrophoresis. Levels of BACE1 and APP in cell extracts was measured by Western blotting revealed with the WO-2 or BACE1 antibody in denaturating conditions (SDS PAGE, upper panel). Arrows indicate the expected position of cellular monomeric BACE1 or APP. The formation of cellular APP/BACE1 complexes was investigated by Blue Native electrophoresis in 6%-10% polyacrylamide gels (BN PAGE, lower panel) followed by Western blotting revealed with the WO-2 or BACE1 antibody. Arrows indicate the expected position of monomeric APP (mon APP), monomeric BACE (mon BACE1) or APP/BACE1 complexes. C. APP/BACE1 interaction analyzed by co-immunoprecipitation. Levels of BACE1 and APP were measured by Western blotting in cell extracts prior to immunoprecipitation. Cell lysates were immunoprecipitated either by the WO-2 or by the anti BACE1 antibody and revealed by Western blotting with the WO-2 or the BACE1 antibody as indicated. Arrows indicate the expected position of cellular BACE1 or APP. D. Soluble αAPP and soluble βAPP were analyzed in the same sample (extracellular medium) by multiplex assays. The production of human sβAPP from non-transfected (Co) cells and cells expressing APP695 and APP mut5 is given in ng/ml of culture medium (left) or as sβAPP/sαAPP ratio represented as percentage of APP 695. Values are means ± sem, n = 6; ns = non-significant. E. Expression of full-length APP and βCTF revealed with the C-terminal antibody.

Impact of GG 625/629 LL mutation on γ-cleavage and Aβ production

To investigate the role of the GxxxG motifs in the γ-cleavage of APP, we measured Aβ production from cells expressing either APP or C99 bearing or not the GG 625/629 LL mutation (Figs. 4C and 4E). The C99 protein corresponds to the β-cleaved C-terminal fragment of APP fused to the signal peptide sequence (Fig. 4A), and thus Aβ is released from C99 by a single cleavage performed by γ-secretase (43). We specifically measured Aβ40 and Aβ42 production by ELISA in CHO cell supernatants. Both Aβ40 and Aβ42 were present in the culture medium of cells expressing APP695 or C99 (Fig. 4C and 4E). Importantly, Aβ40 production was abolished and the level of Aβ42 was also strongly decreased in cells expressing APP mut5 or C99 mut5. The decrease in Aβ40 and Aβ42 in cells expressing APP mut5 was not compensated by an increase in shorter Aβ isoforms (Supp. Fig.1). This was consistent with the overall decrease in Aβ measured by immunoprecipitation (Fig. 1C). Similar levels of C99 or C99 mut5 were detected in the cells. However, in cells expressing C99 mut5, a band corresponding to a possible C99 dimer was detected, whereas such a band was present only at very low levels in C99 cells (Fig. 4D).

Fig. 4.

Fig. 4

Effects of GG 625/629 LL mutation on γ-cleavage and Aβ production. A. Schematic representation of the APP695 and C99 proteins. The epitope recognized by the WO-2 antibody is underlined. B. The expression of APP695, APP mut5 was measured by Western blotting in cell lysates revealed by the WO-2 antibody. The presence of APP is indicated by an arrow. C. Aβ1-40 and Aβ1-42 production was monitored by ELISA in the culture medium of cells expressing APP or APP mut5, and given as Aβ levels in pg/ml. Values are means ± sem, n = 4; * p < 0.05, *** p < 0.001, compared to control. D. The expression of C99, C99 mut5 was measured by Western blotting in cell lysates revealed by the WO-2 antibody. The presence of C99 is depicted by an arrow, the asterisk indicates the presence of possible C99 dimers. E. Aβ1-40 and Aβ1-42 production was monitored by ELISA in the culture medium of cells expressing C99 or C99 mut5, and given as Aβ levels in pg/ml. Values are means ± sem, n = 4; *** p < 0.001, compared to control.

This impairment of APP γ-cleavage detected for both APP mut5 and C99 mut5 led us to investigate the effect of the GxxxG mutation on APP-PS1 interaction in co-immunoprecipitation experiments. PS1 is thought to harbor the catalytic core of the γ-secretase complex (9), and it has been shown to interact with APP C-terminal fragments (44). Our data clearly show that the GG 625/629 LL mutation did not impair the interaction between APP and full length PS1 or PS1 fragments that were reported to represent the catalytically active form of PS1 (Fig. 5).

Fig. 5.

Fig. 5

GG 625/629 LL mutation does not impair APP-PS1 interaction. The interaction between APP, APP mut5 and PS1 was studied in CHO cells co-expressing APP and human PS1. A. The levels of APP (top), PS1 holoprotein and PS1 N-terminal fragment (NTF, bottom) were monitored by Western blotting revealed with the WO-2 and MAB1563 antibodies, respectively. B. Cell lysates were immunoprecipitated by the WO-2 (human APP) antibody and revealed by Western blotting with the human PS1-specific MAB1563 antibody. Arrows indicate the expected position of cellular PS1 holoprotein, and PS1 N-terminal fragment.

Taken together, the complete inhibition of Aβ40 along with the strong decrease in Aβ42 production observed for APP mut5 and C99 mut5 led us to conclude that GG 625/629 LL mutation inhibits specifically the γ-cleavage that produces Aβ40 and Aβ42, but this was not due to a failure of APP-PS1 interaction.

