+ All Categories
Home > Documents > OligosaccharideBindingtoBarley -Amylase1 · OligosaccharideBindingtoBarley -Amylase1*...

OligosaccharideBindingtoBarley -Amylase1 · OligosaccharideBindingtoBarley -Amylase1*...

Date post: 11-Sep-2018
Category:
Upload: vanthien
View: 213 times
Download: 0 times
Share this document with a friend
12
Oligosaccharide Binding to Barley -Amylase 1 * Received for publication, May 19, 2005, and in revised form, July 6, 2005 Published, JBC Papers in Press, July 19, 2005, DOI 10.1074/jbc.M505515200 Xavier Robert , Richard Haser , Haruhide Mori §1 , Birte Svensson §¶ , and Nushin Aghajari ‡2 From the Laboratoire de BioCristallographie, IFR-128 “BioSciences Lyon-Gerland,” Institut de Biologie et Chimie des Prote ´ines, UMR 5086, CNRS-UCBL1, 7 Passage du Vercors, F-69367 Lyon Cedex 07, France, the § Department of Chemistry, Carlsberg Laboratory, Gamle Carlsberg Vej 10, DK-2500 Copenhagen Valby, Denmark, and the Biochemistry and Nutrition Group, BioCentrum-DTU, Technical University of Denmark, Søltofts Plads, Building 224, DK-2800 Kgs. Lyngby, Denmark Enzymatic subsite mapping earlier predicted 10 binding subsites in the active site substrate binding cleft of barley -amylase isozymes. The three-dimensional structures of the oligosaccharide complexes with barley -amylase isozyme 1 (AMY1) described here give for the first time a thorough insight into the substrate binding by describing residues defining 9 subsites, namely 7 through 2. These structures support that the pseudotetrasaccharide inhibitor acarbose is hydrolyzed by the active enzymes. Moreover, sugar binding was observed to the starch granule-binding site previously determined in barley -amylase isozyme 2 (AMY2), and the sugar binding modes are compared between the two isozymes. The “sugar tongs” surface binding site discovered in the AMY1-thio-DP4 com- plex is confirmed in the present work. A site that putatively serves as an entrance for the substrate to the active site was proposed at the glycone part of the binding cleft, and the crystal structures of the catalytic nucleophile mutant (AMY1 D180A ) complexed with acar- bose and maltoheptaose, respectively, suggest an additional role for the nucleophile in the stabilization of the Michaelis complex. Fur- thermore, probable roles are outlined for the surface binding sites. Our data support a model in which the two surface sites in AMY1 can interact with amylose chains in their naturally folded form. Because of the specificities of these two sites, they may locate/orient the enzyme in order to facilitate access to the active site for polysac- charide chains. Moreover, the sugar tongs surface site could also perform the unraveling of amylose chains, with the aid of Tyr-380 acting as “molecular tweezers.” Structural studies of -amylases in complex with substrate analogues have received much attention in the past decade (1–13). The pseudotet- rasaccharide inhibitor acarbose was used for the vast majority of these studies. In contrast, binding of natural substrates was only described in three complexes known to date. These are the -amylase from porcine pancreas complexed with maltopentaose (14, 15), the inactive catalytic mutant E208Q -amylase from Bacillus subtilis in complex with mal- topentaose (16), and the “maltogenic” -amylase from Bacillus stearo- thermophilus in complex with a maltose unit that was derived from a maltotriose (17). -Amylases belong to the glycoside hydrolase family 13 (afmb.cnrs-mrs.fr/CAZY), which together with GH70 and GH77 constitute GH clan H (18). These enzymes share the catalytic site geom- etry with three invariant acid residues as follows: a catalytic nucleophile, a proton donor, and a third catalytic acidic side chain, which polarizes the glucoside unit at subsite 1 (19). Despite a sequence identity of 80% and a nearly identical overall folding (20), AMY1 and AMY2 exhibit important differences as concerns stability (21, 22), enzymatic proper- ties (23), and sensitivity to the proteinaceous inhibitor barley -amylase subtilisin inhibitor, belonging to the Kunitz soybean trypsin inhibitor family (24). Most interestingly, three calcium ions are bound to both isozymes sharing identical ligands (20), although calcium affects the activity of AMY1 3 and AMY2 in distinct manners (21). Numerous AMY1 subsite mutants have been characterized with respect to sub- strate affinity and catalytic capacity, but structural insight for barley -amylases was limited until now to subsites 1 through 2 being experimentally defined with the AMY2-acarbose complex (4). Enzy- matic subsite mapping (25) and computer-aided modeling (26, 27) sug- gested that both AMY1 and AMY2 possess 10 subsites spanning from the nonreducing end at subsite 6 to 4 on the aglycon accommodat- ing part. A more recent study in which crystals of AMY1 soaked in a solution containing thio-maltotetraose was expected to expand our knowledge on subsite-binding modes in the catalytic cleft of plant -amylases. This substrate analogue, however, led to the discovery of a new surface binding site at domain C (see Fig. 1), the so-called “pair of sugar tongs,” but did not bind to the active site (28). This site and an earlier discovered surface binding site on the catalytic domain made up of two consecutive tryptophan residues, Trp-278 and Trp-279 (AMY1 numbering), perform distinct interactions with the sugar rings of differ- ent oligosaccharide ligands (4, 28). Mutational analyses have confirmed that these sites indeed can bind onto starch granules and that they also bind -cyclodextrin (29, 62). Here we report the crystal structures of native AMY1 as well as an inactive mutant of the catalytic nucleophile Asp-180 in complex with acarbose (Fig. 2, A and B) and maltoheptaose (Fig. 1), respectively. This is the first report on sugar binding to the active site of barley -amylase 1 and on binding of the substrate malto- heptaose to the active site region. EXPERIMENTAL PROCEDURES Preparation of Recombinant AMY19—A truncated form, AMY19 (nonapeptide deletion from the C terminus), of recombinant AMY1 was prepared as described previously (30) to overcome difficulties encoun- tered in growing three-dimensional crystals of full-length AMY1 (30). AMY19 is henceforth referred to as AMY1. * This work was supported by the European Union Framework IV and V Programmes Alpha-Glucan Active Designer Enzymes Grant BIO4-980022, Combinatorial Engineer- ing of Glycoside Hydrolases from the -Amylase Superfamily Grant QLK3-CT-2001- 00149, CNRS, and travel support from Institut Franc ¸ais, Copenhagen. The costs of publication of this article were defrayed in part by the payment of page charges. This article must therefore be hereby marked “advertisement” in accordance with 18 U.S.C. Section 1734 solely to indicate this fact. The atomic coordinates and structure factors (codes 1RPK, 1RP8, and 1RP9) have been depos- ited in the Protein Data Bank, Research Collaboratory for Structural Bioinformatics, Rut- gers University, New Brunswick, NJ (http://www.rcsb.org/). 1 Present address: Division of Applied Bioscience, Graduate School of Agriculture, Hok- kaido University, Sapporo 060-8589, Japan. 2 To whom correspondence should be addressed: Laboratoire de BioCristallographie, Institut de Biologie et Chimie des Prote ´ ines, UMR 5086, CNRS-UCBL1, 7 Passage du Vercors, F-69367 Lyon, France. Tel.: 33-4-72-72-26-33; Fax: 33-4-72-72-26-16; E-mail: [email protected]. 3 The abbreviations used are: AMY1, barley -amylase isozyme 1; AMY2, barley - amylase isozyme 2; MES, 4-morpholineethanesulfonic acid; CGTase, cyclomaltodex- trin glucanotransferases; G7, maltoheptaose; G8, malto-octaose. THE JOURNAL OF BIOLOGICAL CHEMISTRY VOL. 280, NO. 38, pp. 32968 –32978, September 23, 2005 © 2005 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in the U.S.A. 32968 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 280 • NUMBER 38 • SEPTEMBER 23, 2005 by guest on September 11, 2018 http://www.jbc.org/ Downloaded from
Transcript

Oligosaccharide Binding to Barley �-Amylase 1*

Received for publication, May 19, 2005, and in revised form, July 6, 2005 Published, JBC Papers in Press, July 19, 2005, DOI 10.1074/jbc.M505515200

Xavier Robert‡, Richard Haser‡, Haruhide Mori§1, Birte Svensson§¶, and Nushin Aghajari‡2

From the ‡Laboratoire de BioCristallographie, IFR-128 “BioSciences Lyon-Gerland,” Institut de Biologie et Chimie des Proteines,UMR 5086, CNRS-UCBL1, 7 Passage du Vercors, F-69367 Lyon Cedex 07, France, the §Department of Chemistry, CarlsbergLaboratory, Gamle Carlsberg Vej 10, DK-2500 Copenhagen Valby, Denmark, and the ¶Biochemistry and Nutrition Group,BioCentrum-DTU, Technical University of Denmark, Søltofts Plads, Building 224, DK-2800 Kgs. Lyngby, Denmark

Enzymatic subsite mapping earlier predicted 10 binding subsitesin the active site substrate binding cleft of barley �-amylaseisozymes. The three-dimensional structures of the oligosaccharidecomplexes with barley�-amylase isozyme 1 (AMY1) described heregive for the first time a thorough insight into the substrate bindingby describing residues defining 9 subsites, namely �7 through �2.These structures support that the pseudotetrasaccharide inhibitoracarbose is hydrolyzed by the active enzymes. Moreover, sugarbinding was observed to the starch granule-binding site previouslydetermined in barley �-amylase isozyme 2 (AMY2), and the sugarbindingmodes are compared between the two isozymes. The “sugartongs” surface binding site discovered in the AMY1-thio-DP4 com-plex is confirmed in the presentwork.A site that putatively serves asan entrance for the substrate to the active site was proposed at theglycone part of the binding cleft, and the crystal structures of thecatalytic nucleophile mutant (AMY1D180A) complexed with acar-bose andmaltoheptaose, respectively, suggest an additional role forthe nucleophile in the stabilization of the Michaelis complex. Fur-thermore, probable roles are outlined for the surface binding sites.Our data support a model in which the two surface sites in AMY1can interact with amylose chains in their naturally folded form.Because of the specificities of these two sites, theymay locate/orientthe enzyme in order to facilitate access to the active site for polysac-charide chains. Moreover, the sugar tongs surface site could alsoperform the unraveling of amylose chains, with the aid of Tyr-380acting as “molecular tweezers.”

