J. Appl. Glycosci.: Advance Publication doi:10.5458/jag.jag.JAG-2012_023 Review (Received January 17, 2013; Accepted March 6, 2013) J-STAGE Advance Published Date: March 21, 2013

Review

Surface Binding Sites (SBSs), Mechanism and Regulation of Degrading Amylopectin and α-limit

Marie S. Møller,1 Darrell Cockburn,1 Jonas W. Nielsen,2 Johanne M. Jensen,1,* Malene B. Vester-Christensen,1,** Morten M. Nielsen,1,*** Joakim M. Andersen,1 Casper Wilkens,1 Julie Rannes,1 Per Hägglund,1 Anette Henriksen,3,**** Maher Abou Hachem,1 Martin Willemoës2 and Birte Svensson,1,*****

1Enzyme and Protein Chemistry, Department of Systems Biology, Technical University of Denmark (Søltofts Plads, Bldn. 224, DK-2800 Kgs. Lyngby, Denmark) 2Department of Biology, University of Copenhagen, (Ole Maaløes Vej 5, DK-2200 Copenhagen N, Denmark) 3Protein Chemistry Group, Carlsberg Laboratory (Gamle Carlsberg Vej 10, DK-1799 Copenhagen V, Denmark)

Running title: GH13 Structure, Surface Sites and Regulation

*Present address: Department of Drug Design and Pharmacology, Biostructural Research, University of Copenhagen (Universitetsparken 2, DK-2100 Copenhagen Ø, Denmark) **Present address: Department of Cellular and Molecular Medicine, University of Copenhagen (Panum Instituttet, Blegdamsvej 3, DK-2200, Copenhagen N, Denmark) ***Present address: Carlsberg Laboratory (Gamle Carlsberg Vej 10, DK-1799 Copenhagen V, Denmark) ****Present address: Protein Structure and Biophysics, Novo Nordisk A/S (Novo Nordisk Park, DK-2760 Måløv, Denmark) *****Corresponding author (Tel. +4545252740, Fax. +4545886307, E-mail: [email protected]

1 J. Appl. Glycosci.: Advance Publication

Abbreviations: -CD, -cyclodextrin; AE, affinity electrophoresis; AMY1 and AMY2, barley α- isozymes 1 and 2; BASI, barley -amylase/subtilisin inhibitor; β-CD, β- cyclodextrin; CBM, carbohydrate binding module; CBM21 and CBM48, carbohydrate binding module families 21 and 48; CM-, family of chloroform-methanol soluble cereal seed proteins; DP, degree of polymerization; GH13, glycoside family 13; GH13_13 and GH13_14, family 13 subfamilies 13 and 14; HvTrxh1 and HvTrxh2, barley (Hordeum vulgare) thioredoxin isozymes 1 and 2; LD, limit dextrinase; LDI, limit dextrinase inhibitor; NTR, NADPH-dependent thioredoxin reductase; SBD, starch binding domain; SBS, surface ; SBS1 and SBS2, surface binding site 1 and 2 in AMY1 and AMY2; SPR, surface plasmon resonance; trxh, thioredoxin h.

2 Møller et al.: GH13 Structure, Surface Sites and Regulation

Abstract: Certain enzymes interact with polysaccharides at surface binding sites (SBSs) situated outside of their active sites. SBSs are not easily identified and their function has been discerned in relatively few cases. Starch degradation is a concerted action involving GH13 . New insight into barley seed α-amylase 1 (AMY1) and limit dextrinase (LD) includes i) kinetics of bi-exponential amylopectin -1 by AMY1, one reaction having low Km (8 µg/ml) and high kcat (57 s ) and the other -1 high Km (97 µg/ml) and low kcat (23 s ). β-Cyclodextrin (β-CD) inhibits the first reaction by binding to an SBS (SBS2) on domain C with Kd = 70 µM, which for the SBS2 Y380A mutant increases to 1.4 mM. SBS2 thus has a role in the fast, high- affinity component of amylopectin degradation. ii) The N-terminal domain of LD, the debranching in germinating seeds, shows distant structural similarity with domains including CBM21 present in other proteins and involved in various molecular interactions, but no binding site identity. LD is controlled by barley limit dextrinase inhibitor (LDI) which belongs to the cereal-type inhibitor family and forms a tight 1:1 complex with LD. iii) LDI in turn is regulated by disulfide reduction mediated by the barley thioredoxin h (trxh) NADPH-dependent thioredoxin reductase (NTR) system. Based on the progress monitored by released free thiol groups from LDI and its failure to inhibit LD as elicited by trxh, the LDI inactivation is proposed to stem from loss of structural integrity due to reduction of all four disulfides.

Key words: surface binding sites, α-amylase, amylopectin hydrolysis, limit dextrinase, limit dextrinase inhibitor, thioredoxin h

3 J. Appl. Glycosci.: Advance Publication

Although carbohydrate surface binding sites (SBSs) were first described more than three decades ago in phosphorylase and later in α-glucan phosphorylases from plants,1-3) functional roles of SBSs (sometimes referred to as secondary binding sites) situated outside of the area in polysaccharide metabolizing enzymes for many years received only sporadic attention. Increasing interest in SBSs, however, emerged during the past few years along with identification of several SBSs as well as further assignment of their functional properties. The slow progress with learning more about SBSs probably reflects that they have not been considered essential for the enzyme catalytic activities combined with technical difficulties at several levels, i.e., firstly with the actual identification of SBSs, and secondly in characterizing their mechanism of action and impact on the enzyme function and activity. While SBSs originally were observed in both starch-related α-glucan3-18) and glycogen1,2) converting enzymes, the later disclosed presence of SBSs in certain cell wall polysaccharide degrading enzymes of potential application in biomass utilization,19-22) indicates SBSs to be a more widely occurring generic concept in polysaccharide metabolizing enzymes.

Two SBSs (SBS1 and SBS2) in barley α-amylase were first identified by differential chemical modification23) and crystallography6,8,24) and previously demonstrated by aid of site-directed mutagenesis in barley α-amylase isozyme 1 (AMY1) to be important for activity on starch granules and polysaccharide substrates.4,9-11,25) Most recently, further distinction between the individual functional roles of SBS1 and SBS2 was achieved through detailed analysis of the effect of SBS2 in amylopectin catalyzed hydrolysis by

AMY1 showing a non-Michaelis-Menten kinetic behaviour.26)

4 Møller et al.: GH13 Structure, Surface Sites and Regulation

Regulation of hydrolases acting on starch and other poly- and oligosaccharides can also happen through interaction with proteinaceous inhibitors27-34) as well as by ligands binding at allosteric sites15,35-37) or by post-translational modification.38) Structural insights into complex formation between polysaccharide hydrolyzing enzymes and their specific proteinaceous inhibitors were first gained on α-,27-31) followed by xylanases31-33) and pectin methyl esterases34) and furthermore includes characterization of a related glycosidase/ inhibitor.39) Very recently, we solved the crystal structure of a unique complex of a debranching enzyme and the corresponding inhibitor, i.e. barley limit dextrinase (LD)40-42) and its endogenous inhibitor (LDI)43-47) that belongs to the cereal-type inhibitor protein family.29,30,48-50) Most remarkably, the LD:LDI complex (Møller et al., unpublished) is distinct from the structure already determined for barley α-amylase isozyme

2 (AMY2) in complex with barley α-amylase/subtilisin inhibitor (BASI)28 while noticeably both BASI and LDI are targets of the cytosolic h-type thioredoxin (trxh) present in mature barley seeds (HvTrxh).51-54) This latter redox-related control by trxh confers an additional level of regulation that similarly to the enzyme:inhibitor complex formation involves protein-protein interaction. By contrast to the very clear preference of HvTrxh for one of the two disulfide bonds found in BASI,51,52) apparently all four disulfides of LDI as well as the Cys59-glutathione disulfide bond are susceptible to reduction by HvTrxh1 and

HvTrxh2,54) as will be further discussed below.