GG 625/629 LL mutation does not change AICD release

The cleavage of APP C-terminal fragments by the γ-secretase complex also releases AICD, the APP Intracellular C-terminal Domain (18). Since GG 625/629 LL mutation decreases Aβ production, we studied its impact on AICD release. This was first analyzed by using APPGal4 fusion proteins. This system allows to measure the release of AICD in a quantitative and sensitive manner by a Gal4 transactivation assay (15;16). APP Gal4 and corresponding GG 625/629 LL mutant (APPGal4 mut5) were transiently expressed in CHO cells (Fig. 6A and 6B). We measured Aβ40 production in the culture medium of APPGal4-expressing cells by ELISA. We observed a strong decrease in Aβ production in extracellular media from cells expressing APPGal4 mut5 (Fig. 6C). This was in line with the results obtained from cells expressing APP695. Very interestingly, transactivation assays showed that the mutation of the GxxxG motif had no detectable effect on AICD release (Fig. 6D).

Fig. 6.

Fig. 6

GG 625/629 LL mutation does not affect AICD release. The release of AICD by ε-cleavage of APP was studied in CHO cells expressing either APP695 or APPGal4 fusion proteins. A. Schematic representation of the APPGal4 protein in comparison to APP695, along with the epitope recognized by the WO-2 and APP C-terminal antibody (C-ter). B. Expression of APP695, APPGal4 and APPGal4 mut5 was measured in transfected CHO cells by Western blotting revealed by the WO-2 antibody. C. Aβ1-40 production was monitored by ELISA in the culture medium of cells expressing APPGal4, APPGal4 mut5, and given as Aβ levels in pg/ml, nd = non-detectable (below 15 pg/ml). Values are means ± sem, n = 4; *** p < 0.001, compared to control. D. The release of AICD was measured by a Gal4 transactivation assay (bottom). Luciferase activity was normalized and represented as activity relative to control (APP695). Values are means ± sem, n = 4, ns = non-significant, *** p < 0.001, compared to control. E. Forty hours after transfection, CHO cells were treated for 8 h by 100μM of 1-10 Phenanthroline monohydrate (orthophenantroline, PNT), a metalloprotease inhibitor. AICD levels were measured in cell lysates (Western blotting) revealed by the C-ter antibody. Arrows indicate the expected position of αCTF and AICD. F. The same experiments were performed in cells expressing APP695 or APP mut5. AICD levels were measured in cell lysates (Western blotting) revealed by the C-ter antibody. Arrows indicate the expected position of APP, αCTF and AICD.

The intracellular AICD levels were next monitored by Western blotting in cells expressing C99 and C99 mut5. AICD is a very labile peptide that is rapidly degraded by intracellular metalloproteases and particularly by Insulin-Degrading Enzyme (45). The AICD levels were therefore measured in the presence of 100 μM orthophenantroline (PNT), a metalloprotease inhibitor (46). Results shown in Figure 6E indicate that the levels of AICD are similar in cells expressing C99 and C99 mut5. We carried out the same experiments in cells expressing APP695 and APP mutant 5, and showed that AICD levels were again similar in cells expressing APP and APP mut5 (Fig. 6F).

In summary, we have found that the glycine to leucine mutation decreases Aβ formation (Fig. 1C, 4C, 4E and 6C), whereas it has no effect on AICD release (Fig. 6D, 6E and 6F) even though AICD and Aβ are believed to be produced by γ-secretase activity. In addition, the interaction of APP with PS1 was not disrupted by the GG 625/629 LL mutation (Fig. 5). This led us to search for a possible role of the GxxxG motifs in homodimerization of APP C-terminal fragments, which could account for the observed effects on Aβ and AICD formation

GG 625/629 LL enhances the formation of C99 homodimers

One way for the GxxxG motifs to modulate Aβ production could be to promote close apposition of TM domains in APP, or especially in C99. We asked whether amyloidogenic processing simply requires dimerization, or whether the GxxxG motifs impose a specific interface of TM dimerization, which then promotes amyloidogenic processing. Since our GG 625/629 LL mutation inhibited amyloidogenic processing, we tested its effect on C99 dimerization. Unexpectedly, results shown in Fig. 4D indicate that the GG 625/629 LL mutation is likely to trigger the formation of SDS-resistant C99 dimers.

We employed HA- and Myc-tagged C99 constructs in order to be able to test by co-immunoprecipitation experiments whether C99 or mutants of C99 do oligomerize. These proteins contain the signal peptide (residues 1-19) of APP, fused to the tag (HA or Myc) followed by a linker corresponding to the 4 amino acids preceding the β-cleavage site and the C99 sequence (Fig. 7A). When expressed in CHO cells, these constructs produce Aβ40 and Aβ42 (Fig. 7B), indicating that they are correctly processed by β- and γ-secretase. Here again, for the tagged versions of C99, the GG 625/629 LL mutation completely blocked Aβ40 production and very strongly decreased Aβ42 (Fig. 7B). Western blotting on cell lysates revealed that the mutated HA- and Myc-C99 were forming high levels of SDS-resistant dimers when compared to the corresponding non-mutated proteins (Fig. 7C), which is in accordance with our results with non-tagged C99 and C99 mut5. These dimers where recognized specifically by both anti-tag (HA- and Myc-) and WO-2 antibodies.

Fig. 7.