Structural studies of �-amylases in complex with substrate analogueshave receivedmuch attention in the past decade (1–13). The pseudotet-rasaccharide inhibitor acarbose was used for the vast majority of thesestudies. In contrast, binding of natural substrates was only described inthree complexes known to date. These are the �-amylase from porcinepancreas complexed with maltopentaose (14, 15), the inactive catalyticmutant E208Q �-amylase from Bacillus subtilis in complex with mal-topentaose (16), and the “maltogenic” �-amylase from Bacillus stearo-

thermophilus in complex with a maltose unit that was derived from amaltotriose (17). �-Amylases belong to the glycoside hydrolase family13 (afmb.cnrs-mrs.fr/CAZY), which together with GH70 and GH77constitute GH clanH (18). These enzymes share the catalytic site geom-etry with three invariant acid residues as follows: a catalytic nucleophile,a proton donor, and a third catalytic acidic side chain, which polarizesthe glucoside unit at subsite�1 (19). Despite a sequence identity of 80%and a nearly identical overall folding (20), AMY1 and AMY2 exhibitimportant differences as concerns stability (21, 22), enzymatic proper-ties (23), and sensitivity to the proteinaceous inhibitor barley �-amylasesubtilisin inhibitor, belonging to the Kunitz soybean trypsin inhibitorfamily (24). Most interestingly, three calcium ions are bound to bothisozymes sharing identical ligands (20), although calcium affects theactivity of AMY13 and AMY2 in distinct manners (21). NumerousAMY1 subsite mutants have been characterized with respect to sub-strate affinity and catalytic capacity, but structural insight for barley�-amylases was limited until now to subsites �1 through �2 beingexperimentally defined with the AMY2-acarbose complex (4). Enzy-matic subsite mapping (25) and computer-aided modeling (26, 27) sug-gested that both AMY1 and AMY2 possess 10 subsites spanning fromthe nonreducing end at subsite �6 to �4 on the aglycon accommodat-ing part. A more recent study in which crystals of AMY1 soaked in asolution containing thio-maltotetraose was expected to expand ourknowledge on subsite-binding modes in the catalytic cleft of plant�-amylases. This substrate analogue, however, led to the discovery of anew surface binding site at domain C (see Fig. 1), the so-called “pair ofsugar tongs,” but did not bind to the active site (28). This site and anearlier discovered surface binding site on the catalytic domain made upof two consecutive tryptophan residues, Trp-278 and Trp-279 (AMY1numbering), perform distinct interactions with the sugar rings of differ-ent oligosaccharide ligands (4, 28). Mutational analyses have confirmedthat these sites indeed can bind onto starch granules and that they alsobind �-cyclodextrin (29, 62). Here we report the crystal structures ofnative AMY1 as well as an inactive mutant of the catalytic nucleophileAsp-180 in complex with acarbose (Fig. 2, A and B) and maltoheptaose(Fig. 1), respectively. This is the first report on sugar binding to theactive site of barley �-amylase 1 and on binding of the substrate malto-heptaose to the active site region.

EXPERIMENTAL PROCEDURES

Preparation of Recombinant AMY1�9—A truncated form, AMY1�9(nonapeptide deletion from theC terminus), of recombinantAMY1wasprepared as described previously (30) to overcome difficulties encoun-tered in growing three-dimensional crystals of full-length AMY1 (30).AMY1�9 is henceforth referred to as AMY1.

* This work was supported by the European Union Framework IV and V ProgrammesAlpha-Glucan Active Designer Enzymes Grant BIO4-980022, Combinatorial Engineer-ing of Glycoside Hydrolases from the �-Amylase Superfamily Grant QLK3-CT-2001-00149, CNRS, and travel support from Institut Francais, Copenhagen. The costs ofpublication of this article were defrayed in part by the payment of page charges. Thisarticle must therefore be hereby marked “advertisement” in accordance with 18 U.S.C.Section 1734 solely to indicate this fact.

The atomic coordinates and structure factors (codes 1RPK, 1RP8, and 1RP9) have been depos-ited in the Protein Data Bank, Research Collaboratory for Structural Bioinformatics, Rut-gers University, New Brunswick, NJ (http://www.rcsb.org/).

1 Present address: Division of Applied Bioscience, Graduate School of Agriculture, Hok-kaido University, Sapporo 060-8589, Japan.

2 To whom correspondence should be addressed: Laboratoire de BioCristallographie,Institut de Biologie et Chimie des Proteines, UMR 5086, CNRS-UCBL1, 7 Passage duVercors, F-69367 Lyon, France. Tel.: 33-4-72-72-26-33; Fax: 33-4-72-72-26-16; E-mail:[email protected].

3 The abbreviations used are: AMY1, barley �-amylase isozyme 1; AMY2, barley �-amylase isozyme 2; MES, 4-morpholineethanesulfonic acid; CGTase, cyclomaltodex-trin glucanotransferases; G7, maltoheptaose; G8, malto-octaose.

THE JOURNAL OF BIOLOGICAL CHEMISTRY VOL. 280, NO. 38, pp. 32968 –32978, September 23, 2005© 2005 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in the U.S.A.

32968 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 280 • NUMBER 38 • SEPTEMBER 23, 2005

by guest on September 11, 2018

http://ww

w.jbc.org/

Dow

nloaded from

Preparation of the Inactive Mutant AMY1D180A—Escherichia coliDH5� (Invitrogen) was used for standard cloning. Pichia pastorisGS115 and pPICZA (Invitrogen), carrying the ZeocinR selectionmarker, were used for expression. AMY1 cDNA was from an in-housecollection (31). pPICZA-AMY1 harbored the AMY1-encoding insertflanked by EcoRI and KpnI sites (32). Standard culture media were usedfor E. coli (33) and P. pastoris (31).

Construction, Site-directed Mutagenesis, Transformation, andScreening—The D180A mutation was introduced into pPICZA-AMY1(32) by themegaprimer procedure (34) as done for other mutations (32,35). AMY1 cDNA was amplified using primer A (32) and the mutagen-esis primer C (5�-CCCCTAGCGAAAGCAAGGCGCCACG-3�, posi-tions 695–719, antisense orientation; mutant anticodon double-under-lined) to generate the megaprimer used in the second PCR with primerB (32). All PCRs were performed using high fidelity Pfu DNA polymer-ase (Stratagene), and the complete sequence of pPICZA-AMY1D180Awas confirmed (Applied Biosystems 377 DNA Sequencer and Taq

DyeDeoxy Terminator Cycle Sequencing kit, PerkinElmer Life Sci-ences). The plasmid was linearized at the BstXI site for P. pastoris trans-formation (31). Transformants were screened on YPDS/Zeocin (100�g � ml�1) plates, transferred to an MDH plate, grown in 3 ml ofbuffered minimal glycerol complex (BMGY) at 30 °C for 1 day, centri-fuged (1,500 � g, 8 min, room temperature), and resuspended (3 ml ofBMGY containing 0.5 % methanol but without glycerol (BMMY)) forinduction with vigorous shaking for 2 days. SDS-PAGE and silver stain-ing (PhastSystem, Amersham Biosciences) of supernatants confirmedthe secretion of AMY1D180A.

Production and Purification of AMY1D180A—The selected transfor-mant was grown (0.5 liters of BMGY, 2 days in a 5-liter flask) to A600 �20, and the medium was replaced by BMMY (1 liter) for induction for32 h under vigorous shaking. The supernatant was kept, and the cellswere resuspended in BMMY (1 liter) for a second induction culture (39h). AMY1D180Awas purified from the combined supernatants by affinitychromatography on �-cyclodextrin-Sepharose (31, 32) and anion

FIGURE 1. Stereo drawing of the overall struc-ture of AMY1D180A inactive mutant in complexwith maltoheptaose. Calcium ions are indicatedas green spheres. Catalytic residues are highlightedin pink. The full maltoheptaose molecule bound inthe catalytic cleft is represented by a gray transpar-ent surface. To the left, the starch granule-bindingsurface site (occupied by a maltopentaose mole-cule) is shown with the two tryptophan residues(Trp-278 and Trp-279) highlighted in blue. Indomain C (bottom part of the figure), another mal-topentaose molecule is curved around Tyr-380 (inblue) at the level of the sugar tongs surface site.

FIGURE 2. A, atoms labeling convention for glu-cosyl residues and acarbose. The acarviosine unitis constituted by rings A and B, including theamino group valienamine, and are �-1,4-linked torings C and D representing a maltose unit. B,superimposition of residues implicated in the sub-strate binding in active sites of AMY1 and AMY2(stereo view). Only residues interacting directly byhydrogen bonds are shown. The complex AMY2-acarbose (4) (Protein Data Bank entry 1BG9) is pre-sented in blue, AMY1-acarbose (this work, ProteinData Bank entry 1RPK) in red, AMY1-thio-DP4 (28)(Protein Data Bank entry 1P6W) in green, andAMY1D180A-acarbose (this work, Protein Data Bankentry 1RP9) in yellow.

Barley �-Amylase Oligosaccharide Complexes

SEPTEMBER 23, 2005 • VOLUME 280 • NUMBER 38 JOURNAL OF BIOLOGICAL CHEMISTRY 32969

by guest on September 11, 2018

http://ww

w.jbc.org/

Dow

nloaded from

exchange chromatography (32). The enzyme was concentrated (Cen-triprep YM10, Millipore, Bedford, MA) to �2 mg � ml�1 and stored in10 mM MES, 5 mM CaCl2, pH 6.8, 0.02% (w/v) sodium azide at 4 °C.Isoelectric focusing (pI 4–6.5) and SDS-PAGE (PhastGels and Phast-System, Amersham Biosciences) of AMY1D180A were silver-stainedaccording to the manufacturer’s recommendation (36). Enzyme con-centrationswere calculated from amino acid contents of protein (25�g)hydrolysate (32). Purified AMY1D180A was obtained (6.5 mg � liter�1)and gave one band in SDS-PAGE (45 kDa) and native PAGE (data notshown).

