Survey
* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
doi:10.1016/j.jmb.2003.12.071 J. Mol. Biol. (2004) 337, 367–386 High-resolution Crystal Structure of Arthrobacter aurescens Chondroitin AC Lyase: An Enzyme –Substrate Complex Defines the Catalytic Mechanism Vladimir V. Lunin1, Yunge Li1, Robert J. Linhardt2, Hirofumi Miyazono3 Mamoru Kyogashima3, Takuji Kaneko3, Alexander W. Bell4 and Miroslaw Cygler1* 1 Biotechnology Research Institute, National Research Council of Canada, and Montréal Joint Centre for Structural Biology, Montréal Québec, 6100 Royalmount Ave. Montréal, Québec, Canada H4P 2R2 2 Department of Chemistry Division of Medicinal Chemistry and Department of Chemical and Biochemical Engineering, The University of Iowa, 115 S. Grand Ave, PHAR S328, Iowa City, IA 52242-1112, USA 3 Central Research Laboratories Seikagaku Corporation, Tateno 3-1253, Higashiyamato-shi Tokyo 207-0021, Japan 4 Montréal Proteomics Network 740 Dr Penfield Ave., Montréal Québec, Canada H3A 1A4 *Corresponding author Chondroitin lyases (EC 4.2.2.4 and EC 4.2.2.5) are glycosaminoglycandegrading enzymes that act as eliminases. Chondroitin lyase AC from Arthrobacter aurescens (ArthroAC) is known to act on chondroitin 4-sulfate and chondroitin 6-sulfate but not on dermatan sulfate. Like other chondroitin AC lyases, it is capable of cleaving hyaluronan. We have determined the three-dimensional crystal structure of ArthroAC in its native form as well as in complex with its substrates (chondroitin 4-sulfate tetrasaccharide, CStetra and hyaluronan tetrasaccharide) at resolution varying from 1.25 Å to 1.9 Å. The primary sequence of ArthroAC has not been previously determined but it was possible to determine the amino acid sequence of this enzyme from the highresolution electron density maps and to confirm it by mass spectrometry. The enzyme – substrate complexes were obtained by soaking the substrate into the crystals for varying lengths of time (30 seconds to ten hours) and flash-cooling the crystals. The electron density map for crystals soaked in the substrate for as short as 30 seconds showed the substrate clearly and indicated that the ring of central glucuronic acid assumes a distorted boat conformation. This structure strongly supports the lytic mechanism where Tyr242 acts as a general base that abstracts the proton from the C5 position of glucuronic acid while Asn183 and His233 neutralize the charge on the glucuronate acidic group. Comparison of this structure with that of chondroitinase AC from Flavobacterium heparinum (FlavoAC) provides an explanation for the exolytic and endolytic mode of action of ArthroAC and FlavoAC, respectively. Crown Copyright q 2004 Published by Elsevier Ltd. All rights reserved. Keywords: chondroitin lyase; chondroitinase AC; Arthrobacter aurescens; substrate binding; catalytic mechanism Introduction Glycosaminoglycans (GAGs) are the carbohydrate components of proteoglycans, which are a major component of the extracellular matrix.1 They are highly negatively charged polysaccharSupplementary data associated with this article can be found at doi: 10.1016/j.jmb.2003.12.071 Abbreviations used: GAG, glycosaminoglycan; MS, mass spectrometry. E-mail address of the corresponding author: [email protected] ides, composed of disaccharide repeating units of a substituted glucosamine or galactosamine attached through (1,4) linkage to a uronic acid molecule. These disaccharide units are linked (1,3) or (1,4) into a polysaccharide chain.2 The glucosamine/galactosamine units are sulfated extensively, and their synthesis requires the concerted action of a large number of enzymes.3,4 Glycosaminoglycans are degraded enzymatically by two types of enzymes, hydrolases and lyases.5 Hydrolases catalyze cleavage of the glycosyloxygen bond by addition of water, producing a saturated disaccharide. Lyases cleave the 0022-2836/$ - see front matter Crown Copyright q 2004 Published by Elsevier Ltd. All rights reserved. 368 oxygen– aglycone linkage through proton abstraction, producing an unsaturated disaccharide product with a double bond between C4 and C5. The enzymatic mechanisms of hydrolases are well understood and reactions proceed either according to the retaining or inverting mechanism.6 On the other hand, the molecular details of the enzymatic mechanism of GAG lyases are still poorly understood. A chemically plausible mechanism for the b elimination reaction has been proposed;7 however, the constitution of the active site and the roles of individual amino acids are not clear. A number of bacterial species synthesize GAG lyases, enzymes used to degrade and utilize glycosaminoglycans as a source of carbon in the bacterium’s natural environment.5,8 Polysaccharide lyases with known three-dimensional structures fall into two architectures: the right-handed parallel b-helix (pectate/ pectin lyases, chondroitinase B, rhamnoglucuronan lyase) and (a/a)n toroid (n ¼ 5 for Flavobacterium