Establishing a toolbox for identification and characterization of carbohydrate surface

binding sites in polysaccharide processing enzymes.

5 J. Appl. Glycosci.: Advance Publication

The SBSs are different from carbohydrate binding modules (CBMs), but in many respects SBSs seem to exert the same functions as CBMs. An obvious short-coming for understanding enzyme-polysaccharide interactions is the lack of structural insight at the atomic resolution level into the binding of macromolecular ligands to the enzyme surface.

This situation encourages the development of a dedicated toolbox comprising procedures for identifying as well as characterizing SBSs and their mode of action. In a recent review,

Cuyvers et al.55) cover the occurrence of SBSs in polysaccharide processing enzymes based on published crystal structures of carbohydrate-active enzymes having oligosaccharides accommodated at putative SBSs. These authors list seven possible functional roles of SBSs; i) targeting the enzyme to its substrate; ii) guiding substrate chains into the enzyme’s catalytic site; iii) disruption of substrate interactions to facilitate hydrolysis of its single polysaccharide chains; iv) increasing the frequency of multiple attacks; v) introduction of sites for ; vi) delivery of reaction products, e.g. to another enzyme or a membrane transport unit; and vii) anchoring of the enzyme to the cell wall of the parent microorganism. A couple of additional roles of SBSs can be included; viii) bias enzyme specificity by giving preference for a particular substrate reaction, as will be expanded below in relation to amylopectin hydrolysis catalyzed by AMY1;26) and ix) acting as binding sites for pharmaceutical molecular chaperones as demonstrated for GH27 human α- galactosidase, where an SBS interacting with β-galactose is proposed to advance folding of the recombinant therapeutic enzyme.56)

A number of techniques facilitate recognition of SBSs. Crystallography is particularly informative as it can disclose carbohydrate ligands bound on the enzyme surface at a certain distance from the active site.12-18,20-22) While most cases of putative SBSs have been

6 Møller et al.: GH13 Structure, Surface Sites and Regulation

discovered by protein crystallography,2,5-8,12-22) at least one example was reported for the structure of a xylanase in complex with xylooligosaccharides solved by NMR.57)

Amino acid residues proposed to play a role in putative SBSs can be further investigated by site-directed mutagenesis followed by various assays validating their properties, e.g. surface plasmon resonance (SPR), which is both sensitive and flexible in accurately analyzing oligosaccharide binding even when the affinity is rather modest and which may also provide useful insight into interaction with polysaccharide.4,9-11,25,58-61) Another simple albeit less exact characterization technique is affinity electrophoresis (AE), which has been used early on to determine binding constants of glycogen phosphorylases to glycogen and to validate functional roles of SBSs using engineered chimeric forms of plant phosphorylases.1,3,62,63) In AE, the migration rate of the investigated enzyme in native

PAGE gels cast to contain the interacting polysaccharide candidate is compared with the migration rate of the enzyme in a gel run in parallel but without the potential polysaccharide substrate. While about 10% of starch-metabolizing enzymes contain one or more starch binding domains (SBDs) that provide contact with starch granules and related polymeric substrates and which based on sequence similarity are organized in 10 CBM

64) families in the CAZy database (http://www.cazy.org/), there is no good estimate of the number of starch-active enzymes possessing SBSs. Barley AMY1 has no SBD, but contains two SBSs and its migration is retarded in AE performed by native PAGE with amylopectin included in the gel (Fig. 1). Insoluble polysaccharide “pull-down” is another

SBS screening technique and it is anticipated that also carbohydrate microarrays will be useful for identification of the presence of potential SBSs.

7 J. Appl. Glycosci.: Advance Publication

Obviously, a weakness in most SBS screening techniques is the interference of the carbohydrate ligand with the active site. For both SPR and AE, this interaction may be prevented by blocking the active site through reaction with a mechanism-based inhibitor that gets covalently bound to a catalytic residue and blocks substrate binding to the active site. Such an approach has enabled discrimination of binding to SBSs and at the active site of xylanase A from Bacillus subtilis.58) While this type of strategy may in principle work for any carbohydrate-active enzymes, the methods and availability of inhibitors for blocking active sites are enzyme-dependent. In fact these reagents are unavailable for the majority of carbohydrate-active enzymes and require synthesis of suitable, functionalized carbohydrate derivatives. Alternatively, if an enzyme inhibitory carbohydrate derivative is accessible which has significantly higher affinity for the active site than for the SBSs, this may offer a way to selectively block the active site. However, the discrimination between active sites and SBSs is tricky and acarbose, a powerful inhibitory pseudotetrasaccharide for several amylolytic enzymes, has also been observed to bind at SBSs.6,7,18,24,65)

SBS1 and SBS2 in barley α-amylase 1 (AMY1) and the role of SBS2 in hydrolysis of

amylopectin and amylose.

SBS1 and SBS2 in AMY1, previously named “the starch granule-binding surface site” and “the pair of sugar tongs”, respectively,6,8-11,24,25) both interact with maltoheptaose, other maltooligosaccharides and acarbose as observed in crystal structures of different complexes with wild-type and the inactive catalytic nucleophile D180A AMY1 mutant, respectively (Fig. 2).8,24) A series of single, double, and triple SBS mutants involving replacement of aromatic residues stacking onto and/or hydrogen bonding with carbohydrate

8 Møller et al.: GH13 Structure, Surface Sites and Regulation

ligands at SBS1 and SBS2 in the inactive D180A and wild-type AMY1,4,9-11,24) previously allowed us to propose functional roles of SBS1 and SBS2, notably including effects on the degree of multiple attack on amylose66) and also on the subsite map.11) SBS1 is a carbohydrate binding site where Trp278 and Trp279 (Trp277 and Trp278 in AMY26)) stack onto two adjacent glucosyl units of maltooligosaccharides and acarbose in crystal structures of AMY1 and AMY2.6,8,24) SBS2 of AMY1 revealed a 3.1 Å movement of the side chain of the central residue Tyr380 upon carbohydrate binding. Tyr380 contributes with a large number of ligand contacts. 8,24) Single W278A and W279A AMY1 mutants at SBS1 caused greater loss of starch granule affinity and hydrolysis than in case of Y380A at SBS2 and the affinity of the double W278A/W279A mutant was further reduced (Fig. 3).9,11) The SBS1 and SBS2 triple mutant W278A/W279A/Y380A showed no detectable binding to starch granules and the rate of release of soluble reducing hydrolysis products decreased to about

0.25% of that of AMY1 wild-type.11) The two SBSs thus possess differential affinity to starch granules and they synergistically advance mobilization of starch. Single α-glucan chains in amorphous regions on the starch granules may bind to SBS2. Furthermore recent data propose that SBS2 has a specific role in the kinetics of amylopectin hydrolysis.11,25,26)

GH13 is the largest glycoside hydrolase family64) containing more than 30 subfamilies67) and the SBS2 in AMY18-11,24,25,66) and AMY225) represents a remarkable case of a very important function residing in a C-terminal domain of a GH13 member.