Enzyme Activity Assays—Activity toward soluble potato starch(Sigma), 9mg�ml�1 in 20mM sodium acetate buffer, 5mMCaCl2, 0.05mg � ml�1 bovine serum albumin, pH 5.5, was measured forAMY1D180A (4 �M) at 37 °C and incubated for 45 s, 10 min, 30 min, and2 h (140 �l) using the copper bicinchoninate procedure (35, 37) andmaltose as standard. Wild-type AMY1 (3 nM) was analyzed under thesame conditions (10 min).

Crystallization, Soaking, and Cryo-protection—The complexbetween AMY1 and the inhibitor acarbose was obtained by co-crystal-lization at conditions derived from those of the native enzyme (30). 0.5�l of a 100 mM acarbose stock solution was mixed with 2 �l of proteinstock and 3 �l of well solution, thus resulting in a final concentration of9.1 mM acarbose in the drop, and crystals grew to a final size of 0.35 �0.1 � 0.05 mm3 within 3 weeks.

The inactive AMY1D180A mutant was crystallized by the hangingdrop vapor diffusion method at 19 °C, using protein stock (2.9 mg �ml�1) in 10mMMES, pH 6.7, 5mMCaCl2, and 0.02%NaN3. Drops wereprepared bymixing 2�l of protein stockwith 0.5�l of 3% (v/v) isopropylalcohol and 2.5 �l of well solution containing 20% (w/v) polyethyleneglycol 8000. Thin crystals typically grew to a final size of 0.8� 0.1� 0.05mm3 after �6 months. Soaking was done by adding a solution of mal-toheptaose (Roche Applied Science) directly to the drop to a final con-centration of 10 mM and leaving it for 24 h at 19 °C.The AMY1D180A-acarbose complex was obtained by adding acarbose

directly to the drop at a final concentration of 10 mM and leaving it for24 h at 19 °C. All crystals were cryo-protected prior to data collection by

rapid soaking in three successive steps in mother liquor containingincreasing concentrations of ethylene glycol (5, 10, and 15% (v/v)) and10 mM acarbose or maltoheptaose, respectively.

Diffraction Data Collection—AMY1-acarbose diffraction data werecollected on a MARresearch 345 Image Plate System associated to aNonius FR591 rotating anode (CuK� radiation), operating at 44 kV and100 mA and coupled to Osmic confocal mirrors. Data processing andreduction was carried out using DENZO (38) and SCALA from theCCP4 package (39).AMY1D180A-maltoheptaose data were collected at the ID14-1 beam-

line at the European Synchrotron Radiation Facility, Grenoble, France,on a MarCCD detector, and data on the AMY1D180A-acarbose werecollected at the FIP BM30Abeamline (European Synchrotron RadiationFacility) on a MarCCD detector. Diffracted intensities were integratedwith the programMOSFLM (40) as implemented in the CCP4 softwarepackage (39) and scaled with SCALA (39). All crystals belong to theorthorhombic space group P21212, and one molecule is present in theasymmetric unit. Data collection statistics are presented in TABLEONE.

Structure Determination and Refinement—Because of crystal iso-morphism with AMY1 (28), the latter was used as starting model (Pro-tein Data Bank entry 1HT6) in a difference Fourier, where all watermolecules, calcium ions, and ligand molecules had been removed. Forall data sets, an initial rigid body refinement, including data to 4 Åresolution, was performed, and in the remaining refinements a simu-lated annealing protocol was used extending the data to 2.0 Å resolu-tion. All refinements were done with the programCNS (41). In order toavoid over-refinement, free and conventionalR-factors weremonitored(42). Alternating with these refinement steps, visual examination ofelectron density maps and manual building was carried out using thegraphic software TURBO-FRODO (43). Based on the inspection of 2Fo� Fc and Fo � Fc maps (contoured at 1 and 3�, respectively), calciumions were inserted, and water molecules were manually added respect-ing hydrogen bonds with standard distances and angles formed toappropriate atoms. Ligands (sugarmoieties from acarbose ormaltohep-taose) were manually inserted in the electron density maps. Because of

TABLE ONE

Data collection and refinement statistics for AMY1-acarbose, AMY1D180A-acarbose, and AMY1D180A-maltoheptaose complexes

EnzymeAMY1-acarbose AMY1D180A-acarbose AMY1D180A-malthoheptaose

Protein data bank entry code 1RPK 1RP9 1RP8Wavelength (Å) 1.5418 0.9500 0.9340Data collection temperature (K) 100 100 100Unit cell dimensions, a, b, c (Å) 93.4, 72.9, 61.4 93.7, 73.5, 61.1 93.0, 72.5, 62.2Space group P21212 P21212 P21212Resolution range (Å) 17.4 to 2.0 39.5 to 2.0 39.1 to 2.0Completeness of data (%) 94.2 (94.7) 88.0 (88.9) 99.9 (100.0)Multiplicity of data (overall) 3.5 (3.1) 4.4 (4.4) 7.5 (7.3)Outermost shell (Å) 2.05 to 2.00 2.05 to 2.00 2.05 to 2.00Total no. reflections 96,013 114,564 211,756No. unique reflections 27,344 25,774 29,146Rsyma (%) 18.3 (36.6) 8.3 (31.5) 9.5 (31.1)Overall I/�(I) 4.2 7.5 7.5Rfactorb (%) 18.0 18.3 17.2Rfreec (%) 23.1 23.1 22.1Root mean square deviation from ideal geometryBonds lengths (Å) 0.010 0.013 0.010Bond angles (°) 1.50 1.60 1.40

a Rsym � (�hkl�i�I(hkl)i � I(hkl)�/�hkl�iI(hkl)).b R � (�h�Fobs(h)� � k�Fcalc(h)�)/(�h�Fobs(h)�) with k, a scaling factor.c Rfree � (�h�Fobs(h)� � k�Fcalc(h)�)/(�h�Fobs(h)�) with k, a scaling factor. Rfree is calculated from a test set constituted by 10% of the total number of reflections randomly selected.

Barley �-Amylase Oligosaccharide Complexes

32970 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 280 • NUMBER 38 • SEPTEMBER 23, 2005

by guest on September 11, 2018

http://ww

w.jbc.org/

Dow

nloaded from

the high quality and resolution of the diffraction data, the orientationand nature of each sugar ring was unequivocally determined. Models ofacarbose and glucose units were found in the HIC-Up data base (44).The geometry and coordinate errors of the three final structures wereexamined with the programs PROCHECK (45) and WHATCHECK(46). Refinement statistics are listed in TABLE ONE.Structural superimposition of the structures was performed using the

“rigid” option in TURBO-FRODO (43). Figures shown in this paperwere rendered using ISIS-Draw (freeware from Elsevier MDL, SanLeandro, CA; www.mdli.com), TURBO-FRODO (43), and VIEWER-LITETM 4.2 (freeware from Accelrys Inc., San Diego; www.accelry-s.com).

RESULTS

AMY1 in Complex with Acarbose and Comparison with AMY2-Acar-bose—The three-dimensional structure of the complex between AMY1and acarbose was solved to 2.0 Å resolution. As for AMY2 (4), threesubstrate binding subsites are occupied by amolecule of component 2 (apseudotrisaccharide derived from acarbose after cleavage of a glucoseunit from the reducing end) meaning that unit A to C occupies subsites�1 to �2. The starch granule-binding surface site containing Trp-278and Trp-279 also binds a component 2 molecule, whereas the “sugar

tongs” site at domain C (28) displays the entire acarbose molecule.Direct hydrogen bonds between enzyme and inhibitor molecules arelisted for the three sites in TABLE TWO.The electron density map clearly suggests that acarbose was cleaved

byAMY1 resulting in a component 2molecule binding to the active site.At the starch granule-binding surface site, the fourth sugar ring is notdefined, reflecting either that this sugar partly points into the solventand therefore is flexible or a difference in affinity for this and the sugartongs-binding site, respectively. This conclusion is supported by thestructures, the surface site with Trp-278 to Trp-279 being widelyopened and exposed to solvent. Only a relatively low number of inter-actions are present for stabilizing long flexible substrates; therefore, ashort and less flexible substrate may have a higher affinity. In contrast,the sugar tongs site is defined by a cleft in which oligosaccharide iscaptured by Tyr-380 and has an increased number of interactions com-pared with the former surface site, thus a priori providing superiorstabilization for longer ligands. Remarkably, when comparing the activesites for AMY1-acarbose and AMY2-acarbose, structural conservationbetween the enzymes is high as only two amino acid side chains differbetween the isozymes (see Fig. 2B). Arg-183AMY1 replaces Lys-182AMY2and Asn-209AMY1 corresponds to Ser-208AMY2, without major conse-quence for substrate binding, because all side chains form hydrogen

TABLE TWO

Comparison of direct hydrogen bond contacts in AMY1-acarbose and AMY2-acarboseThe underlined residues are not conserved between AMY1 and AMY2.