Fig. 7

GG 625/629 LL mutation triggers the formation of C99 homodimers. A. Schematic representation of the HA- and Myc-tagged C99 proteins. The epitope recognized by the WO-2 antibody is underlined. B. Aβ1-40 and Aβ1-42 production was monitored by ELISA in the culture medium of cells expressing HA-C99, Myc-C99, or the corresponding mutants 5, and given as Aβ levels in pg/ml. Values are means ± sem, n = 4; *** p < 0.001, compared to control. C. The expression of HA- and Myc-C99 and HA- and Myc-C99 mut5 was measured by Western blotting in cell lysates revealed by the anti-HA (left), anti-Myc (middle) and WO-2 (right) antibodies. The presence of tagged C99 at the expected molecular weight is indicated by an arrow. The asterisk indicates the presence of possible C99 dimers. D. The formation of C99 homodimers was analyzed in CHO cell lines stably expressing HA-C99, Myc-C99 or the corresponding mutants 5. Levels of C99 was measured by Western blotting revealed by the anti-HA antibody in cell extracts prior to immunoprecipitation (Input, direct lysates). Extracts from cells expressing separately the HA-C99 and Myc-C99 were mixed and analyzed by Western blotting prior to immunoprecipitation (Input, post-lysate mix). Cell lysates were further immunoprecipitated by the anti-Myc and revealed by Western blotting with the anti-HA antibody as indicated. Arrows indicate the expected position of HA-C99 monomers (HA-C99) and HA-C99 dimers (HA-C99*).

Cell lysates were immunoprecipitated with the anti-Myc antibody and then subsequently analyzed by Western blotting with the anti-HA antibody. C99 proteins were detected by Western blotting with anti-HA antibodies only in cell lysates co-expressing HA-C99 and Myc-C99 constructs, and not in those expressing only HA-C99. This demonstrated the specificity of the antibody for the tag (Myc) used for immunoprecipitation. Bands were detected in Western blots at the expected molecular weight of tagged-C99 monomers. Their intensity was similar for C99 and C99 mut5 (containing the GG 625/629 LL mutation). Importantly, very high levels of C99 dimers were immunoprecipitated in cells co-expressing HA- and Myc-C99 mut5, whereas these dimers were present at low levels in cells expressing the non-mutated HA- and Myc-C99 proteins (Fig. 7D). The formation of C99 dimers did not result from the post-lysate aggregation of C99 proteins. When cell lysates separately expressing HA- and Myc-C99 were mixed and immunoprecipitated (Post-lysate mix lanes), no C99 dimer was detected (Fig. 7D). Altogether, our data show that the Gly-to-Leu mutation at the positions 625 and 629 promoted the formation of SDS-resistant C99 dimers, and thus promoted dimerization of the C-terminus domain of APP.

We further analyzed the dimerization state in relation to Aβ production for C99 proteins that carried a single G to L mutation in the G625xxxG629 motif or for G to I mutants that were recently described (22). The different mutants are depicted in Fig. 8A. The Gly-to-Leu mutation at the position 625 had weak effects on the formation of SDS-resistant C99 dimers, whereas G 629 L mutation lead to formation of similar levels of SDS-resistant C99 dimers as those observed for the GG 625/629 LL mutants (Fig 8B). Strikingly, the decrease in Aβ secretion (Aβ40 and Aβ42) was inversely proportional to the levels of C99-SDS resistant dimers measured for these mutants. In addition, identical results were obtained for glycine to isoleucine mutants, among which the very recently described G 629 I mutant (22) (Fig. 8B and 8C). There was no significant difference observed in AICD levels produced by any of these mutants (Fig. 8B). These results indicate that the GxxxG motif may impose a specific dimer interface, which is required for amyloiogenic processing. Mutations that rotate this interface, even if they would enhance dimerization, would impair amyloidogenic processing and Aβ formation. We have used computational searches of low energy dimer structures to assess the likely dimerization interface imposed by the GxxxG motifs and the possible influence of mutation of the glycine residues on TM helix dimerization and helix orientation within a dimer. For the wild-type APP sequence (residues 620-648), we typically observed a low energy structure (Fig. 9A) where helix dimerization was mediated by the GxxxG motifs involving glycines 621, 625, 629 and 633. In contrast, for the mut5 sequence (GG 625/629 LL) and the G 629 I sequence, the presence of leucines at positions 625 and 629 or isoleucine at position 629 led to a rotation of the helices and a low energy dimer structure with the small residues Gly634 and Ala638 in the dimer interface (Figs. 9B and 9C, respectively). Fig. 9D shows a plot of the interaction energies that stabilize the helix dimers in Figures 9A-C. The close proximity of the helices in the wild-type APP dimer allows Ser622 to form a stabilizing interhelical hydrogen bond. In the mut5 dimer, the most stabilizing interaction is an interhelical hydrogen bond formed by Asn623. Thus, the dimeric interfaces are very likely to differ between the wild type APP TM domain the APP mut5 TM domain.

Fig. 8.

Fig. 8

G to L and G to I mutations display similar effects on Aβ production, AICD release and C99 oligomers formation. A. Schematic representation of the transmembrane (TM) and juxtamembrane domains of the C99 G to L or G to I mutants. B. The expression of C99 and C99 mutants was measured by Western blotting in cell lysates revealed by the WO-2 antibody (Top). The presence of C99 is depicted by an arrow, the asterisk indicates the presence of C99 SDS-resistant dimers. The presence of AICD and αCTFs was measured in the same cell lysates by Wetern blotting revealed with the C-ter antibody (Bottom). D. Aβ1-40 and Aβ1-42 production was monitored by ELISA in the culture medium of cells expressing the indicated G to L or G to I C99 mutants. Results are given as Aβ levels in pg/ml. Values are means ± sem, n = 4; *** p < 0.001, compared to control.

Fig. 9.