-1 Wild-type and Y380A AMY1 hydrolyzed amylose with kcat and Km of 132 s and 0.22 mg/ml and 108 s-1 and 0.37 mg/ml, respectively, indicating that Y380A is not seriously affecting the steady-state kinetics of the hydrolysis of the linear substrate.9,24) In contrast to conventional Michaelis-Menten kinetics of product formation from amylose, however, the

9 J. Appl. Glycosci.: Advance Publication

initial rates of hydrolysis of amylopectin could not be determined, since the Km was too low to allow colorimetric quantification of reducing maltooligosaccharide products, of which most had a degree of polymerization (DP) of 7.26) Entire progress curves were therefore monitored to determine kinetic parameters for amylopectin hydrolysis. While the progress curve for degradation of amylose followed a simple mono-exponential function, the hydrolysis of amylopectin appeared to follow a bi-exponential function that described the kinetics of two reactions, “a” and “b”, for the degradation corresponding to fractions about

0.40 and 0.60 of the oligosaccharide products.26) These two reactions for catalysis of

-1 amylopectin hydrolysis by wild-type AMY1, were characterized by kcat and Km of 57 s and 8 µg/ml (for “a”) and 23 s-1 and 97 µg/ml (for “b”) (Fig. 4).26) The AMY1 Y380A mutant essentially retained these kcat values, whereas the Km of the fast reaction (“a”) increased by a factor of 20, while Km was only slightly increased for the slow reaction (“b”) compared to wild-type AMY1. A similar effect was obtained by adding 5 mM β- cyclodextrin (β-CD ) to the amylopectin hydrolysis catalyzed by wild-type AMY1. Kd of β-CD binding by wild-type AMY1 is about 10 times lower than the corresponding Kd for

Y380A AMY1 as measured by SPR.9-11,25) Noticeably, in agreement with the kinetics of the hydrolysis of amylopectin by the Y380A mutant, AMY1 was not critically affected by the presence of 5 mM β-CD.26) It is concluded that SBS2 provides a functional advantage to

AMY1 by specific binding to amylopectin clusters, which leads to saturation of the active site at a lower substrate concentration. AMY1 thus seems to belong to a small group of α- amylases containing a functional SBS2, which might explain the dramatic difference in the

68) apparent Km of this enzyme and that of α-amylase from Aspergillus oryzae.

10 Møller et al.: GH13 Structure, Surface Sites and Regulation

Regulatory processes in germinating barley seeds involving proteinaceous enzyme

inhibitors and redox reactions.

Numerous interplaying processes in the germinating barley seed result in an optimal utilization of storage compounds and ensure the successful development of the plantlet.

Regulation of germination takes place at several levels. At the gene level mediated by phytohormones, e.g. gibberellic acid that is synthesized by the embryo after water uptake at the onset of germination and diffused to the aleurone layer. In addition, several enzyme activities involved in germination are regulated by processes that involve protein-protein interactions. In the case of starch mobilization, these include the straightforward inhibition of hydrolases of starch and limit dextrins, i.e. barley α-amylase 2 (AMY2) by barley α- amylase/subtilisin inhibitor (BASI)27,28) and limit dextrinase (LD) by limit dextrinase inhibitor (LDI).43,46) Moreover, regulation of the enzyme activities by the endogenous proteinaceous inhibitors BASI and LDI appears to occur at an additional level consisting in disulfide bond reduction in the inhibitory proteins as catalyzed by the thioredoxin system composed of thioredoxin type h (two isoforms are found in barley)69) and NADPH- dependent thioredoxin-reductase (NTR; also two isoforms in barley).70) During the past decade we identified and characterized the key proteins in the barley thioredoxin system

(HvTrxh1 and HvTrxh2 and HvNTR1 and HvNTR2) as well as a large number of trxh potential target proteins, including the identification of individual specific target disulfide bonds in these proteins.51,71,72) Noticeably, the structure was solved of the disulfide-bonded trapped complex between HvTrxh2 and barley α-amylase inhibitor (BASI)52) that involves one of the two disulfide bonds present in BASI. This trapped complex of HvTrxh2 and

BASI was the first crystal structure providing detailed insight into the elements that

11 J. Appl. Glycosci.: Advance Publication

determine recognition between a thioredoxin and its target protein. Very recently, the reaction of HvTrxh1 and HvTrxh2 with LDI was investigated in solution54) as will be further described below.

Three-dimensional structure of native barley limit dextrinase (LD).

A large group of enzymes belonging to GH13 specifically catalyze hydrolysis of endo--1,6 linkages such as branch points in amylopectin and/or related poly- and oligosaccharides,64,67) e.g. α- and β-limit dextrins as well as of -1,6 linkages in the linear fungal exopolysaccharide pullulan from Aureobasidium pullulans.42,73,74) Limit dextrinase

(LD) is the sole enzyme that debranches α-limit dextrins in the germinating barley seed. It also shows high activity towards pullulan,40-42) but has much lower activity towards amylopectin itself.75) LD plays an important role in starch mobilization and is perhaps participating in starch biosynthesis by aiding with restructuring the organization and clustering of branches in intermediate structures of amylopectin sometimes referred to as phytoglycogen.76)

LD was produced recombinantly in high yields by secretory expression using Pichia pastoris as host and applying a high cell-density fermentation procedure.42) The quality of

-1 -1 the recombinant LD was excellent as its activity (kcat,app/Km,app = 488 ± 23 ml·mg ·s ) was

3.5 fold higher than of LD previously purified from germinating barley seeds.42) Kinetic

-1 parameters for hydrolysis of pullulan were Km,app = 0.16 ± 0.02 mg·ml and kcat,app = 79 ±

10 s-1 as fitted to a Michaelis-Menten equation including a substrate inhibition term.42) This distinct kinetic behavior may reflect rather than a direct inhibition by substrate that LD has

12 Møller et al.: GH13 Structure, Surface Sites and Regulation

a high transglycosylation activity, as supported by preliminary experiments, e.g. including

LD catalyzed maltosylation of -CD (Vester-Christensen et al., unpublished).