AMY1 atom Acarbose atom Distance AMY2 atom Acarbose atom Distance

Å Å

A. Active siteTyr-52-O aca-O6A 3.6 Tyr-51-O aca-O6A 3.2His-93-N�2 aca-O6A 3.0 His-92-N�2 aca-O6A 3.2Arg-178-N�2 aca-O2A 3.0 Arg-177-N�1 aca-O2A 3.0Asp-180-O�2 aca-O6A 2.8 Asp-179-O�2 aca-O6A 2.9Arg-183-N�1 aca-O2B 2.8 Lys-182-N� aca-O3C 3.0Glu-205-O�1a aca-O3B 2.7 Glu-204-O�1 aca-N4B 2.8Glu-205-O�2a aca-N4B 2.8 Glu-204-O�2 aca-O3B 2.6Trp-207-O aca-O3C 2.7 Trp-206-O aca-O3C 2.8Asn-209-N aca-O2C 2.8 Ser208-N aca-O2C 3.5Asn-209-O�1 aca-O2C 3.6 Ser208-O aca-O2C 3.2His-290-N�2 aca-O2A 2.8 His-288-N�2 aca-O2A 2.9His-290-N�2 aca-O3A 2.9 His-288-N�2 aca-O3A 2.8Asp-291-O�1 aca-O3A 2.6 Asp-289-O�1 aca-O3A 2.8Asp-291-O�2 aca-N4B 3.1 Asp-289-O�2 aca-N4B 3.2Asp-291-O�2 aca-O2A 2.5 Asp-289-O�2 aca-O2A 2.7

B. Starch granule surface binding siteGln-227-O�1 aca-O3C 3.0 Gln-226-O�1 aca-O3B 2.8Gln-227-N�2 aca-O2C 2.7 Gln-226-N�2 aca-O2B 2.9Asp-234-O�1 aca-O3B 2.7 Asp-233-O�1a aca-O2A 2.9Asp-234-O�2 aca-O2B 2.5 Asp-233-O�2a aca-O3A 3.1

Trp-276-O aca-O6A 3.0

C. Sugar tongs surface binding siteLys-375-N� aca-O2B 2.9Lys-375-N� aca-O3B 3.2Tyr-380-O aca-O2A 2.8Tyr-380-O aca-O3B 2.5Asp-381-O�1 aca-O3A 3.0Val-382-N aca-O2A 2.8Thr-392-O1 aca-O6A 3.0Asp-398-N aca-O3C 2.9His-395-N�1 aca-O6C 2.9

a The inversions in hydrogen bonding partners between AMY1 and AMY2 have no functional significance.

Barley �-Amylase Oligosaccharide Complexes

SEPTEMBER 23, 2005 • VOLUME 280 • NUMBER 38 JOURNAL OF BIOLOGICAL CHEMISTRY 32971

by guest on September 11, 2018

http://ww

w.jbc.org/

Dow

nloaded from

bonds to acarbose. Some variation, however, is seen in the conformationof the oligosaccharides as demonstrated by the torsion angles,� and�,describing the flexibility around the �-(1,4)-glucosidic linkage. �

between the cyclitol (ring A) and the 4,6-dideoxy-�-D-glucose (ring B)units in acarbose is defined by C6-C1-N4�-C4� and � is defined byC1-N4�-C4�-C5�, whereas between 4,6-dideoxy-�-D-glucose (ring B)and �-D-glucose (ring C) units, these are� �O5-C1-O4�-C4� and� �

C1-O4�-C4�-C5�. In AMY1-acarbose, these values are (�,�) �

(18°,�142°), ring A(�1) to ring B(�1), and (�,�) � (97°,�120°), ringB(�1) to ring C(�2), compared with AMY2-acarbose with (�,�) �

(�2°,�122°), ring A(�1) to ring B(�1), and (�,�) � (90°,�114°), ringB(�1) to ring C(�2). The �,� values found in the AMY1-acarbosecomplex are in good agreement with those reported for other �-amy-lase-acarbose complexes around the glucosidic bond corresponding tothe scissile bond in substrates (47), with (�,�) � (18°,�148°), in thecomplex between the psychrophilic Pseudoalteromonas haloplanktis�-amylase and the transglycosylation product derived from acarbose(8), and (�,�) � (32°,�151°) for a homologue complex with Thermo-actinomyces vulgaris R-47 �-amylase 2 (48). The difference in �,� val-ues found in the AMY2-acarbose complex may be related to the lowerresolution of the crystal structure (2.8 Å) as compared with the others.

AMY1D180A in Complex with Acarbose—The purified inactivemutant AMY1D180A showed a specific activity of 6.7 � 10�4 �mol �

min�1 � mg�1, corresponding to 6.7 � 10�4 % of wild-type AMY1having a specific activity of 100 �mol � min�1 � mg�1. In the complexbetween acarbose and AMY1D180A, with Asp-180 being the catalyticnucleophile, an intact acarbosemolecule occupies the three sugar bind-ing sites as follows: the active site, the starch granule-binding surfacesite, and the sugar tongs. The high quality of the electron density mapsallows unambiguous definition and orientation of all sugar units. In theactive site, acarbose (Fig. 2B) covers subsites �4 to �1. Eleven directhydrogen bonding interactions exist between AMY1D180A and theinhibitor, eight of these being in subsite �1 (ring D), one to ring C(subsite �2), and finally two to ring A (subsite �4). The conformationof the tetrasaccharide differs from analogue sugar moieties in the acar-bose-derived trisaccharide in the AMY1-acarbose complex. �,� valuesfor the acarbose chain are (107°,�116°), ring A(�4) to ring B(�3);(70°,�152°), ring B(�3) to ring C(�2); and (83°,�153°), ring C(�2) toring D(�1). These �,� values clearly demonstrate that the acarbosesugar chain adopts dissimilar conformations depending on the subsiteto which it is bound. An important difference that probably is a conse-quence of the mutation of the catalytic nucleophile is the loss of hydro-gen bonding between His-93 in subsite �1 and the aca-O6A atom asfound in the complex AMY1-acarbose or an analogue sugar moietyoccupying this subsite (see TABLESTWO, part A, and THREE, part A).His-93 also lacks this hydrogen bond in the maltoheptaose complexdescribed below. This suggests that the nucleophile is essential for thestabilization of the Michaelis complex to obtain the “V”-shaped orien-tation of glucosemolecules residing in subsites�1 and�1. The nucleo-phile furthermore plays a key role in the stabilization of the above-mentioned conserved His-93 for which the imidazole ring must adopt aparticular conformation in order to hydrogen bond to the substrate, andthereby stabilize the complex. Earlier mutational analysis indeed indi-cated that this histidine, which belongs to a conserved motif in GH13(49), is important for transition state stabilization (29).

AMY1D180A in Complex with Maltoheptaose—The three-dimen-sional structure of inactive AMY1D180A-maltoheptaose exhibits thesame three distinct sugar-binding sites as the complexes with acarbose(see Figs. 1, 3, and 4). In the active site, an entire molecule of maltohep-taose resides at subsites �1 to �7. Both the starch granule-binding

surface site and the sugar tongs, however, only show five of the sevenglucose units.In the active site the orientation of the glucose units (Glc2000 to

Glc2006, from the nonreducing toward the reducing end) is readilyinterpretable. The overall conformation of maltoheptaose can be con-sidered as a twisted “S” (see Fig. 3A). All glucose units at subsites �1 to�7 adopt a chair conformation. The enzyme kinetics subsite maps didnot identify subsite �7 that was proposed in a study reporting 10 sub-sites from �7 to �3 (50). For Glc2000 (in subsite �7) to Glc2003 (sub-site �4), the hydroxyl groups at C6 point into the bulk solvent, whereasthose of Glc2004 (subsite�3) to Glc2006 (subsite�1) point toward theinterior of the protein. The conformation of the heptasaccharide asdescribed by �,� values are (118°,�113°) � Glc2006(�7) toGlc2005(�6); (123°,�111°) � Glc2005(�6) to Glc2004(�5); (�62°,�73°) �Glc2004(�5) to Glc2003(�4); (122°,�122°) � Glc2003(�4) toGlc2002(�3); (69°,�148°)�Glc2002(�3) to Glc2001(�2); (83°,�155°)�Glc2001(�2) toGlc2000(�1).When compared with the analogue valuesin AMY1D180A-acarbose, � and � are rather similar between subsites�3 to �1 as expected. However, when approaching the nonreducingend of the glycone part of the active site, only �,� values betweensubsites �2 to �1 are similar to the analogue ones as found in othercomplexes (47). Hereafter they differ, which is particularly remarkablewhen comparing to (�,�) for the interglycosidic bond between subsites�3 to �2 in the B. subtilis �-amylase-maltopentaose structure (16)being (123°,�107°). This seems to be mainly due to the torsion anglesbetween subsite �5 to �4 glucoses, being subjected to a very drasticshift resulting in (�,�) � (�62°,�73°) deviating to a very high degreefrom those observed in a regular helical structure of amylose. In thisstructure the repeating unit consists of a maltotriose where the confor-mation of the glycosidic linkage is (91.8°,�153.2°), (85.7°,�145.3°), and(91.8°,�151.3°) (51). The interactions between maltoheptaose andactive site residues are listed in TABLE THREE, part A. Eight of the 16direct hydrogen bonds betweenmaltoheptaose andAMY1D180A involvethe reducing end ring (Glc2006), whereas Glc2004 has no direct inter-actions with the enzyme, and an aromatic stacking exists betweenTrp-10 and Glc2005 (subsite �2). In comparison to this relatively lownumber of contacts, AMY1-acarbose (see TABLE TWO, part A) has 15direct hydrogen bonds between the enzyme and the three observedrings of acarbose, as well as a certain number of interactions mediatedby water molecules. Fig. 4 shows a schematic representation, summa-rizing these interactions. Furthermore, a number of internal hydrogenbonds are observed for maltoheptaose in the complex (see TABLETHREE, part D). A summary of residues defining subsites �1 to �7 isgiven in TABLE FOUR. The fact that no direct interactions existbetween sugar moieties in subsite �3 (Glc2004) and the enzyme is inexcellent agreement with the binding kinetics of malto-oligosaccha-rides, which allows determination of subsite affinities as listed inTABLEFOUR. Two independent groups have arrived at somewhat differentresults (25, 50), although some features are shared. These studies showthat the affinity at subsite �3 is low compared with the other subsites(25) and even negative (50). An unexplained density around Cys-95 inthe native structure of AMY1 (28) is seen in the electron density here aswell but does not seem to affect the interaction with Glc2002.The presence of subsite �7 in the AMY1D180A-maltoheptaose com-

plex was not unequivocally defined by subsite mapping (25, 50). In thissubsite, His-45 and Val-47 make direct hydrogen bonds to the nonre-ducing end of maltoheptaose (Glc2000), see Fig. 4. Furthermore, twoindirect hydrogen bonds to Arg-56 and Lys-64 mediated by Wat-1592and Wat-1088, respectively, are observed. The ensemble of glucoseunits 2000, 2001, 2002, and 2003 adopt a half-circle conformation withthe Val-47-C1 atom as the approximate center. Clearly defined elec-

Barley �-Amylase Oligosaccharide Complexes

32972 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 280 • NUMBER 38 • SEPTEMBER 23, 2005

by guest on September 11, 2018

http://ww

w.jbc.org/

Dow

nloaded from

tron density for the glucose unit, Glc2000, strengthens the existence ofsubsite�7. Themean value of B-factors for the atoms of this ring is 42.9Å2, which is in accordance with a relatively flexible ring.