Fig. 9

Dimer models for the wild-type APP transmembrane domain, the GG 625/629 LL and the G 629 I mutants. A. Low energy dimer of the TM domain of APP. The dimer interface is lined by glycines at positions 621, 625, 629 and 633. B-C. Low energy dimer of the TM domain of the GG 625/629 LL and G 629 I mutants of APP, respectively. Mutation of glycines 625 and 629 to leucine, or glycine 629 to isoleucine results in rotation of the transmembrane helices. C. Helix-helix interaction energies for the dimer structures for wild-type APP (solid line, full circles), the GG 625/629 LL mutant (dashed line, open circles) and G 629 I mutant (dashed line, open squares). The glycines in the interface of wild-type APP dimer allow Ser622 to form a strong interhelical hydrogen bond. In the GG/LL and G/I mutants, the Gly634xxxAla638 sequence allows close approach of the helices. In these mutants, interhelical hydrogen bonding of Asn623 provides the most stabilizing interaction.

Discussion

Our key finding is that the TM GxxxG motifs are required for Aβ40 and Aβ42 production from APP. The GxxxG motifs are not required for the generation of AICD, although both Aβ peptides and AICD are normally produced by a γ-secretase-mediated process. Our APP mut5, where the middle GxxxG motif is mutated to LxxxL, is indistinguishable from wild type APP with respect to AICD generation, but exhibits severely impaired generation of the Aβ40 and Aβ42 peptides. Since the leucine substitutions are not compatible with the close helix packing observed by the GxxxG motifs, our results demonstrate that helix dimerization and orientation differently affect processing of APP leading to Aβ or AICD production.

The GxxxG motif and ‘GxxxG-like’ motifs (in which one or both glycine residues are substituted by other small residues, such as alanine or serine), were found to be essential for mediating close interactions between TM α-helices, due to the very small size of glycine (47). Both alanines and glycines are small residues frequently found in the contact interfaces between TM helices (20), and AxxxA motifs have been reported to also promote interaction of TM α-helices (42).

The APP TM domain and extracellular juxtamembrane region, which are thought to adopt α-helical conformations, contain three in-register GxxxG motifs. It is unclear whether these motifs contribute to the dimerization of APP, which contains a bulky extracellular domain that has been reported to mediate APP-APP interactions (48). However, after ectodomain shedding (by β-cleavage for instance), when the extracellular portion is small, the TM domains are likely to play a major role in dimerization of βCTF (C99). To investigate the role of GxxxG motifs in APP processing, we mutated the glycine residues of each GxxxG motif either to alanine or leucine and found that the middle motif is crucial for the amyloidogenic processing of APP. Impairment of the amyloidogenic processing of APP in the GxxxG to LxxxL mutants might therefore be due to the disruption of close TM helical packing when glycines were substituted by bulky leucine residues. Indeed, amyloidogenic processing was normal when glycines were mutated to alanines.

Three lines of evidence indicated that the GxxxG mutation might affect the γ-cleavage of APP: (i) β-cleavage and APP-BACE1 interaction were not affected by GxxxG to LxxxL mutation (Fig. 3B, and 3C); (ii) higher levels of βCTFs were detected in APP mut5 in the context of unchanged β-cleavage (Fig. 3E); and (iii) a similar decrease in Aβ was observed for APP mut5 and C99 mut5, in which Aβ is produced by the single γ-cleavage (Fig. 4C and 4E). Thus, the replacement of glycine by leucine in the middle GxxxG motif (GG 625/629 LL) of APP or C99 leads to impaired γ-cleavage, which results in the inhibition of Aβ40 and Aβ42 production.

Since mutation of G625xxxG629 to LxxxL resulted in decreased γ-cleavage, it was conceivable that this mutation might impair the interaction of APP with the γ-secretase. The multicomponent γ-secretase complex contains at least four membrane proteins. The interaction of APP with PS1, the catalytic core of γ-secretase complex, was not altered as far as association is concerned (Fig. 5). It is not clear whether the relative positioning of APP towards PS1 was modified by the leucine mutations, and this should be further investigated.

Very importantly, we also showed that in the context of decreased Aβ production, the intracellular levels of AICD were not affected. Aβ and AICD production require a functional presenilin-dependent γ-secretase complex (18;49;50). However, Aβ is mainly produced by cleavage at the positions 40 and 42, whereas AICD results from the cleavage at position 49, distal to the γ-cleavage site (51). Thus, the cleavage at position 40 or 42 is now referred as γ-cleavage, whereas the cleavage at position 49 releasing AICD is referred to as ε-cleavage. The interrelation between γ- and ε-cleavage is a matter of debate. A previous study reported an equimolar production of Aβ and AICD by γ-secretase, suggesting a direct relationship between γ- and ε-cleavage (52). On the other hand, loss of a component (TMP21) associated to the γ-secretase complex results in an increase in Aβ, dependent on γ-cleavage of APP, without affecting the level of AICD, dependent on ε-cleavage (53). Moreover, it has been reported that PS1 mutations increase the production of Aβ42 but inhibit cleavage at the ε position (54). In the current study, we have found that the GG 625/629 LL mutation of APP results in a decrease in total Aβ without affecting the level of AICD. These results are completely in line with a very recent study showing that the extracellular/luminal juxtamembrane region of APP is an important regulatory domain that differentially regulates γ- and ε-cleavage (55). This point should be further investigated in order to understand how the GxxxG motifs contribute to proper positioning of APP for γ-cleavage, without influencing ε-cleavage.