The availability of the recombinant protein made it possible to determine the three- dimensional structure of LD, initially in complex with α- and β-CDs (PDB entries 2y5e and

42) 2y4s), which are inhibitors with Kd of 27 and 0.7 µM, respectively. LD is a multidomain enzyme of subfamily GH13_1367) composed of an N-terminal domain, followed by a

CBM48 connected to the catalytic (β/α)8-barrel domain followed by a C-terminal β- sandwich domain of two sheets each with four β-strands flanked by four short α-helices.77)

In the LD α- and β-CD complexes, the structure of LD was not defined for three short loops in the N-terminal domain. Recently, however, a crystal structure was obtained of free LD

(PDB entry 4aio) in which these loops were defined (Fig. 5).78) Catalytic and key ligand binding residues at the active site superimposed very well between the free and the β-CD complex LD structures (PDB entry 2y4s).78) Compared, however, with from

Klebsiella pneumonia that belongs to the same subfamily (GH13_13) in free state and with ligand bound at the active site, two different main chain conformations were seen in the loop EGWDS containing the catalytic acid/base Glu706.73,78) Moreover, Trp708 made a

~90° rotation to stack at the subsite +2 of the active site with bound maltose, maltotriose or maltotetraose (PDB entries 2fhb, 2fhc or 2fhf, respectively) relative to the rotamer of

Trp708 in native pullulanase (PDB entry 2fgz) or in complexes with glucose (PDB entry

2fh6) or isomaltose (PDB entry 2fh8), which was seen to adopt the “inactive” rotamer conformation.73) Noticeably, the conserved barley LD sequence motif EGWDF that contains the catalytic acid/base protein donor Glu510, in structures of both free LD and

13 J. Appl. Glycosci.: Advance Publication

complexes with α- or β-CD LD adopts an “active” conformation and LD Trp512 (Fig. 5) was also found as the “active” rotamer.77,78) This apparently fixed conformation of the central element from the LD active site proposes that the active site has less flexibility in

LD as compared to the pullulanase, possibly explaining the low activity of LD towards amylopectin and the high activity towards limit dextrins as compared to bacterial .77,78)

The unresolved issue concerning the possible role of the LD N-terminal domain motivated applying several alignment methods against the entire PDB archive, which led to identification of nine unique structures.78) Five of these were domains found in various pullulanases of GH13_13 and GH13_14, but the sequence identity to LD was generally low

(7–13%) except for the domain from K. pneumonia pullulanase which showed 31% sequence identity with the N-terminal domain of LD. Three other domains have documented albeit diverse functions including binding of a peptide ligand, domain multimerization and N-acetyl-D-glucosamine binding (Fig. 6).78) The ligand binding sites in these domains, however, do not overlap. Noticeably the starch binding domain of CBM21 from Rhizopus oryzae glucoamylase showed 6% sequence identity and was identified with slightly lower score as the structurally similar domains mentioned above (Fig. 7).78,79) Only three shared residues were surface exposed among the identical residues from the structure- based sequence alignment to the CBM21 of the N-terminal domain of LD, and these residues were located in a part of CBM21 that was not structurally conserved. Furthermore, the identified starch binding residues in the CBM21 were neither conserved in the N- terminal domain of LD nor replaced by residues of similar biophysical properties.78,79)

Therefore, the N-terminal domain of LD may participate in molecular interactions, but

14 Møller et al.: GH13 Structure, Surface Sites and Regulation

there is no evidence that suggests whether this is with substrate, other proteins, or in multimerization.78)

The activity of LD which is implicated both in the biosynthesis and degradation of starch in vivo76) is regulated by the endogenous LDI43) that belongs to the cereal-type inhibitor protein family containing numerous inhibitors of α-amylases and other hydrolases, typically the proteases trypsin or chymotrypsin.27) These inhibitors are also referred to as

CM-proteins, which are abundant in barley and other cereal seeds. The LDI was recently produced recombinantly in high yields using P. pastoris as host.47) It has a molecular weight of about 13 kDa and contains four intramolecular disulfide bonds and one glutathionylated cysteinyl residue, which in the LDI purified from barley seeds was found both in glutathionylated and cysteinylated forms.44,47,54) LD:LDI forms a 1:1 molar complex45,47) (Fig. 8) of very high affinity.46) In barley, the LD:LDI complex has a counterpart in AMY2:BASI.27,28,80,81) Both of these inhibitors may undergo simultaneous regulation since they share the sensitivity to the disulfide reductase trxh.51,52) It appears, however, that the mode of action of HvTrxh on these two inhibitors is distinctly different.

Furthermore, LDI possessed a remarkably high thermal stability of importance for the beer production.47)

The interaction between barley thioredoxin h (trxh) and limit dextrinase inhibitor

(LDI).

The LDI is a target for barley thioredoxin.53) However, binding of LDI to LD prevented HvTrxh1 from inactivating LDI.54) In the case of two cereal-type inhibitors related to LDI (BMA1-1 and BMA1-2)51) as well as of BASI52) HvTrxh1 reduced a specific

15 J. Appl. Glycosci.: Advance Publication

disulfide bond. This was apparently not the case for LDI. By contrast the loss of LDI inhibitory activity accumulated with increasing release of free thiol groups by treatment with HvTrxh1 or HvTrxh2 (Fig. 9).54) Noticeably, the glutathione disulfide-linked to Cys59 was reduced before complete release of 10 thiol groups per LDI molecule. The reduction of the Cys59-glutathione disulfide seemed not accompanied by loss of LDI activity.54)

Concluding remarks

There is increasing acceptance that SBSs contribute to characteristic functional properties of polysaccharide converting enzymes, including amylolytic enzymes, in their interplay with polymeric and supramolecular carbohydrate structures such as starch granules and plant cell walls. The lack of high-resolution structures illustrating enzyme complexes with polysaccharides is compensated by mutational analysis including assay of both carbohydrate binding and hydrolysis to gain knowledge of the mechanism of action and importance of SBSs in various multivalent substrate binding events. The assignment of a specific SBS in AMY1 (SBS2) for a role in the fast, high affinity-part of amylopectin degradation is one example of SBSs constituting a field open for discoveries of fundamental properties. Such new understanding of SBSs may be tested by introduction of

SBSs using rational protein engineering strategies to confer manipulated or new functionalities and possesses strong potential for advancing industrial-scale conversions of biomass and also of starch and related sugars.

The progress in analysis of structure/function relationships of LDI and the interaction with LD fundamentally expands insight into the mode of action of proteinaceous inhibitors with starch-degrading enzymes; in particular it reveals an unexpected diversity of target

16 Møller et al.: GH13 Structure, Surface Sites and Regulation

enzyme recognition motifs in members of the cereal inhibitor family. The high thermal stability of LDI47) and its fold-similarity with related proteinaceous inhibitors suggest that the LDI fold can provide a starting point for engineering recognition and diverse specificities to be confer a robust molecule with desired requirements.

We are grateful to our colleagues Andreas Blennow, Camilla Christiansen, Rob Field, Martin Rejzek, Hans

Erik Mølager Christensen, Hiroyuki Nakai, Richard Haser and Nushin Aghajari for stimulating discussions.

This work was supported by the Danish Natural Science Research Council, the Danish Research Council for

Technology and Production Sciences, the Carlsberg Foundation, an H. C. Ørsted postdoctoral fellowship from the Technical University of Denmark (to DC), Ph. D. fellowships from the Technical University of Denmark

(to MSM, JMA, MBVC, MMN, CW), Danish National Advanced Technology Foundation (HTF), Maersk Oil and Gas A/S and the University of Copenhagen (all to MW and LWN), and an Oticon Foundation Master thesis stipend (to JMJ).

1) G. Philip, G. Gringel and D. Palm: Rabbit muscle phosphorylase derivatives with

oligosaccharides covalently bound to the glycogen storage site. Biochemistry, 21, 3043–

3050 (1982).