Most interestingly, Glc2001 is located between two hydrophobic res-iduesVal-47 andTyr-105, but no hydrophobic contacts exist to either of

these. The space between these two residues may be considered as thesubstrate “entrance” to the active site cleft. The plane of ring Glc2001 isnot parallel to that of the aromatic group of Tyr-105 and an imperfectaromatic stacking is present between these two entities. Probably, thisstacking would be optimized if the additional glucose unit (Glc2000) at

FIGURE 3. Views of 2 Fo � Fc electron densitymap (contoured at 1�) of the 2.0 Å resolutionstructure of the AMY1D180A-maltoheptaosecomplex. A, catalytic cleft and active site occupiedby a full substrate molecule; B, starch granule-binding surface site containing a maltopentaosemolecule; C, sugar tongs surface binding site witha maltopentaose molecule.

Barley �-Amylase Oligosaccharide Complexes

SEPTEMBER 23, 2005 • VOLUME 280 • NUMBER 38 JOURNAL OF BIOLOGICAL CHEMISTRY 32973

by guest on September 11, 2018

http://ww

w.jbc.org/

Dow

nloaded from

TABLE THREE

Direct hydrogen bond contacts in the complexes AMY1D180A-maltoheptaose and AMY1D180A-acarboseAtom labeling convention for glucosyl residues and acarbose is presented in Fig. 2A.

AMY1D180A atom Maltoheptaose atom Distance AMY1D180A atom Acarbose atom Distance

Å Å

A. Active siteHis-45-N�2 Glc2000-O3 3.0Val-47-O Glc2000-O4 3.3Cys-95-S Glc2002-O2 3.1Ala-96-O Glc2001-O2 2.6Ala-146-O Glc2002-O2 2.9 Ala-146-O Aca-O3A 3.0Ala-146-O Glc2003-O4 2.8 Ala-146-O aca-O4A 3.3Arg-178-N�1 Glc2006-O1 3.0 Arg-178-�1 aca-O1D 3.0Arg-178-N�2 Glc2006-O2 3.0 Arg-178-�2 aca-O2D 3.0Phe-181-N Glc2006-O1 3.4 Phe-181-N aca-O1D 3.3Glu-205-O�1 Glc2006-O1 2.6 Glu-205-O�1 aca-O1D 2.4His-290-N�2 Glc2006-O2 3.0 His-290-N�2 aca-O2D 2.8His-290-N�2 Glc2006-O3 3.1 His-290-N�2 aca-O3D 3.0Asp-291-O�1 Glc2006-O3 2.8 Asp-291-O�1 aca-O3D 2.7Asp-291-O�2 Glc2006-O2 2.6 Asp-291-O�2 aca-O2D 2.5Gln-296-O�1 Glc2005-O2 2.9 Gln-296-O�1 aca-O2C 2.8

B. Starch granule surface-binding site

Gln-227-N�2 Glc4003-O2 2.5 Gln-227-N�2 aca-O3C 3.0Gln-227-O�1 Glc4003-O3 2.8 Gln-227-O�1 aca-O2C 2.6Asp-234-O�1 Glc4002-O3 2.7 Asp-234-O�1 aca-O3B 2.7Asp-234-O�2 Glc4002-O2 2.7 Asp-234-O�2 aca-O2B 2.6

Lys-271-N� aca-O2D 3.2Trp-278-O Glc4002-O6 2.6

C. Sugar tongs surface-binding site

Lys-375-N� Glc3002-O2 3.0 Lys-375-N� aca-O2B 2.9Lys-375-N� Glc3002-O3 3.3 Lys-375-N� aca-O3B 3.1Tyr-380-O Glc3001-O2 2.8 Tyr-380-O aca-O2A 2.8Tyr-380-O Glc3001-O3 2.8 Tyr-380-O aca-O3B 2.6Asp-381-O�1 Glc3001-O3 2.7 Asp-381-O�1 aca-O3A 2.6Val-382-N Glc3001-O2 3.1 Val-382-N aca-O2A 2.9Thr-392-O1 Glc3001-O6 2.9His-395-N�1 Glc3003-O6 2.7 His-395-N�1 aca-O6C 2.7Gly-396-O Glc3003-O6 3.3Asp-398-N Glc3003-O3 2.8 Asp-398-N aca-O3C 2.9Asp-398-O�1 Glc3003-O3 3.2 Asp-398-O�1 aca-O3C 3.3

Trp-402-N�1 aca-O6A 3.0

D. Intramolecular hydrogen bonds

Maltoheptaose (active site) Acarbose (active site)Glc2000-O2 Glc2001-O3 2.7Glc2001-O2 Glc2002-O3 2.7Glc2001-O5 Glc2002-O6 3.3

aca-O2A aca-O3B 2.9Glc2003-O2 Glc2004-O3 2.5Glc2004-O6 Glc2005-O6 3.0

Maltoheptaose fragment (starch granule-binding surface site) Acarbose (starch granule-binding surface site)Glc4000-O2 Glc4001-O3 3.1Glc4001-O2 Glc4002-O3 3.0 aca-O2A aca-O3B 3.5Glc4002-O2 Glc4003-O3 2.9 aca-O2B aca-O3C 3.1

Maltoheptaose fragment (sugar tongs binding site) Acarbose (sugar tongs binding site)Glc3002-O2 Glc3003-O3 3.2 aca-O2A aca-O3B 3.7

aca-O2B aca-O3C 3.1Glc3003-O2 Glc3004-O6 3.4 aca-O2C aca-O6D 3.5

aca-O3D aca-O6C 2.8

Barley �-Amylase Oligosaccharide Complexes

32974 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 280 • NUMBER 38 • SEPTEMBER 23, 2005

by guest on September 11, 2018

http://ww

w.jbc.org/

Dow

nloaded from

the nonreducing end of Glc2001 inmaltoheptaose was lacking. It seemsthat Glc2000 forces Glc2001 to move away from Tyr-105 and therebyweakens the aromatic stacking interaction.Comparative studies of active site residues in AMY1D180A-maltohep-

taose and native AMY1 show that the backbones are superimposableand that no drastic reorientation of side chains occurred. All residuesinvolved in interactions with the substrate and their neighbors arehighly superimposable with the exception of Arg-183. In native AMY1,this residue displays a double conformation, for which the side chainpoints in two opposite directions. In the AMY1D180A-maltoheptaosecomplex, Arg-183 has only one conformation with the side chain point-ing toward Glc2006. Although Arg-183 does not interact directly withGlc2006, it contributes to the formation of a network ofwatermoleculesthat mediates contacts between the above-mentioned glucose unit andneighboring residues. The side chain of Lys-130 is slightly reorientedcompared with native AMY1 and makes a water-mediated hydrogenbond (Wat-896) to Glc2003. Finally, a very small shift is observed forTyr-105, which approaches Glc2001 (see above). Therefore, it appearsthat substrate binding to AMY1 does not induce a significant reorgani-zation of the protein residues.When comparing the maltoheptaose complex to AMY1D180A-acar-

bose, all atoms from rings C and D (the maltose unit in acarbose) areperfectly superimposed, in agreement with their respective �,� values.Moreover, all contacts established in the complex with maltoheptaose

are invariant in the acarbose complex (see TABLE THREE, part A) forrings C and D, and the hydrogen bond lengths are very similar. Accord-ingly, backbones and side chains of both complexes are perfectly super-imposable at subsites �1 and �2. Noticeably, the previously discussedArg-183 is orientated in an opposite direction in the complexAMY1D180A-maltoheptaose. This reorientation results in a slight mod-ification of the watermolecule network surrounding ring D in acarbose,without any major influence on the indirect water-mediated protein-inhibitor interactions. As it was shown in the AMY1D180A-maltohep-taose complex, acarbose ring B, which corresponds to glucose unitGlc2004, occupies subsite �3 and has no interaction at all with theprotein. Once again, these two sugar rings are perfectly superimposed.Finally, ring A from acarbose is located in subsite �4 and its conforma-tion slightly differs from Glc2003, its counterpart in AMY1D180A-mal-toheptaose, thus leading to an additional interaction with Ala-146-O.This difference is due to the inter-cyclic nitrogen atom in the acarvi-osine unit. Moreover, the presence of extra glucose units in maltohep-taose occupying subsites�5 to�7 seems to force the substrate to adopta half-circle conformation centered on Val-47 (see above), which is notthe case for acarbose. This is consistent with residues Val-47 and Tyr-105 defining the entrance of the catalytic pathway, thus serving as“guides” for the substrate and conferring the spatial organization for itsapproach toward the active site.

TABLE FOUR

Amino acid residues defining subsites in the active site cleft of AMY1 based on the structural studies of AMY1D180A/maltoheptaose, AMY1D180A-acarbose, and AMY1-acarbose complexesCatalytic residues are underlined. Binding kinetics in the various subsites as obtained by two independent groups (25, 50) are given in the bottom of the table.