APP was reported to form homodimers/homo-oligomers (32;48). The isolated TM sequence of APP and other γ-secretase substrates like Notch were shown to self-associate and form dimers in TOXCAT assays (56). A very recent study reported that the GxxxG motifs of APP are required for Aβ42 production and that the same motif supports homodimerization of the APP TM domain in isolation in TOXCAT assays in bacteria (22). Mutation of G629 to isoleucine impaired both dimerization in TOXCAT assays of a short sequence taken from the APP TM domain and the production of Aβ42. In contrast, our results demonstrate that the G629I mutant promotes homodimerization of C99 (Fig. 8B), and provide the first experimental evidence that C99 homodimers are present in transfected cells. These observations do not fit to a model in which GxxxG mutations weaken the dimerization of TM helices, thus facilitating the access of γ-secretase to the cleavage sites that are buried in the membrane (22). A very recent study might explain why our results about dimerization of APP TM domain are not in line with those previously reported (22). We analyzed homodimerization of the βCTF of APP (C99) by co-immunoprecipitation, whereas dimerization of APP TM domain was previously studied in TOXCAT assays (22;56). Importantly, in the TOXCAT assays, the dimerization of only the APP 29-42 TM sequence (Aβ numbering) was analyzed (22), and this sequence is lacking the first juxtamembrane GxxxG motif. As shown in Fig. 9, this motif contains a serine (Serine 622, APP 695 numbering) that forms the strongest interhelical contact in the calculated APP homodimer, and is thus likely to be a major determinant of APP dimerization. More to the point, the first APP GxxxG motif was shown to be critically involved in Aβ, but not in AICD, production (57), with the single mutation of Serine 622 to leucine drastically reducing Aβ production without affecting AICD release (57). These results agree with our observations and establish that the dimerization of the APP TM domain needs to be investigated in the context of the surrounding juxtamembrane region.

This raises the hypothesis that, although GxxxG motifs are required for amyloidogenic processing, enhancing dimerization of APP or C99 per se does not necessarily lead to an increase in the production of Aβ and particularly Aβ42 (22;32), since we show that increased dimerization of C99 is associated with a very strong decrease in Aβ40 and Aβ42. Dimerization that occurs after ectodomain shedding has also been shown to decrease presenilin-dependent intramembrane cleavage (58). Molecular dynamics simulations of the TM domain of APP or of the GG 625/629 LL and G 629 I mutants show that a different dimer interface can be adopted by the wild type APP TM helix versus GG 625/629 LL and G 629 I TM helices. All the Gly residues of the GxxxG motifs are predicted to be in the interface of the wild type TM dimer, while introduction of Leu or Ileu residues will cause a rotation and the placement of other small residues in the interface (i.e. G634 and A638). This conformation, which allows for strong and SDS-resistant dimerization, has no effect on ε-cleavage but deeply affects Aβ production. This prediction provides a new general model to explain our observations and results reported by other groups (22).

Taken together, these data lead to a comprehensive model for the role of the GxxxG motifs in APP on the pathogenesis of Alzheimer's disease. We propose that the sequence represented by Aβ when present in the context of full-length APP may adopt an α-helical structure. Once BACE has cleaved APP at the β position, the βCTF will likely assume a dimer conformation, with the GxxxG motifs in the interface, as these motifs are classical mediators of TM dimerization (27). The βCTF dimers will then bind to the presenilin complex(es) and will be processed to form Aβ peptides and AICD. For these processing events, two distinct models have been proposed. In the first, progressive cleavage occurs from the ε to the γ sites (59), with expected equimolar production of Aβ and AICD, as previously described (52). In the second, the γ and ε-secretase activities may be associated with different presenilin complexes, and PS1-dependent unraveling of the α-helical TM sequence occurring around positions 40 and 42 to generate the γ-cleavage. In this case, one can imagine situations where Aβ production can increase or decrease relative to the level of AICD. Once Aβ is generated, the GxxxG motifs would promote a conformational change from α-helix to β-strand, as we previously reported (21). Amyloid fibrils associated with Alzheimer's disease and a wide range of other neurodegenerative diseases have a cross β-sheet structure, where main chain hydrogen bonding occurs between β-strands in the direction of the fibril axis. We have shown that the packing of Met35 against Gly33 in the C-terminus of Aβ40 and against Gly37 in the C-terminus of Aβ42 (21) leads to the formation of strongly neurotoxic amyloid fibrils. Since a small fraction of these Aβ peptides may never leave the membrane lipid bilayer, they may also bind other GxxxG or AxxxA transmembrane proteins like APP or the γ-secretase complex as it has been suggested (19) and thus affect their function.

In summary, our present and previous data (21) indicate that the APP TM GxxxG motifs mediate close interactions between TM α-helices of APP and between β-sheets in fibrils formed by the Aβ40 and Aβ42 peptides, and also that glycine facilitates the α-helix to β-sheet conversion, which all are key features required for amyloid deposition. As a result, a motif initially discovered to promote dimerization of glycophorin A via self-association of its TM domain, may play a major role in Aβ production and pathogenic effects of APP processing in Alzheimer's disease (19). In addition, the finding that AICD release by ε-cleavage can be dissociated from Aβ production is also of particular interest in the physiopathological function of APP. AICD is thought to act as a regulator of gene transcription (16;17;60), and our data indicate that the genetic program controlled by the amyloidogenic processing of APP might be distinguished from Aβ production, a key event in AD.