2) E.J. Goldsmith, R.J. Fletterick and S.G. Withers: The three-dimensional structure of

acarbose bound to glycogen phosphorylase. J. Biol. Chem., 262, 1449–1455 (1987).

3) H. Mori, K. Tanizawa and T. Fukui: Engineered plant phosphorylase showing

extraordinarily high affinity for various α-glucan molecules. Protein Sci., 2, 1621–1629

(1993).

4) M. Søgaard, A. Kadziola, R. Haser and B. Svensson: Site-directed mutagenesis of

histidine 93, aspartic acid 180, glutamic acid 205, histidine 290, and aspartic acid 291 at

17 J. Appl. Glycosci.: Advance Publication

the active site and tryptophan 279 at the raw starch binding site in barley α-amylase 1. J.

Biol. Chem., 268, 22480–22484 (1993).

5) M. Qian, R. Haser and F. Payan: Carbohydrate binding sites in a pancreatic α-amylase-

substrate complex, derived from X-ray structure analysis at 2.1 Å resolution. Protein

Sci., 4, 747–755 (1995).

6) A. Kadziola, M. Søgaard, B. Svensson and R. Haser: Molecular structure of a barley α-

amylase-inhibitor complex: implications for starch binding and catalysis. J. Mol. Biol.,

278, 205–217 (1998).

7) Z. Dauter, M. Dauter, A.M. Brzozowski, S. Christensen, T.V. Borchert, L. Beier, K.S.

Wilson and G.J. Davies: X-ray structure of Novamyl, the five-domain “maltogenic” α-

amylase from Bacillus stearothermophilus: Maltose and acarbose complexes at 1.7 Å

resolution. Biochemistry, 38, 8385–8392 (1999).

8) X. Robert, R. Haser, T.E. Gottschalk, F. Ratajczak, H. Driguez, B. Svensson and N.

Aghajari: The structure of barley -amylase isozyme 1 reveals a novel role of domain C

in substrate recognition and binding: “a pair of sugar tongs”. Structure, 11, 973984

(2003).

9) S. Bozonnet, M.T. Jensen, M.M. Nielsen, N. Aghajari, M.H. Jensen, B. Kramhøft, M.

Willemoës, S. Tranier, R. Haser and B. Svensson: The "pair of sugar tongs" site on the

non-catalytic domain C of barley α-amylase participates in substrate binding and

activity. FEBS J., 274, 50555067 (2007).

18 Møller et al.: GH13 Structure, Surface Sites and Regulation

10) M.M. Nielsen, E.-S. Seo, S. Bozonnet, N. Aghajari, X. Robert, R. Haser and B.

Svensson: Multi-site substrate binding and interplay in barley α-amylase 1. FEBS Lett.,

582, 25672571 (2008).

11) M.M. Nielsen, S. Bozonnet, E.-S. Seo, J.A. Mótyán, J.M. Andersen, A. Dilokpimol, M.

Abou Hachem, G. Gyémánt, H. Næsted, L. Kandra, B.W. Sigurskjold and B. Svensson:

Two secondary carbohydrate binding sites on the surface of barley α-amylase 1 have

distinct functions and display synergy in hydrolysis of starch granules. Biochemistry, 48,

76867697 (2009).

12) C. Ragunath, S.G. Manuel, V. Venkataraman, H.B.R. Sait, C. Kasinathan and N.

Ramasubbu: Probing the role of aromatic residues at the secondary saccharide-binding

sites of human salivary α-amylase in substrate hydrolysis and bacterial binding. J. Mol.

Biol., 384, 1232–1248 (2008).

13) S.B. Larson, J.S. Day and A. McPherson: X-ray crystallographic analyses of pig

pancreatic α-amylase with limit , oligosaccharide, and α-cyclodextrin.

Biochemistry, 49, 3101–3115 (2010).

14) A. Abe, T. Tonozuka, Y. Sakano and S. Kamitori: Complex structures of

Thermoactinomyces vulgaris R-47 α-amylase 1 with malto-oligosaccharides

demonstrate the role of domain N acting as a starch-binding domain. J. Mol. Biol., 335,

811–822 (2004).

15) L.K. Skov, O. Mirza, D. Sprogøe, I. Dar, M. Rémaud-Siméon, C. Albenne, P. Monsan

and M. Gajhede: Oligosaccharide and sucrose complexes of amylosucrase. J. Biol.

Chem., 227, 47741–47747 (2002).

19 J. Appl. Glycosci.: Advance Publication

16) B. Mikami, M. Adachi, T. Kage, E. Sarikaya, T. Nanmori, R. Shinke and S. Utsumi:

Structure of raw starch-digesting Bacillus cereus β-amylase complexed with maltose.

Biochemistry, 38, 7050–7061 (1999).

17) T. Oyama, H. Miyake, M. Kusunoki and Y. Nitta: Crystal structures of β-amylase from

Bacillus cereus var. mycoides in complexes with substrate analogs and affinity-labeling

reagents. J. Biochem. (Tokyo), 133, 467–474 (2003).

18) I. Przylas, Y. Terada, K. Fujii, T. Takaha, W. Saenger and N. Sträter: X-ray structure of

acarbose bound to amylomaltase from Thermus aquaticus - Implications for the

synthesis of large cyclic glucans. Eur. J. Biochem., 267, 6903–6913 (2000).

19) D. De Vos, T. Collins, W. Nerinckx, S.N. Savvides, M. Claeyssens, C. Gerday, G.

Feller and J. Van Beeumen: Oligosaccharide binding in family 8 glycosidases: Crystal

structures of active-site mutants of the β-1,4-xylanase pXyl from Pseudoalteromonas

haloplanktis TAH3a in complex with substrate and product. Biochemistry, 45, 4797–

4807 (2006).

20) L. Lo Leggio, S. Kalogiannis, K. Eckert, S.C.M. Teixeira, M.K. Bhat, C. Andrei, R.W.

Pickersgill and S. Larsen: Substrate specificity and subsite mobility in T. aurantiacus

xylanase 10A. FEBS Lett., 509, 303–308 (2001).

21) E. Vandermarliere, T.M. Bourgois, S. Rombouts, S. Van Campenhout, G. Volckaert,

S.V. Strelkov, J.A. Delcour, A. Rabijns and C.M. Courtin: Crystallographic analysis

shows substrate binding at the –3 to +1 active-site subsites and at the surface of

glycoside hydrolase family 11 endo-1,4-β-xylanases. Biochem. J., 410, 71–79 (2008).

20 Møller et al.: GH13 Structure, Surface Sites and Regulation

22) J. Allouch, W. Helbert, B. Henrissat and M. Czjzek: Parallel substrate binding sites in a

β-agarase suggest a novel mode of action on double-helical agarose. Structure, 12, 623–

632 (2004).

23) R.M. Gibson and B. Svensson: Identification of tryptophanyl residues involved in

binding of carbohydrate ligands to barley α-amylase 2. Carlsberg Res. Commun., 52,

373–379 (1987).

24) X. Robert, R. Haser, H. Mori, B. Svensson and N. Aghajari: Oligosaccharide binding to

barley -amylase 1. J. Biol. Chem., 280, 3296832978 (2005).