AMY1 subsites �7 �6 �5 �4 �3 �2 �1 �1 �2 �3 �4

Residues involved in direct hydrogen bond His-45Val-47

Ala-96 Cys-95Ala-146

Ala-146 Gln-296 Tyr-52His-93Arg-178Asp-180Phe-181Glu-205His-290Asp-291

Arg-183Glu-205Asp-291

Trp-207 ND ND

Residues involved in indirect hydrogen bondsInteractions mediated by water

Ser-46Arg-56Lys-64

Asp-97 Ala-96Tyr-105

Lys-130Tyr-131Ala-145

Glu-50 Glu-50Gln-296Ser-48

Ser-48Glu-50

Glu-205Trp-207

ND ND

Residues involved in aromatic stacking Tyr-105 Tyr-52 Trp-207 ND NDSubsite affinity (kJ � mol�1)Based on MacGregor et al. (50)

�0.26 7.68 5.19 4.42 1.25 8.46 �10.20 6.50 4.94 0.32 3.24

Subsite affinity (kJ � mol�1)Based on Ajandouz et al. (25)

1.6 11.0 1.8 0.5 �0.9 3.5 ND ND 6.0 �1.2

FIGURE 4. Schematic representation of hydro-gen bonding network in the catalytic cleftof the AMY1D180A-maltoheptaose complex.Amino acid residues are in rectangles and watermolecules in ellipsoids. Catalytic residues are high-lighted. Figure was rendered using the ISIS-Drawsoftware.

Barley �-Amylase Oligosaccharide Complexes

SEPTEMBER 23, 2005 • VOLUME 280 • NUMBER 38 JOURNAL OF BIOLOGICAL CHEMISTRY 32975

by guest on September 11, 2018

http://ww

w.jbc.org/

Dow

nloaded from

StarchGranule Surface-binding Site—Asecondary binding site at thesurface of the enzyme was visualized earlier in crystal structures ofAMY2 (4) and AMY1 (28) in complex with the substrate analoguesacarbose and thio-maltotetraose, respectively. Two adjacent trypto-phans, 278 and 279 (Trp-276 and Trp-277 in AMY2), define this so-called starch granule-binding site and were identified by differentiallabeling and UV spectroscopy (52). It has been shown that these twotryptophan residues are spatially locked because of their environment.An angle of 135° found between the two planes defined by the aromatictryptophan side chains appears to be constant in both AMY1 andAMY2, in complexed as well as in native states (28). It was thereforesuggested that Trp-278 andTrp-279 (AMY1) function as a kind of “geo-metric filter” determining whether substrates should bind to the activesite.The complex between the inactive mutant AMY1D180A and malto-

heptaose displays electron density corresponding to five of the sevenglucose units (Glc4000 to Glc4004, numbered from the nonreducing tothe reducing end, Fig. 3B). For this sugar molecule which forms a half-circle, no electron density was observed for the two remaining glucoserings, probably because of their location in the bulk solvent, and thussuggesting a highly disordered state for these two rings. Gln-227 andAsp-234 are performing two major interactions as shown (TABLETHREE, part B). Furthermore, stacking interactions are presentbetween Trp-278/Glc4003 and Trp-279/Glc4002, respectively, and anetwork of water molecules mediates several indirect contacts betweenAMY1 andmaltoheptaose. Comparative studies show that the only dif-ferences between native and complexed AMY1 at this site concern asmall reorientation of the side chain of Gln-227 (to interact withGlc4003), which in turn gives rise to a small reorientation of Asn-231.In the complex between the inactive mutant and acarbose, the starch

granule-binding surface site also accommodates an intact acarbosemolecule with rings B and C stacking onto Trp-279 and Trp-278,respectively. These interactions are in excellent agreement with thosedefined in AMY1D180A-maltoheptaose (see TABLE THREE, part B),where rings B and C perfectly superimpose with maltoheptaose glucoseunits Glc4002 and Glc4003, but where the spatial location is less con-served for acarbose rings A andD, leading to a less curved conformationof the molecule as compared with maltoheptaose. Consequently, anadditional interaction is found between Lys-271 and acarbose (seeTABLE THREE, part B).

The Pair of Sugar Tongs—The third substrate-binding site is theAMY1-specific sugar tongs located in domain C (28). At this site, iden-tified in a complex between AMY1 and a thio-maltotetraose substrateanalogue, thio-DP4, Tyr-380 shows a shift upon sugar binding, resultingin a movement of the short loop defined by residues surrounding Tyr-380. This substrate binding ability is lacking in the known structures ofAMY2, and in part may stem from Pro-376AMY2, which replaces Ser-378AMY1 and thus rigidifies the loop and impedes its movement. More-over, comparative studies of the tertiary structure of domainC in several�-amylases (28) showed that this domain C-binding site is unique toAMY1-type plant �-amylases. In fact, the presence of two additional�-strands in �-amylases from other species blocks the putative accessfor the substrate at this site. Moreover, Tyr-380, one of the key residuesin the sugar tongs site, is conserved only in plant �-amylases (28).In the AMY1D180A-maltoheptaose complex, electron density is seen

for five of the seven glucose units (labeledGlc3000 toGlc3004, number-ing from the nonreducing end to the reducing end), see Fig. 3C. As inAMY1-acarbose, the sugar molecule has the shape of a half-circle withthe Tyr-380 side chain pointing into the center. This malto-oligosac-charide is exposed into the solvent, which explains the lack of electrondensity for two sugar moieties. Direct hydrogen bond interactions are

listed in TABLE THREE, part C along with intramolecular hydrogenbonds stabilizing the molecule (TABLE THREE, part D). Moreover,several indirect interactions mediated by water molecules are present(results not shown). The binding of the substrate in this site causes smallrearrangements as indicated by shifts of 0.8 and 3Å for Tyr-380-C� andTyr-380-OH, respectively. Finally, Thr-392 reorients its side chain tointeract with Glc3001.When comparing to the AMY1D180A-acarbose complex, an entire

acarbose molecule superimposes quite well with maltoheptaose on glu-cose units Glc3001, 3002, 3003, and 3004, respectively, with a largenumber of conserved interactions for the inactive mutant complexes atthis site (see TABLETHREE, part C). Ring A in acarbose is not perfectlysuperimposed with Glc3001 because of the different chemical structureof the valienamine compared with a glucose ring. Ring D is rotated 180°compared with Glc3004, resulting in its hydroxyl group at C6 pointingin the opposite direction of its homologue in Glc3004. Remarkably, thelocations of backbone and side chains, including Tyr-380, are fully con-served between the two complexes. Comparedwith the native structureof AMY1, the shift observed for Tyr-380-C� and Tyr-380-OH atoms inthe acarbose complex is 1.2 and 3.4 Å, respectively, which is in excellentagreement with those observed in AMY1D180A-maltoheptaose, as wellas with those reported for AMY1-thio-DP4 being 1.2 and 3.1 Å, respec-tively (28).

DISCUSSION

The combination of structural studies of enzyme-substrate andenzyme-inhibitor complexes as revealed by the AMY1D180A-maltohep-taose, AMY1D180A-acarbose, and AMY1-acarbose complexes leads to astructural definition of subsites �7 to �2, as summarized in TABLEFOUR. The present study reveals subsite �7, which was not unambig-uously determined neither by substrate mapping (25) nor by molecularmodeling (26, 27). Accordingly, only subsites �3 and eventually subsite�4 remain to be visualized experimentally by a complex in AMY1 orAMY2. Our results are predominantly consistent with earlier observa-tions from computer-aidedmodeling studies showing the interaction ofa maltodecaose in the active site cleft of AMY2 (26) and AMY1 (27).These modeling studies proposed a hypothetical fork shape of the non-reducing end of the binding area, which should allow the substrate toreach the catalytic site by two distinct ways, or accommodation of�-1,6-branched substrates. Neither is confirmed by the present struc-tures because no electron density corresponding to sugar rings isobserved in the putative alternative binding region. The subsite map-ping studies show low affinity at subsite �7, but biochemical data canprove the existence of a functional �7 subsite as small amounts ofp-nitrophenyl and substantial amounts of glucose are released fromPNPG7 (25) andmalto-octaose (50), respectively. AMY1 has similarKm

values for G7 and G8 (1.9-fold higher for G7) despite the 18-fold lowerkcat/Km for G7 compared with G8 (50), which may be explained bynonproductive binding of G7 to subsites �7 to �1 as illustrated here inthe AMY1D180A-maltoheptaose complex. Binding modes of G7 andacarbose covering subsites �7 to �1 and �4 to �1, respectively, inAMY1D180A are unexpected, however, from the subsite mapping. Theaffinity at subsite�2 is obviously higher than at subsites�3 and�4 (25,50), suggesting acarbose binding to occur in subsites �2 to �2, or oth-erwise that the nonreducing end valienamine ring that mimics the dis-torted glucose should be accommodated in subsite �1 as found for thewild-type AMY1-acarbose complex. The actual binding modes there-fore must be due to the loss of the nucleophile. In both complexes theglucose moiety at the reducing end is located in subsite �1, is undis-torted, and lacks the hydrogen bond between O6 and His-93, asobserved in all complexes of the �-amylase family enzymes including