Supplementary Material

Supplementary_Data

Acknowledgments

The authors thank L. Mercken for his help for sαAPP and sβAPP quantification, N. Sergeant and A. Delacourte for the APP C-ter antibodies and F. N'Kuli for excellent technical assistance. L.M. Munter and C. Weise are greatly acknowledged for mass spectrometry analysis. This work was supported by an Action de Recherche Concertée (ARC 03/08-299) from the French Community of Belgium (JNO and PJC), Interuniversity Attraction Poles Programme-Belgian Sate-Belgian Science Policy, the Belgian Fonds de Recherche Scientifique Médicale (JNO), the Queen Elisabeth Medical Foundation (JNO), NIH grant AG027317 (SOS), funds (to SNC) from the De Hovre Foundation, the Fonds Speciaux de Recherche, the Christian de Duve Institute of Cellular Pathology (ICP) and the F.N.R.S. Belgium. SNC is a Research Associate of the F.N.R.S. Belgium.

Footnotes

1

The abbreviations used are:APP, Amyloid Precursor Protein; Aβ, β-amyloid peptide; AD, Alzheimer's disease; CNS, Central Nervous System; TM, Transmembrane domain.

References

  • 1.Seubert P, Vigo-Pelfrey C, Esch F, Lee M, Dovey H, Davis D, Sinha S, Schlossmacher M, Whaley J, Swindlehurst C. Nature. 1992;359:325–327. doi: 10.1038/359325a0. [DOI] [PubMed] [Google Scholar]
  • 2.Shoji M, Golde TE, Ghiso J, Cheung TT, Estus S, Shaffer LM, Cai XD, McKay DM, Tintner R, Frangione B. Science. 1992;258:126–129. doi: 10.1126/science.1439760. [DOI] [PubMed] [Google Scholar]
  • 3.Kang J, Lemaire HG, Unterbeck A, Salbaum JM, Masters CL, Grzeschik KH, Multhaup G, Beyreuther K, Muller-Hill B. Nature. 1987;325:733–736. doi: 10.1038/325733a0. [DOI] [PubMed] [Google Scholar]
  • 4.Sinha S, Anderson JP, Barbour R, Basi GS, Caccavello R, Davis D, Doan M, Dovey HF, Frigon N, Hong J, Jacobson-Croak K, Jewett N, Keim P, Knops J, Lieberburg I, Power M, Tan H, Tatsuno G, Tung J, Schenk D, Seubert P, Suomensaari SM, Wang S, Walker D, John V. Nature. 1999;402:537–540. doi: 10.1038/990114. [DOI] [PubMed] [Google Scholar]
  • 5.Vassar R, Bennett BD, Babu-Khan S, Kahn S, Mendiaz EA, Denis P, Teplow DB, Ross S, Amarante P, Loeloff R, Luo Y, Fisher S, Fuller J, Edenson S, Lile J, Jarosinski MA, Biere AL, Curran E, Burgess T, Louis JC, Collins F, Treanor J, Rogers G, Citron M. Science. 1999;286:735–741. doi: 10.1126/science.286.5440.735. [DOI] [PubMed] [Google Scholar]
  • 6.Hussain I, Powell D, Howlett DR, Tew DG, Meek TD, Chapman C, Gloger IS, Murphy KE, Southan CD, Ryan DM, Smith TS, Simmons DL, Walsh FS, Dingwall C, Christie G. Mol Cell Neurosci. 1999;14:419–427. doi: 10.1006/mcne.1999.0811. [DOI] [PubMed] [Google Scholar]
  • 7.Yan R, Bienkowski MJ, Shuck ME, Miao H, Tory MC, Pauley AM, Brashier JR, Stratman NC, Mathews WR, Buhl AE, Carter DB, Tomasselli AG, Parodi LA, Heinrikson RL, Gurney ME. Nature. 1999;402:533–537. doi: 10.1038/990107. [DOI] [PubMed] [Google Scholar]
  • 8.Lin X, Koelsch G, Wu S, Downs D, Dashti A, Tang J. Proc Natl Acad Sci U S A. 2000;97:1456–1460. doi: 10.1073/pnas.97.4.1456. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Fortini ME. Nat Rev Mol Cell Biol. 2002;3:673–684. doi: 10.1038/nrm910. [DOI] [PubMed] [Google Scholar]
  • 10.Cupers P, Orlans I, Craessaerts K, Annaert W, De Strooper B. J Neurochem. 2001;78:1168–1178. doi: 10.1046/j.1471-4159.2001.00516.x. [DOI] [PubMed] [Google Scholar]
  • 11.Muresan Z, Muresan V. Hum Mol Genet. 2004;13:475–488. doi: 10.1093/hmg/ddh054. [DOI] [PubMed] [Google Scholar]
  • 12.Annaert W, De Strooper B. Trends Neurosci. 1999;22:439–443. doi: 10.1016/s0166-2236(99)01455-1. [DOI] [PubMed] [Google Scholar]
  • 13.Ebinu JO, Yankner BA. Neuron. 2002;34:499–502. doi: 10.1016/s0896-6273(02)00704-3. [DOI] [PubMed] [Google Scholar]
  • 14.Leissring MA, Murphy MP, Mead TR, Akbari Y, Sugarman MC, Jannatipour M, Anliker B, Muller U, Saftig P, De Strooper B, Wolfe MS, Golde TE, LaFerla FM. Proc Natl Acad Sci U S A. 2002;99:4697–4702. doi: 10.1073/pnas.072033799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Cao X, Sudhof TC. Science. 2001;293:115–120. doi: 10.1126/science.1058783. [DOI] [PubMed] [Google Scholar]