25) E.-S. Seo, J.M. Andersen, M.M. Nielsen, M.B. Vester-Christensen, C. Christiansen,

J.M. Jensen, J.A. Mótyán, M.A. Glaring, A. Blennow, L. Kandra, G. Gyémánt, S.

Janecek, R. Haser, N. Aghajari, M. Abou Hachem and B. Svensson: New insight into

structure/function relationships in plant α-amylase family GH13 members. J. Appl.

Glycosci., 57, 157–162 (2010).

26) J.W. Nielsen, B. Kramhøft, S. Bozonnet, M. Abou Hachem, S.L.S. Stipp, B. Svensson

and M. Willemoës: Degradation of the starch components amylopectin and amylose by

barley α-amylase 1: Role of surface binding site 2. Arch. Biochem. Biophys., 528, 1–6

(2012).

27) B. Svensson, K. Fukuda, P.K. Nielsen and B.C. Bønsager: Proteinaceous -amylase

inhibitors. Biochim. Biophys. Acta, Proteins Proteomics, 1696, 145–156 (2004).

28) F. Vallée, A. Kadziola, Y. Bourne, M. Juy, K.W. Rodenburg, B. Svensson and R. Haser:

Barley α-amylase bound to its endogenous protein inhibitor BASI: crystal structure at 1.9

Å resolution. Structure Fold. Des., 6, 649–659 (1998).

21 J. Appl. Glycosci.: Advance Publication

29) S. Strobl, K. Maskos, G. Wiegand, R. Huber, F.X. Gomis-Rüth and R. Glockshuber: A

novel strategy for inhibition of α-amylases: yellow meal worm α-amylase in complex

with the Ragi bifunctional inhibitor at 2.5 angstrom resolution. Structure Fold. Des., 6,

911–921 (1998).

30) F. Payan: Structural basis for the inhibition of mammalian and insect α-amylases by

plant protein inhibitors. Biochim. Biophys. Acta, Proteins Proteomics, 1696, 171–180

(2004).

31) N. Juge and B. Svensson: Proteinaceous inhibitors of carbohydrate-active enzymes in

cereals – Implication in agriculture, cereal-processing and nutrition. J. Sci. Food. Agric.,

86, 1573–1586 (2006).

32) M. Vardakou, C. Dumon, J.W. Murray, P. Christakopoulos, D.P.Weiner, N. Juge, R.J.

Lewis, H.J. Gilbert and J.E. Flint: Understanding the structural basis for substrate and

inhibitor recognition in eukaryotic GH11 xylanases. J. Mol. Biol., 375, 1293–1305

(2008).

33) S. Sansen, C.J. De Ranter, K. Gebruers, K. Brijs, C.M. Courtin, J.A. Delcour and A.

Rabijns: Structural basis for inhibition of Aspergillus niger xylanase by Triticum

aestivum xylanase inhibitor-I. J. Biol. Chem., 279, 36022–36038 (2004).

34) A. Di Matteo, A. Giovane, A. Raiola, L. Camardella, D. Bonivento, G. De Lorenzo, F.

Cervone, D. Bellincampi, and D. Tsernoglou: Structural basis for the interaction

between pectin methylesterase and a specific inhibitor protein. Plant Cell, 17, 849–858

(2005).

22 Møller et al.: GH13 Structure, Surface Sites and Regulation

35) N. Oudjeriouat, Y. Moreau, M. Santimone, B. Svensson, G. Marchis-Mouren and V.

Desseaux: On the mechanism of barley -amylase. Acarbose and cyclodextrin

inhibition of barley amylase isozymes. Eur. J. Biochem., 270, 3871–3879 (2003).

36) K. Fujii, H. Minagawa, Y. Terada, T. Takaha, T. Kuriki, J. Shimada and H. Kaneko:

Function of second glucan binding site including tyrosines 54 and 101 in Thermus

aquaticus amylomaltase. J. Biosci. Bioeng., 103, 167–173 (2007).

37) N. Pinotsis, D.D. Leonidas, E.D. Chrysina, N.G. Oikonomakos and I.M. Mavridis: The

binding of β- and γ-cyclodextrins to glycogen phosphorylase b: Kinetic and

crystallographic studies. Protein Sci., 12, 1914–1924 (2003).

38) L.N. Johnson: The regulation of protein phosphorylation. Biochem. Soc. Trans., 37,

627–641 (2009).

39) M. Hothorn, W. Van den Ende, W. Lammens, V. Rybin and K. Scheffzek: Structural

insights into the pH-controlled targeting of plant cell-wall invertase by a specific

inhibitor protein. Proc. Natl. Acad. Sci. U.S.A., 107, 17427–17432 (2010).

40) M. Kristensen, F. Lok, V. Planchot, I. Svendsen, R. Leah and B. Svensson: Isolation and

characterization of the gene encoding starch debranching enzyme limit dextrinase from

germinating barley. Biochim. Biophys. Acta, 1431, 538546 (1999).

41) R.A. Burton, X.Q. Zhang, M. Hrmova and G.B. Fincher: A single limit dextrinase gene

is expressed both in the developing endosperm and in germinated grains of barley.

Plant Physiol., 119, 859–871 (1999).

23 J. Appl. Glycosci.: Advance Publication

42) M.B. Vester-Christensen, M. Abou Hachem, H. Næsted and B. Svensson: Secretory

expression of functional barley limit dextrinase by Pichia pastoris using high cell-

density fermentation. Protein Express. Purif., 69, 112119 (2010).

43) A.W. MacGregor, L.J. Macri, S.W. Schroeder and S.L. Bazin: Purification and

characterization of limit dextrinase inhibitors from barley. J. Cereal Sci., 20, 33–41

(1994).

44) E.A. MacGregor, S.L. Bazin, E.W. Ens, J. Lahnstein, L.J. Macri, N.J. Shirley and A.W.

MacGregor: Structural models of limit dextrinase inhibitors from barley. J. Cereal Sci.,

31, 79–90 (2000).

45) A.W. MacGregor, L.J. Donald, E.A. MacGregor and H.W. Duckworth: Stoichiometry

of the complex formed by barley limit dextrinase with its endogenous inhibitor.

Determination by electrospray time-of-flight mass spectrometry. J. Cereal Sci., 37,

357–362 (2003).

46) E.A. MacGregor: The proteinaceous inhibitor of limit dextrinase in barley and malt.

Biochim. Biophys. Acta, Proteins Proteomics, 1696, 165–170 (2004).

47) J.M. Jensen, M.B. Vester-Christensen, M.S. Møller, B.C. Bønsager, H.E.M.

Christensen, M. Abou Hachem and B. Svensson: Efficient secretory expression of

functional barley limit dextrinase inhibitor by high cell-density fermentation of Pichia

pastoris. Protein Expr. Purif., 79, 217–222 (2011).

48) J. Iulek, O. Franco, M. Silva, C. Slivinski, C. Bloch, D. Rigden and M. de Sa:

Purification, biochemical characterisation and partial primary structure of a new α-

24 Møller et al.: GH13 Structure, Surface Sites and Regulation

amylase inhibitor from Secale cereale (rye). Int. J. Biochem. Cell Biol., 32, 1195–1204

(2000).