heparinum chondroitin AC and chondroitin ABC lyases, bacterial hyaluronate lyases, xanthan lyase, and n ¼ 6 for alginate lyases). A catalytic mechanism has been proposed for pectate lyases, Ca2þdependent enzymes,9 but it remains to be seen if it applies to other lyases having the b-helix topology. Several plausible mechanisms have been proposed for the lyases with the (a/a)5 toroidal fold with a histidine or a tyrosine residue in the role of a general base abstracting the proton from the C5 atom of glucuronic acid, and a tyrosine or an arginine residue acting as a general acid donating a proton to the bridging O4 atom.10,11 However, there is insufficient evidence to indicate which of these proposed mechanisms is utilized by the enzymes. The nature of the group presumed to be necessary to neutralize the charge of the glucuronic acid carboxylic group is not clear, since these enzymes do not require Ca2þ and there is no positively charged group in the vicinity of the uronic acid, as expected from the accepted chemical mechanism.7 Glycosaminoglycan-degrading enzymes with defined specificity have found widespread applications as analytical tools for the analysis of the structure of glycosaminoglycans and other polysaccharides.12 Chondroitin AC lyases are used frequently for this purpose. These enzymes cleave the glycosidic bond on the non-reducing end of an uronic acid and use as a substrate either chondroitin 4-sulfate or chondroitin 6-sulfate but not dermatan sulfate. They display a varied degree of activity toward hyaluronan.13 Enzymes from two sources, chondroitin AC lyase from Arthrobacter aurescens (ArthroAC) and from F. heparinum (FlavoAC), are commercially available (Seikagaku Corporation) and used frequently. The latter enzyme has been cloned and overexpressed in F. heparinum14 and in Escherichia coli.15 We previously determined the three-dimensional structure of this enzyme on its own16 and in complex with several dermatan sulfate oligosaccharides.10 Here, we present the three-dimensional structure of chondroitin AC lyase from A. aurescens as well Chondroitinase AC Crystal Structure and Mechanism as the complexes with chondroitin tetrasaccharide (DUAp (1 ! 3)-b-D -GalpNAc4S (1 ! 4)-b-D -GlcAp (1 ! 3)-a,b-D -GalpNAc4S, where DUAp is the unsaturated sugar residue, 4-deoxy-a-L -threo-hex4-enopyranosyluronic acid; GlcAp, glucopyranosyluronic acid; GalpN, 2-deoxy-2-aminogalactopyranose; S, sulfate; and Ac, acetate) and hyaluronan tetrasaccharide substrates. Despite the purification and characterization of ArthroAC many years ago13,17,18 and its extensive use as an analytical tool in glycosaminoglycan analysis, this enzyme has not been cloned and its amino acid sequence has not been determined, although its amino acid composition and carbohydrate content were reported.18 We obtained crystals that diffract up to 1.25 Å resolution. The high-resolution data led to high-quality electron density maps of native enzyme and several complexes, and allowed us to deduce confidently the amino acid sequence for 99% of the amino acid residues and to propose the molecular details of the catalytic mechanism, which is common to FlavoAC and hyaluronate lyases. Here, we follow the nomenclature introduced by Davies et al.19 and designate the sugars on the reducing end of the break with a þ sign and the sugars on the nonreducing end with a 2 sign. In this nomenclature, the enzymes break the bond between sugars 2 1 and þ 1, the latter being an uronic acid. Results and Discussion Amino acid sequence and its conservation The molecular mass of the entire molecule was measured by ion spray mass spectrometry. Two species were present with molecular masses of 79,502 Da and 79,840 Da. The greater mass corresponds well to the molecular mass of 79,785 Da (average mass) calculated from the amino acid sequence. Based on the analysis of MS/MS spectra, we believe that the smaller mass corresponds to the fragment missing three N-terminal amino acid residues. Such a good agreement between the predicted and measured molecular mass indicated that (1) no major assignment errors were made and (2) no glycosylation or other modification were present (FlavoAC is glycosylated).16 Peptides extracted from an in-gel trypsin digest of purified A. aurescens chondroitin AC lyase were analyzed by LC-QToF mass spectrometry as described in Materials and Methods. The resulting peaklist of fragmentation spectra was matched inhouse against the sequence deduced from the crystal structure employing Mascot (MatrixScience) software.20 Matched tandem MS spectra were confirmed manually and the remaining spectra were interpreted manually. This process was repeated several times, using the two sets of experimental data iteratively to determine the optimal sequence of the lyase (Supplementary Material). The results of MS/MS analysis of 202 tandem 369 Chondroitinase AC Crystal Structure and Mechanism mass spectra covering 