  • 16.Cao X, Sudhof TC. J Biol Chem. 2004;279:24601–24611. doi: 10.1074/jbc.M402248200. [DOI] [PubMed] [Google Scholar]
  • 17.Pardossi-Piquard R, Petit A, Kawarai T, Sunyach C, Alves dC, Vincent B, Ring S, D'Adamio L, Shen J, Muller U, St George HP, Checler F. Neuron. 2005;46:541–554. doi: 10.1016/j.neuron.2005.04.008. [DOI] [PubMed] [Google Scholar]
  • 18.Hebert SS, Serneels L, Tolia A, Craessaerts K, Derks C, Filippov MA, Muller U, De Strooper B. EMBO Rep. 2006;7:739–745. doi: 10.1038/sj.embor.7400704. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Marchesi VT. Proc Natl Acad Sci U S A. 2005;102:9093–9098. doi: 10.1073/pnas.0503181102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Liu W, Crocker E, Zhang W, Elliott JI, Luy B, Li H, Aimoto S, Smith SO. Biochemistry. 2005;44:3591–3597. doi: 10.1021/bi047827g. [DOI] [PubMed] [Google Scholar]
  • 21.Sato T, Kienlen-Campard P, Ahmed M, Liu W, Li H, Elliott JI, Aimoto S, Constantinescu SN, Octave JN, Smith SO. Biochemistry. 2006;45:5503–5516. doi: 10.1021/bi052485f. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Munter LM, Voigt P, Harmeier A, Kaden D, Gottschalk KE, Weise C, Pipkorn R, Schaefer M, Langosch D, Multhaup G. EMBO J. 2007;26:1702–1712. doi: 10.1038/sj.emboj.7601616. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Bormann BJ, Knowles WJ, Marchesi VT. J Biol Chem. 1989;264:4033–4037. [PubMed] [Google Scholar]
  • 24.Lemmon MA, Treutlein HR, Adams PD, Brunger AT, Engelman DM. Nat Struct Biol. 1994;1:157–163. doi: 10.1038/nsb0394-157. [DOI] [PubMed] [Google Scholar]
  • 25.Smith SO, Song D, Shekar S, Groesbeek M, Ziliox M, Aimoto S. Biochemistry. 2001;40:6553–6558. doi: 10.1021/bi010357v. [DOI] [PubMed] [Google Scholar]
  • 26.Russ WP, Engelman DM. J Mol Biol. 2000;296:911–919. doi: 10.1006/jmbi.1999.3489. [DOI] [PubMed] [Google Scholar]
  • 27.Eilers M, Patel AB, Liu W, Smith SO. Biophys J. 2002;82:2720–2736. doi: 10.1016/S0006-3495(02)75613-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Javadpour MM, Eilers M, Groesbeek M, Smith SO. Biophys J. 1999;77:1609–1618. doi: 10.1016/S0006-3495(99)77009-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Niimura M, Isoo N, Takasugi N, Tsuruoka M, Ui-Tei K, Saigo K, Morohashi Y, Tomita T, Iwatsubo T. J Biol Chem. 2005;280:12967–12975. doi: 10.1074/jbc.M409829200. [DOI] [PubMed] [Google Scholar]
  • 30.Lee SF, Shah S, Yu C, Wigley WC, Li H, Lim M, Pedersen K, Han W, Thomas P, Lundkvist J, Hao YH, Yu G. J Biol Chem. 2004;279:4144–4152. doi: 10.1074/jbc.M309745200. [DOI] [PubMed] [Google Scholar]
  • 31.Hendriks L, van Duijn CM, Cras P, Cruts M, Van Hul W, van Harskamp F, Warren A, McInnis MG, Antonarakis SE, Martin JJ. Nat Genet. 1992;1:218–221. doi: 10.1038/ng0692-218. [DOI] [PubMed] [Google Scholar]
  • 32.Scheuermann S, Hambsch B, Hesse L, Stumm J, Schmidt C, Beher D, Bayer TA, Beyreuther K, Multhaup G. J Biol Chem. 2001;276:33923–33929. doi: 10.1074/jbc.M105410200. [DOI] [PubMed] [Google Scholar]
  • 33.Macq AF, Czech C, Essalmani R, Brion JP, Maron A, Mercken L, Pradier L, Octave JN. J Biol Chem. 1998;273:28931–28936. doi: 10.1074/jbc.273.44.28931. [DOI] [PubMed] [Google Scholar]
  • 34.Sergeant N, David JP, Champain D, Ghestem A, Wattez A, Delacourte A. J Neurochem. 2002;81:663–672. doi: 10.1046/j.1471-4159.2002.00901.x. [DOI] [PubMed] [Google Scholar]
  • 35.Octave JN, Essalmani R, Tasiaux B, Menager J, Czech C, Mercken L. J Biol Chem. 2000;275:1525–1528. doi: 10.1074/jbc.275.3.1525. [DOI] [PubMed] [Google Scholar]
  • 36.Kienlen-Campard P, Miolet S, Tasiaux B, Octave JN. J Biol Chem. 2002;277:15666–15670. doi: 10.1074/jbc.M200887200. [DOI] [PubMed] [Google Scholar]
  • 37.Schagger H, von Jagow G. Anal Biochem. 1991;199:223–231. doi: 10.1016/0003-2697(91)90094-a. [DOI] [PubMed] [Google Scholar]