Barley �-Amylase Oligosaccharide Complexes

32976 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 280 • NUMBER 38 • SEPTEMBER 23, 2005

by guest on September 11, 2018

http://ww

w.jbc.org/

Dow

nloaded from

the wild-type AMY1-acarbose complex. Asp-180 therefore seems to beessential both for distortion of the glucose ring in subsite �1 and forbinding this distorted glucose ring in the appropriate orientation, result-ing in substrate binding in a productivemode spanning subsites�1 and�1. In AMY1D180A, the loss of Asp-180 hence causes lack of distortionand thereby from binding in the productive binding mode.AMY1 is more efficient than AMY2 in degrading starch granules and

has higher affinity for substrates in general. Structural analysis of indi-vidual subsites within the substrate binding crevice of the two isozymesdoes not provide an explanation for this variation. However, the pres-ence of the sugar tongs-binding site in AMY1, so far not observed inAMY2, makes a fundamental difference. This study demonstrates theability of the sugar tongs surface site to recognize and bind naturalsubstrates, whereas the previous study only showed binding of a sub-strate analogue (28). It also confirms binding of acarbose or maltohep-taose to the starch granule surface-binding site. The orientation of mal-toheptaose molecules on the two surface substrate-binding sites inAMY1 suggests that both of them a priori are independent of eachother. First, they are separated by a considerable distance, and it isdifficult to imagine a single polysaccharide chain connecting these sites.Second, the reducing ends of oligosaccharides located in these two sitesare facing each other if we try to build a link between these sites by theshortest path. The same conclusion ismade for connectingmolecules atthe active site and the surface sites. This is only possible if the polysac-charide chain coils up around the enzyme, which seems highly improb-able. AMY1 thus could interact with three distinct sugar chains, asopposed to AMY2 binding only two sugar chains.The identification of two surface binding sites in AMY1 leads to the

question of their role in vivo. Hypothetically, these two sites may allowinteraction of the enzyme with amylose and amylopectin molecules instarch. We have shown that the substrates have a salient tendency tocircularize themselves when binding to these sites, and if trying to com-plete this curved sugar chain to obtain a full cycle, it can be shown that�-cyclodextrin (7 glucose rings) and most probably �-cyclodextrin (6glucose rings) are obtained at the sugar tongs-binding site. At the starchgranule-binding site, �-cyclodextrin seems the more probable, inaccordance with earlier studies demonstrating that AMY2 can bind�-cyclodextrin (52, 53). This, however, needs experimental confirma-tion. The structure of amylose is a helicoidally arranged chain, each turncontaining 6 glucosyl residues (54). It therefore seems plausible that thetwo surface sites in AMY1 can interact with amylose in its natural con-formation.Because of the specificity of these two sites, they may locate/orient

the enzyme in order to facilitate access to the active site for polysaccha-ride chains. In addition, the sugar tongs surface site could also disentan-gle polysaccharide chains, Tyr-380 acting as “molecular tweezers” by itsinsertion in the helical and/or lamellar structure of starchy substrates(55).These conclusions are supported by comparative studies with other

family 13 glycoside hydrolases. For example, the structures ofT. vulgarisR-47 �-amylase 1 (TVAI) complexed with malto-oligosaccharidesreveal the presence of a domain “N” putatively acting as a starch-bindingdomain (56). As compared with the �-amylase domains A–C, this extraN-terminal domainNwas shown to bindmalto-oligosaccharides at twodistinct sites, site N and site NA. The first one could interact with theouter surface of the starch helix,mainly through stacking interactions toaromatic residues, whereas the second one holds the saccharide unitsfrom both outside and inside of the helix by stacking interactions andhydrogen bonds. These authors suggest that site N is implicated inrecognizing the surfaces of rigid helical structure of starch, whereas siteNA may recognize the loose helical structure region or contribute to

unravel helical starch. Also, specific hydrolytic activity of TVAI com-pared with that of �-amylase from Aspergillus oryzae (lacking domainN) is around 18-fold higher on raw starch, supporting the crucial role ofdomain N (56). The architecture of sites N and NA exhibit no similarityto that of the sugar tongs site in AMY1, because the substrate is notcaptured by a flexible aromatic residue entering its inner curvature.These sites resemble more closely the starch granule-binding surfacesite in AMY1 and AMY2, notably by the implication of at least onetryptophan residue performing aromatic stacking interactions withsubstrate. We suggest that the sugar tongs from AMY1 and site NAfrom TVAI may share a common role, whereas the starch granule-binding surface site in AMY1/AMY2 and site N could have a similarfunction.Structural comparative studies were furthermore performed to

cyclomaltodextrin glucanotransferases (CGTase, EC 2.4.1.19), alsobelonging to the glycoside hydrolase family 13, and having, in additionto the property of hydrolyzing �-D-(1,4)-glycosidic bonds, the ability ofcircularizing oligosaccharides into �-, �-, or -cyclodextrins. The crys-tal structure of the double mutant E257Q/D229N of CGTase fromBacillus circulans (strain 251) in complex with -cyclodextrins showedthat Tyr-195 in the active site is essential for the circularization mech-anism (57). Remarkably, the phenolic ring of this residue is very close tothe center of the -cyclodextrin, and the binding of the moleculeinduces an important shift (2.6Å) of this key residue. The similaritywiththe AMY1 sugar tongs, however, ends here, as Tyr-195 from theCGTase onlymakes hydrogen bonds with -cyclodextrin but no hydro-phobic interactions. CGTases bind �- or �-cyclodextrins on two dis-tinct surface sites located in domain E (58, 59) of the carbohydrate-bindingmodule family 20 (18). Two adjacent tryptophans (Trp-616 andTrp-662) in the CGTase perform aromatic stacking onto two glucosylunits, constituting one of the binding sites of this domain (site 1). Tyr-663, whichmakes hydrophobic interactions, and Leu-600, which insertsin the cyclodextrin cylinder (59), define the second site. The structure ofbinding site 1 shares the two tryptophanyl residueswith the starch gran-ule surface-binding sites in AMY1 and AMY2, but the CGTase has avery different environment, and the two tryptophans are less stabilizedby neighboring residues. The side chains of these two residues possess ahigh degree of flexibility, allowing the accommodation of both �-, �-,and -cyclodextrins in contrast to AMY1 and AMY2 counterparts,which are in a perfectly locked position (20, 28). Evolutionary aspectsmay explain these common structural features between this CGTaseand AMY1, as CGTases derived from �-amylases, keeping their hydro-lytic activity and gaining their circularization property by the addition ofnew domains (60).C domains of AMY1 and the glucosyltransferase amylosucrase from

Neisseria polysaccharea have been compared. Amylosucrase bind sub-strates to its domain C (61), but with a distinct binding mode and loca-tion of the substrate as compared with AMY1. The �-sandwich domainC of amylosucrase does not confer enough space for accommodatingthe substrate, and the polysaccharide chain is bound onto the side of thedomain, implicating among others a hydrophobic interaction with aphenylalanine residue.Finally, bindingmodes in a complex between acarbose and amyloma-

ltase from Thermus aquaticus (47) were compared with those reportedherein. Amylomaltase is a member of GH 77, which together with GH13 is a part of the clan H. It catalyzes either the transglycosylation withtransfer from one �-1,4-glucan to another or an intramolecular cycliza-tion resulting in much larger cyclodextrins than produced by CGTases.Two acarbose molecules are bound in this structure (47), one in theactive site and a second close to the active center at a distance of 14Å. Inthis latter site, key interactions determining the conformation and bind-

Barley �-Amylase Oligosaccharide Complexes

SEPTEMBER 23, 2005 • VOLUME 280 • NUMBER 38 JOURNAL OF BIOLOGICAL CHEMISTRY 32977

by guest on September 11, 2018

http://ww

w.jbc.org/

Dow

nloaded from

ing of the inhibitor are the hydrophobic contacts of Tyr-54 and Tyr-101with unit C fromacarbose. The authors propose that Tyr-54may help incurving the glucan chain, thus favoring synthesis of cyclic products.When leaving the catalytic center, the chain could “wrap around”Tyr-54 before returning to the active site, which appeared consistentwith the formation of the smallest cycloamyloses having less than 22glucan units. Recognition, binding, and circularization schemes of acar-bose around Tyr-54 are very close to those observed at the sugar tongssite in AMY1. Because this region of amylomaltase cannot be consid-ered as a separate domain, but as an extension of the active site, thesimilarities end here. Moreover, the role of this region seems to becircularization of the glucan chain, rather than recognizing such a con-formation as we suggest for the AMY1 sugar tongs.Major conclusions drawn from this comparative study are that rec-

ognition and bindingmodes of both surface sites in AMY1 are commonto family members. These binding sites do not have their own catalyticproperties but appear to contribute to enhance the activity of theenzyme. AMY1 seems to be the simplest enzyme in terms of the three-dimensional structure that contains two distinct surface binding sitesindirectly implicated in the catalytic activity. As an example, CGTasefrom B. circulans has also two surface sites, but they are located in theextra domain E and are not present in AMY1. Thus, AMY1 appears tobe a highly “optimized” enzyme with an excellent compromise betweencatalytic efficiency on both soluble and insoluble starch-related sub-strates and structural complexity, because of the presence of the newlydiscovered sugar tongs-binding site.

Acknowledgments—We thank Annette J. Gajhede and Tine E. Gottschalk forthe preparations of AMY1�9 and Morten Tovborg Jensen and Kristian SassBak-Jensen for many stimulating discussions. We are grateful to Hassan Bel-rhali from the ID14-1 beamline andMichel Roth and Jean-Luc Ferrer from theFIP BM30A beamline at the European Synchrotron Radiation Facility syn-chrotron (Grenoble, FRANCE) for technical advice during x-ray data collec-tions.

REFERENCES1. Qian, M., Haser, R., Buisson, G., Duee, E., and Payan, F. (1994) Biochemistry 33,

6284–62942. Machius, M., Vertesy, L., Huber, R., and Wiegand, G. (1996) J. Mol. Biol. 260,

409–4213. Brzozowski, A. M., and Davies, G. J. (1997) Biochemistry 36, 10837–108454. Kadziola, A., Søgaard, M., Svensson, B., and Haser, R. (1998) J. Mol. Biol. 278,

205–2175. Brayer, G. D., Sidhu, G., Maurus, R., Rydberg, E. H., Braun, C., Wang, Y., Nguyen,

N. T., Overall, C. M., and Withers, S. G. (2000) Biochemistry 39, 4778–47916. Brzozowski, A. M., Lawson, D. M., Turkenburg, J. P., Bisgaard-Frantzen, H., Svend-

sen, A., Borchert, T. V., Dauter, Z.,Wilson, K. S., andDavies, G. J. (2000) Biochemistry39, 9099–9107

7. Qian, M., Nahoum, V., Bonicel, J., Bischoff, H., Henrissat, B., and Payan, F. (2001)Biochemistry 40, 7700–7709

8. Aghajari, N., Roth, M., and Haser, R. (2002) Biochemistry 41, 4273–42809. Kagawa,M., Fujimoto, Z.,Momma,M., Takase, K., andMizuno,H. (2003) J. Bacteriol.