49) F. García-Maroto, P. Carbonero and F. García-Olmedo: Site-directed mutagenesis and

expression in Escherichia coli of WMAI-1, a wheat monomeric inhibitor of insect α-

amylase. Plant Mol. Biol., 17, 1005–1011 (1991).

50) M. José-Estanyol, F.X. Gomis-Rüth and P. Puigdomènech: The eight-cysteine motif, a

versatile structure in plant proteins. Plant Physiol. Biochem., 42, 355–365 (2004).

51) K. Maeda, C. Finnie and B. Svensson: Cy5 maleimide-labelling for sensitive detection

of free thiols in native protein extracts: Identification of seed proteins targeted by barley

thioredoxin h isoforms. Biochem. J., 378, 497–507 (2004).

52) K. Maeda, P. Hägglund, C. Finnie, B. Svensson and A. Henriksen: Structural basis for

target protein recognition by the protein disulfide reductase thioredoxin. Structure, 14,

17011710 (2006).

53) J.H. Wong, I.A. Iiao, K. Kobrehel and B.B. Buchanan: Thioredoxin dependent

deinhibition of pullulanase of barley malt by inactivation of a specific inhibitor protein.

Plant Physiol., 108, 67 (1995).

54) J.M. Jensen, P. Hägglund, H.E.M. Christensen and B. Svensson: Inactivation of barley

limit dextrinase inhibitor by thioredoxin-catalysed disulfide reduction. FEBS Lett., 586,

2479–2482 (2012).

55) S. Cuyvers, E. Dornez, J.A. Delcour and C.M. Courtin: Occurrence and functional

significance of secondary carbohydrate binding sites in glycoside hydrolases. Crit. Rev.

Biotechnol., 32, 93–107 (2012).

25 J. Appl. Glycosci.: Advance Publication

56) A.I. Guce, N.E. Clark, E.N. Salgado, D.R. Ivanen, A.A. Kulminskaya, H. Brumer and

S.C. Garman: Catalytic mechanism of human α-galactosidase. J. Biol. Chem., 285,

3625–3632 (2010).

57) M.L. Ludwiczek, M. Heller, T. Kantner and L.P. McIntosh: A secondary xylan-binding

site enhances the catalytic activity of a single-domain family 11 glycoside hydrolase. J.

Mol. Biol., 373, 337–354 (2007).

58) S. Cuyvers, E. Dornez, M. Abou Hachem, B. Svensson, M. Horhorn, J. Chory, J.A.

Delcour and C.M. Courtin: Isothermal titration calorimetry and surface plasmon

resonance allow quantifying substrate binding to different binding sites of Bacillus

subtilis xylanase. Anal. Biochem., 420, 90–92 (2011).

59) S. Cuyvers, E. Dornez, M.N. Rezaei, A. Pollet, J.A. Delcour and C.M. Courtin:

Secondary substrate binding strongly affects activity and binding affinity of Bacillus

subtilis and Aspergillus niger GH11 xylanases. FEBS J., 278, 1098–1111 (2011).

60) L.N. Babol, B. Svensson and R. Ipsen: Using surface plasmon resonance technology to

screen interactions between exopolysaccharides and milk proteins. Food Biophys., 6,

468–473 (2011).

61) S.K. Diemer, B. Svensson, L.N.. Babol, D. Cockburn, P. Grijpstra, L. Dijkhuizen, D.M.

Folkenberg, C. Garrigues and R.H. Ipsen: Binding interactions between α-glucans from

Lactobacillus reuteri and milk proteins characterized by surface plasmon resonance.

Food Biophys., 7, 220–226 (2012).

62) K. Takeo and S. Nakamura: Dissociation-constants of glucan phosphorylases of rabbit

tissues studied by polyacrylamide-gel disk-electrophoresis. Arch. Biochem. Biophys.,

153, 1–7 (1972).

26 Møller et al.: GH13 Structure, Surface Sites and Regulation

63) S. Shimomura and T. Fukui: A comparative-study on α-glucan phosphorylases from

plant and animal – interrelationship between the polysaccharide and pyridoxal-

phosphate binding-sites by affinity electrophoresis. Biochemistry, 19, 2287–2294

(1980).

64) B.L. Cantarel, P.M. Coutinho, C. Rancurel, T. Bernard, V. Lombard and B. Henrissat:

The Carbohydrate-Active EnZymes database (CAZy): an expert resource for

glycogenomics. Nucleic Acids Res., 37 (Database issue), D233D238 (2009).

65) J. Ševčík, E. Hostinová, A. Solovicová, J. Gašperík, Z. Dauter and K.S. Wilson:

Structure of the complex of a yeast glucoamylase with acarbose reveals the presence of

a raw starch binding site on the catalytic domain. FEBS J., 273, 2161–2171 (2006).

66) B. Kramhøft, K.S. Bak-Jensen, H. Mori, N. Juge, J. Nøhr and B. Svensson:

Involvement of individual subsites and secondary substrate binding sites in multiple

attack on amylose by barley α-amylase. Biochemistry, 44, 18241832 (2005).

67) M.R. Stam, E.G.J. Danchin, C. Rancurel, P.M. Coutinho and B. Henrissat: Dividing the

large glycoside hydrolase family 13 into subfamilies: towards improved functional

annotations of α-amylase-related proteins. Prot. Eng. Des. Sel., 19, 555–562 (2006).

68) N. Batlle, J.V. Carbonell and J.M. Sendra: Determination of depolymerization kinetics

of amylose, amylopectin, and soluble starch by Aspergillus oryzae α-amylase using a

fluorimetric 2-p-toluidinylnaphthalene-6-sulfonate/flow-injection analysis system.

Biotechnol. Bioeng., 70, 544–552 (2000).

27 J. Appl. Glycosci.: Advance Publication

69) K. Maeda, C. Finnie, O. Østergaard and B. Svensson: Identification, cloning and

characterisation of two thioredoxin h isoforms, HvTrx1 and HvTrx2, from the barley

seed proteome. Eur. J. Biochem., 270, 2633–2643 (2003).

70) A. Shahpiri, B. Svensson and C. Finnie: The NADPH-dependent thioredoxin

reductase/thioredoxin system in germinating barley seeds: gene expression, protein

profiles and interactions between isoforms of thioredoxin h and thioredoxin reductase.

Plant Physiol., 146, 789–799 (2008).

71) P. Hägglund, J. Bunkenborg, K. Maeda and B. Svensson: Identification of thioredoxin

disulfide targets using a quantitative proteomics approach based on isotope-coded

affinity tags. J. Proteome Res., 7, 5270–5276 (2008).

72) P. Hägglund, J. Bunkenborg, F. Yang, L.M. Harder, C. Finnie and B. Svensson:

Identification of thioredoxin target disulfides in proteins released from barley aleurone

layers. J. Proteomics, 73, 1133 –1136 (2010).

73) B. Mikami, H. Iwamoto, D. Malle, H.-J. Yoon, E.D. Sarikaya, Y. Mezaki and Y.

Katsuya: Crystal structure of pullulanase: evidence for parallel binding of

oligosaccharides in the active site. J. Mol. Biol., 359, 690–707 (2006).

74) J.P. Turkenburg, A.M. Brzozowski, A. Svendsen, T.V. Borchert, G.J. Davies and K.S.