  • 38.Pierrot N, Ghisdal P, Caumont AS, Octave JN. J Neurochem. 2004;88:1140–1150. doi: 10.1046/j.1471-4159.2003.02227.x. [DOI] [PubMed] [Google Scholar]
  • 39.Leighton F, Poole B, Beaufay H, Baudhuin P, Coffey JW, Fowler S, De Duve C. J Cell Biol. 1968;37:482–513. doi: 10.1083/jcb.37.2.482. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Adams PD, Engelman DM, Brunger AT. Proteins. 1996;26:257–261. doi: 10.1002/(SICI)1097-0134(199611)26:3<257::AID-PROT2>3.0.CO;2-B. [DOI] [PubMed] [Google Scholar]
  • 41.Perez RG, Soriano S, Hayes JD, Ostaszewski B, Xia W, Selkoe DJ, Chen X, Stokin GB, Koo EH. J Biol Chem. 1999;274:18851–18856. doi: 10.1074/jbc.274.27.18851. [DOI] [PubMed] [Google Scholar]
  • 42.Kleiger G, Grothe R, Mallick P, Eisenberg D. Biochemistry. 2002;41:5990–5997. doi: 10.1021/bi0200763. [DOI] [PubMed] [Google Scholar]
  • 43.Pitsi D, Octave JN. J Biol Chem. 2004;279:25333–25338. doi: 10.1074/jbc.M312710200. [DOI] [PubMed] [Google Scholar]
  • 44.Pitsi D, Kienlen-Campard P, Octave JN. J Neurochem. 2002;83:390–399. doi: 10.1046/j.1471-4159.2002.01138.x. [DOI] [PubMed] [Google Scholar]
  • 45.Edbauer D, Willem M, Lammich S, Steiner H, Haass C. J Biol Chem. 2002;277:13389–13393. doi: 10.1074/jbc.M111571200. [DOI] [PubMed] [Google Scholar]
  • 46.Huysseune S, Kienlen-Campard P, Octave JN. Biochem Biophys Res Commun. 2007;361:317–322. doi: 10.1016/j.bbrc.2007.06.186. [DOI] [PubMed] [Google Scholar]
  • 47.Senes A, Engel DE, DeGrado WF. Curr Opin Struct Biol. 2004;14:465–479. doi: 10.1016/j.sbi.2004.07.007. [DOI] [PubMed] [Google Scholar]
  • 48.Soba P, Eggert S, Wagner K, Zentgraf H, Siehl K, Kreger S, Lower A, Langer A, Merdes G, Paro R, Masters CL, Muller U, Kins S, Beyreuther K. EMBO J. 2005;24:3624–3634. doi: 10.1038/sj.emboj.7600824. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Herreman A, Serneels L, Annaert W, Collen D, Schoonjans L, De Strooper B. Nat Cell Biol. 2000;2:461–462. doi: 10.1038/35017105. [DOI] [PubMed] [Google Scholar]
  • 50.Sastre M, Steiner H, Fuchs K, Capell A, Multhaup G, Condron MM, Teplow DB, Haass C. EMBO Rep. 2001;2:835–841. doi: 10.1093/embo-reports/kve180. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Weidemann A, Eggert S, Reinhard FB, Vogel M, Paliga K, Baier G, Masters CL, Beyreuther K, Evin G. Biochemistry. 2002;41:2825–2835. doi: 10.1021/bi015794o. [DOI] [PubMed] [Google Scholar]
  • 52.Kakuda N, Funamoto S, Yagishita S, Takami M, Osawa S, Dohmae N, Ihara Y. J Biol Chem. 2006;281:14776–14786. doi: 10.1074/jbc.M513453200. [DOI] [PubMed] [Google Scholar]
  • 53.Chen F, Hasegawa H, Schmitt-Ulms G, Kawarai T, Bohm C, Katayama T, Gu Y, Sanjo N, Glista M, Rogaeva E, Wakutani Y, Pardossi-Piquard R, Ruan X, Tandon A, Checler F, Marambaud P, Hansen K, Westaway D, George-Hyslop P, Fraser P. Nature. 2006;440:1208–1212. doi: 10.1038/nature04667. [DOI] [PubMed] [Google Scholar]
  • 54.Chen F, Gu Y, Hasegawa H, Ruan X, Arawaka S, Fraser P, Westaway D, Mount H, George-Hyslop P. J Biol Chem. 2002;277:36521–36526. doi: 10.1074/jbc.M205093200. [DOI] [PubMed] [Google Scholar]
  • 55.Ren Z, Schenk D, Basi GS, Shapiro IP. J Biol Chem. 2007;282:35350–35360. doi: 10.1074/jbc.M702739200. [DOI] [PubMed] [Google Scholar]
  • 56.Vooijs M, Schroeter EH, Pan Y, Blandford M, Kopan R. J Biol Chem. 2004;279:50864–50873. doi: 10.1074/jbc.M409430200. [DOI] [PubMed] [Google Scholar]
  • 57.Ren Z, Schenk D, Basi GS, Shapiro IP. J Biol Chem. 2007 doi: 10.1074/jbc.M702739200. [DOI] [PubMed] [Google Scholar]
  • 58.Struhl G, Adachi A. Mol Cell. 2000;6:625–636. doi: 10.1016/s1097-2765(00)00061-7. [DOI] [PubMed] [Google Scholar]
  • 59.De Strooper B. EMBO Rep. 2007;8:141–146. doi: 10.1038/sj.embor.7400897. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Alves dC, Sunyach C, Pardossi-Piquard R, Sevalle J, Vincent B, Boyer N, Kawarai T, Girardot N, George-Hyslop P, Checler F. J Neurosci. 2006;26:6377–6385. doi: 10.1523/JNEUROSCI.0651-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary_Data

RESOURCES