185, 6981–698410. Linden, A., Mayans, O., Meyer-Klaucke, W., Antranikian, G., and Wilmanns, M.

(2003) J. Biol. Chem. 278, 9875–988411. Ramasubbu, N., Ragunath, C., and Mishra, P. J. (2003) J. Mol. Biol. 325, 1061–107612. Li, C., Begum, A., Numao, S., Park, K. H., Withers, S. G., and Brayer, G. D. (2005)

Biochemistry 44, 3347–335713. Davies, G. J., Brzozowski, A. M., Dauter, Z., Rasmussen, M. D., Borchert, T. V., and

Wilson, K. S. (2005) Acta Crystallogr. Sect. D Biol. Crystallogr. 61, 190–19314. Qian, M., Haser, R., and Payan, F. (1995) Protein Sci. 4, 747–75515. Payan, F., and Qian, M. (2003) J. Protein Chem. 22, 275–28416. Fujimoto, Z., Takase, K., Doui, N., Momma, M., Matsumoto, T., and Mizuno, H.

(1998) J. Mol. Biol. 277, 393–40717. Dauter, Z., Dauter, M., Brzozowski, A. M., Christensen, S., Borchert, T. V., Beier, L.,

Wilson, K. S., and Davies, G. J. (1999) Biochemistry 38, 8385–839218. Coutinho, P. M., and Henrissat, B. (1999) in Recent Advances in Carbohydrate Bio-

engineering (Gilbert, H. J., Henrissat, B., and Svensson, B., eds) pp. 3–12, Royal Society

of Chemistry, Cambridge, UK19. Uitdehaag, J. C., Mosi, R., Kalk, K. H., van der Veen, B. A., Dijkhuizen, L., Withers,

S. G., and Dijkstra, B. W. (1999) Nat. Struct. Biol. 6, 432–43620. Robert, X., Haser, R., Svensson, B., and Aghajari, N. (2002) Biologia (Bratislava) 57,

Suppl. 11, 59–7021. Rodenburg, K. W., Vallee, F., Juge, N., Aghajari, N., Guo, X., Haser, R., and Svensson,

B. (2000) Eur. J. Biochem. 267, 1019–102922. Jensen, M. T., Gottschalk, T. E., and Svensson, B. (2003) J. Cereal Sci. 38, 289–30023. Søgaard, M., and Svensson, B. (1990) Gene (Amst.) 94, 173–17924. Mundy, J., Svendsen, I., and Hejgaard, J. (1983) Carlsberg Res. Commun. 48, 81–9025. Ajandouz, E. H., Abe, J., Svensson, B., and Marchis-Mouren, G. (1992) Biochim.

Biophys. Acta 1159, 193–20226. Andre, G., Buleon, A., Haser, R., and Tran, V. (1999) Biopolymers 50, 751–76227. Bak-Jensen, K. S., Andre, G., Gottschalk, T. E., Paes, G., Tran, V., and Svensson, B.

(2004) J. Biol. Chem. 279, 10093–1010228. Robert, X., Haser, R., Gottschalk, T. E., Ratajczak, F., Driguez, H., Svensson, B., and

Aghajari, N. (2003) Structure (Camb.) 11, 973–98429. Søgaard, M., Kadziola, A., Haser, R., and Svensson, B. (1993) J. Biol. Chem. 268,

22480–2248430. Robert, X., Gottschalk, T. E., Haser, R., Svensson, B., and Aghajari, N. (2002) Acta

Crystallogr. Sect. D Biol. Crystallogr. 58, 683–68631. Juge, N., Andersen, J. S., Tull, D., Roepstorff, P., and Svensson, B. (1996) Protein

Expression Purif. 8, 204–21432. Mori, H., Bak-Jensen, K. S., Gottschalk, T. E., Motawia, M. S., Damager, I., Møller,

B. L., and Svensson, B. (2001) Eur. J. Biochem. 268, 6545–655833. Sambrook, J., Fritsch, E. F., and Maniatis, T. (1989)Molecular Cloning: A Laboratory

Manual, 2nd Ed., p. A1, Cold Spring Harbor Laboratory Press, Cold Spring Harbor,NY

34. Datta, A. K. (1995) Nucleic Acids Res. 23, 4530–453135. Mori, H., Bak-Jensen, K. S., and Svensson, B. (2002) Eur. J. Biochem. 269, 5377–539036. Matsui, I., and Svensson, B. (1997) J. Biol. Chem. 272, 22456–2246337. McFeeters, R. F. (1980) Anal. Biochem. 103, 302–30638. Otwinoswki, Z., and Minor, W. (1997)Methods Enzymol. 276, 307–32639. Collaborative Computational Project Number 4 (1994) Acta Crystallogr. Sect. D Biol.

Crystallogr. 50, 760–76340. Leslie, A. G.W. (1991) inCrystallographyComputing (Moras, D., Podjarny, A. D., and

Thierry, J. C., eds) pp. 50–61, Oxford University Press, Oxford41. Brunger, A. T., Adams, P. D., Clore, G.M., DeLano,W. L., Gros, P., Grosse-Kunstleve,

R. W., Jiang, J. S., Kuszewski, J., Nilges, M., Pannu, N. S., Read, R. J., Rice, L. M.,Simonson, T., andWarren, G. L. (1998) Acta Crystallogr. Sect. D Biol. Crystallogr. 54,905–921

42. Brunger, A. T. (1992) Nature 355, 472–47543. Roussel, A., and Cambillau, C. (1992) Biographics, Architecture et Fonction des

Macremolecules Biologiques, Marseille, France44. Kleywegt, G. J., and Jones, T. A. (1998) Acta Crystallogr. Sect. D Biol. Crystallogr. 54,

1119–113145. Laskowski, R. A., MacArthur, M. W., and Moss, D. S. (1993) J. Appl. Crystallogr. 26,

283–29146. Hooft, R. W., Vriend, G., Sander, C., and Abola, E. E. (1996) Nature 381, 27247. Przylas, I., Tomoo, K., Terada, Y., Takaha, T., Fujii, K., Saenger, W., and Strater, N.

(2000) J. Mol. Biol. 296, 873–88648. Ohtaki, A., Mizuno, M., Tonozuka, T., Sakano, Y., and Kamitori, S. (2004) J. Biol.

Chem. 279, 31033–3104049. MacGregor, E. A., Janecek, S., and Svensson, B. (2001) Biochim. Biophys. Acta 1546,

1–2050. MacGregor, E. A.,MacGregor, A.W.,Macri, L. J., andMorgan, J. E. (1994)Carbohydr.

Res. 257, 249–26851. Imberty, A., Chanzy, H., Perez, S., Buleon, A., and Tran, V. (1988) J. Mol. Biol. 201,

365–37852. Gibson, R. M., and Svensson, B. (1987) Carlsberg Res. Commun. 52, 373–37953. Weselake, R. J., and Hill, R. D. (1983) Cereal Chem. 60, 98–10154. Gessler, K., Uson, I., Takaha, T., Krauss, N., Smith, S.M., Okada, S., Sheldrick, G., and

Saenger, W. (1999) Proc. Natl. Acad. Sci. U. S. A. 96, 4246–425155. Imberty, A., and Perez, S. (1989) Int. J. Biol. Macromol. 11, 177–18556. Abe, A., Tonozuka, T., Sakano, Y., and Kamitori, S. (2004) J. Mol. Biol. 335, 811–82257. Uitdehaag, J. C., Kalk, K. H., van Der Veen, B. A., Dijkhuizen, L., and Dijkstra, B. W.

(1999) J. Biol. Chem. 274, 34868–3487658. Villette, J. R., Krezewinski, F. S., Looten, P. J., Sicard, P. J., and Bouquelet, S. J.-L. (1992)

Biotechnol. Appl. Biochem. 16, 57–6359. Knegtel, R. M., Strokopytov, B., Penninga, D., Faber, O. G., Rozeboom, H. J., Kalk,

K. H., Dijkhuizen, L., and Dijkstra, B. W. (1995) J. Biol. Chem. 270, 29256–2926460. del-Rio, G., Morett, E., and Soberon, X. (1997) FEBS Lett. 416, 221–22461. Skov, L. K.,Mirza,O., Sprogoe, D., Dar, I., Remaud-Simeon,M., Albenne, C.,Monsan,

P., and Gajhede, M. (2002) J. Biol. Chem. 277, 47741–4774762. Bozonnet, S., Bønsager, B., Kramhøft, B.,Mori, H., AbouHachem,M.,Willemoes,M.,

Jensen, M. T., Fukuda, K., Nielsen, P. K., Juge, N., Aghajari, N., Tranier, S., Robert, X.,Haser, R., and Svensson, B. (2005) Biologia (Bratislava) 60/16, in press

Barley �-Amylase Oligosaccharide Complexes

32978 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 280 • NUMBER 38 • SEPTEMBER 23, 2005

by guest on September 11, 2018

http://ww

w.jbc.org/

Dow

nloaded from

Xavier Robert, Richard Haser, Haruhide Mori, Birte Svensson and Nushin Aghajari-Amylase 1αOligosaccharide Binding to Barley

doi: 10.1074/jbc.M505515200 originally published online July 19, 20052005, 280:32968-32978.J. Biol. Chem. 

  10.1074/jbc.M505515200Access the most updated version of this article at doi:

 Alerts:

  When a correction for this article is posted• 

When this article is cited• 

to choose from all of JBC's e-mail alertsClick here

  http://www.jbc.org/content/280/38/32968.full.html#ref-list-1

This article cites 57 references, 10 of which can be accessed free at

by guest on September 11, 2018

http://ww

w.jbc.org/

Dow

nloaded from


Recommended