Wilson: Structure of a pullulanase from Bacillus acidopullulyticus. Prot. Struct. Func.

Bioinf., 76, 516–519 (2009).

75) D.J. Manners and D. Yellowlees: Studies on carbohydrate metabolising enzymes. Part

XXVI. Limit dextrinase from germinated barley. Starch/Stärke, 23, 228–234 (1971).

76) N. Fujita, Y. Toyosawa, Y. Utsumi, T. Higuchi, I. Hanashiro, A. Ikegami, S. Akuzawa,

M. Yoshida, A. Mori, K. Inomata, R. Itoh, A. Miyao, H. Hirochika, H. Satoh and Y.

28 Møller et al.: GH13 Structure, Surface Sites and Regulation

Nakamura: Characterization of pullulanase (PUL)-deficient mutants of rice (Oryza

sativa L.) and the function of PUL on starch biosynthesis in the developing rice

endosperm. J. Exp. Bot., 60, 10091023 (2009).

77) M.B. Vester-Christensen, M. Abou Hachem, B. Svensson and A. Henriksen: Crystal

structure of an essential enzyme in seed starch degradation - barley limit dextrinase in

complex with cyclodextrins. J. Mol. Biol., 403, 739–750 (2010).

78) M.S. Møller, M. Abou Hachem, B. Svensson and A. Henriksen: Structure of the starch

debranching enzyme barley limit dextrinase reveals homology of the N-terminal

domain to CBM21. Acta Crystallogr., Sect. F 68, 1008–1012 (2012).

79) Y.-N. Liu, Y.-T.Lai, W.-I. Chou, M. D.-T. Chang and P.-C. Lyu: Solution structure of

family 21 carbohydrate-binding module from Rhizopus oryzae glucoamylase. Biochem.

J., 403, 21–30 (2007).

80) P.K. Nielsen, B.C. Bønsager, K. Fukuda and B. Svensson: Barley -amylase/subtilisin

inhibitor. Structure, biophysics and protein engineering. Biochim. Biophys. Acta,

Proteins Proteomics, 1696, 157–164 (2004).

81) B.C. Bønsager, P.K. Nielsen, M. Abou Hachem, K. Fukuda, M. Prætorius-Ibba and B.

Svensson: Mutational analysis of the -trefoil fold protein barley -amylase/subtilisin

inhibitor. J. Biol. Chem., 280, 14855–14864 (2005).

29 J. Appl. Glycosci.: Advance Publication

Figure Legends

Figure 1.

Affinity electrophoresis analysis of the interaction between the barley α-amylase isoform 1

(AMY1) and potato amylopectin. On the left is a control native PAGE gel and on the right is an identical gel except that 0.1% amylopectin is included. The outside lanes consist of a protein marker (NativeMark™ from Invitrogen) and the inner lanes are AMY1. The gels are 12% acrylamide and the buffer system is Tris-borate pH 8.7. The proteins were separated at 55 V for 17 h at 4 °C.

Figure 2.

Crystal structure of the inactive catalytic nucleophile mutant Asp180Ala of barley AMY1 in complex with maltoheptaose (PDB entry 1rp8). In red is a cartoon of the secondary structure of the enzyme, with a transparent molecular surface. The maltoheptaose is shown as grey sticks and the SBS1 and SBS2 are shown as thickened sticks and indicated by blue arrows. The active site is indicated by a red arrow.

Figure 3.

Confocal laser scanning microscopy of the interaction between barley starch granules and

Fluorescein-5-EX-labeled AMY1 wt and mutants. The dissociation constants were determined using Langmuir binding analysis towards barley starch granules.9,11) For the combined SBS1+SBS2 mutants (W278A/Y380A and W279A/Y380A) no binding to the starch granule surface was visible (data not shown), which correlates well with that no starch binding was measured using Langmuir binding analysis.9,11)

30 Møller et al.: GH13 Structure, Surface Sites and Regulation

Figure 4.

Saturation of the amylopectin degradation reaction catalyzed by AMY1. Initial rates, represented by v, calculated from progress curves are shown for the combined reaction

(closed squares), the “a” reaction (open squares), and the “b” reaction (open diamonds).

The kinetic parameters (Km; kcat) calculated using the Michaelis-Menten equation

-1 corresponding to the curves shown were: (closed squares) Km 23 µg/ml; kcat 80 s (open

-1 squares) Km 8 µg/ml; kcat 57 s , and (open diamonds) Km 97 µg/ml determined using a fixed

-1 value for kcat of 23 s calculated by subtracting kcat for the “a” reaction from that of the combined reaction. Reprinted from reference 26 with permission from Elsevier.

Figure 5.

Overall structure of LD (PDB entry 4aio). N-domain, green; CBM48, red; catalytic domain, gray; C-domain, blue; Ca2+, purple. The catalytic site residues (Asp473, Glu510, and

Asp642) are shown as orange sticks, while Trp512 is shown as gray sticks. The three short loops, which were unsolved in the structures of LD in complex with the competitive inhibitors α- or β-CD (PDB entries 2y5e and 2y4s), are indicated by arrows.

Figure 6.

Superimposition of the N-domain of LD (orange) with structurally similar domains with known function: (a) Erythropoietin receptor (PDB entry 1eba; red) and the two ligands erythropoietin-mimic peptide, EMP33, and 3,5-dibromotyrosine, DBY-T). (b) Cytokine receptor γ-chain (PDB entry 2b5i; blue) N-acetyl-D-glucosamine (NAG). (c) Esterase (PDB entry 3doi; green) and diethyl phosphonate (DEP). Reprinted from reference 78.

31 J. Appl. Glycosci.: Advance Publication

Figure 7.

Structural alignment of the N-terminal domain of LD (PDB entry 4aio,78) green) and

CBM21 from Rhizopus oryzae (PDB entry 2djm,79) blue) including a β-CD molecule from a later structure of the CBM21 from Rhizopus oryzae (PDB entry 2v8l, purple sticks) superimposed onto the binding site.

Figure 8.

The relative inhibition (%) of LD activity at different LDI molar ratios assayed by using soluble pullulan as substrate. The assay was done at 37°C and pH 5.5.47) The open circles represent the means of a triplicate data set and the error bars depict standard deviations. The solid line is a smooth fit to the data to aid clarity. Reprinted from reference 47 with permission from Elsevier.

Figure 9.

LDI reduction by HvTrxh1 (○) and HvTrxh2 (□) monitored as release of thiol groups (―), and loss of LDI activity (---). Insert shows release of thiol groups (0–60 min). Conditions:

20 µM LDI, 4 µM HvTrxh1 or HvTrxh2, 0.1 µM HvNTR2, 4 mM NADPH in 100 mM

Tris-HCl, pH 8.0 (RT). Reprinted from reference 54 with permission from Elsevier.

32 Møller et al.: GH13 Structure, Surface Sites and Regulation

Figure 1.

Figure 2.

33 J. Appl. Glycosci.: Advance Publication

Figure 3.

Figure 4.

34 Møller et al.: GH13 Structure, Surface Sites and Regulation

Figure 5.

Figure 6.

35 J. Appl. Glycosci.: Advance Publication

Figure 7.

Figure 8.

36 Møller et al.: GH13 Structure, Surface Sites and Regulation

Figure 9.

37