88.5% of the entire ArthroAC sequence (see Materials and Methods) confirmed the amino acid sequence identified from electron density maps. Moreover, this analysis of individual MS/MS fragmentation data allowed us to make side-chain assignments for several residues located in flexible loops for which electron density was not easily interpretable (shown in small letters in Figure 1). The agreement for 18 (4 –21) residues between the residue type derived from the electron density maps and mass spectrometry, and that determined by Edman degradation provides an independent measure underscoring the low level of errors to be expected in our assignment. While the sequence of ArthroAC has not been determined previously, its amino acid composition was reported nearly 30 years ago.18 These data showed good correlation with the amino acid sequence of ArthroAC in this work, supporting our assignments (Table 1). The derived sequence of ArthroAC was used to identify homologous sequences in the NCBI database with the program BLAST.21 There are , 50 such sequences for which the similarity extends along the entire protein. They include hyaluronate, xanthan and chondroitin AC lyases. ArthroAC shows the highest level of sequence identity with various hyaluronan and xanthan lyases (38%) and a lower level of identity with chondroitin AC lyase from F. heparinum (24%). The structures of four proteins representative of these sequences are known; namely, chondroitin AC lyase 1CB8,16 Streptococcus pneumoniae hyaluronate lyase 1EGU,11 Streptococcus agalactiae hyaluronate Table 1. Comparison of percentage distribution of amino acids between chemical amino acid analysis and crystallographic assignment Residue type Ala Arg Asx Glx Trp Val Ser Thr His Phe Gly Pro Ile Leu Lys Tyr Met Cysa Sequenced from the map (%) From Hiyama & Okada18 14.1 5.2 9.1 6.35 2.5 7.0 6.7 8.85 1.85 2.9 10.6 4.0 3.6 9.1 3.3 2.5 1.45 0.9 13.4 4.7 9.0 6.5 2.4 8.2 6.6 8.4 1.7 2.9 11.9 3.3 3.3 8.8 3.3 2.6 1.1 2.3 a The significant difference in the number of cysteine residues between our data and those reported by Hiyama & Okada while there is very good agreement for the other amino acids, is likely related to the fact that hydrolysis was used for the determination of most amino acids while sulfhydryl titration was used for cysteine. lyase 1F1S22 and Bacillus sp. xanthan lyase 1J0M.23 Structure-based alignment of their sequences is shown in Figure 1. The residues important for the integrity of the active site (see below) include Asn183, His233, Tyr242, Arg296 and Glu407 of ArthroAC and are conserved in all of the related enzymes, and Glu412 is replaced by an aspartate residue in one case. The enzyme chondroitin ABC lyase I shows good sequence similarity to the above-mentioned enzymes only for the C-terminal domain. However, its three-dimensional structure (PDB code 1HN0) showed that the catalytic domain has (a/a)5 topology and can be structurally aligned with the catalytic domains of the other lyases.24 While the substrate-binding site shows little sequence conservation, the active-site residues are conserved, suggesting the same enzymatic mechanism. There is no equivalent, however, to the Asn183 of ArthroAC, and there are differences in the local structure in this region. Huang et al. suggested that this enzyme utilizes as a replacement an arginine residue remote in the linear sequence.24 The structure-based alignment of chondroitin ABC lyase I with the other lyases is included in Figure 1. Overall fold The ArthroAC molecule has an overall a þ b architecture and consists of two domains (Figure 2). The N-terminal a-helical domain contains 13 a-helices, ten of which form an incomplete double-layered (a/a)5 toroid as classified within the SCOP database.25 There is a long, deep groove on one side of the toroid that forms the location of the active site and substrate-binding site. Three a-helices at the N terminus precede the (a/a)5 toroid and constrict the cleft on one side. Residues conserved across the sequences of proteins homologous to ArthroAC cluster in the area of this cleft. The C-terminal domain is composed almost entirely of antiparallel b-strands arranged into four b-sheets. The first two sheets contain nine b-strands, some of them rather long. The third sheet has seven b-strands and the last one five b-strands. There is only one short a-helix within this domain (Figure 2). The second domain can be subdivided into two subdomains; the first encompasses the first two large b-sheets and one short a-helix, while the second subdomain is composed of the third and fourth b-sheet. Substrate-binding site Initial experiments of soaking native crystals of ArthroAC in a 5 mM solution of chondroitin 4-sulfate tetrasaccharide substrate for prolonged times before collecting diffraction data showed clear density for only a disaccharide product, indicating that, like other GAG lyases,10 ArthroAC retains enzymatic activity in the crystals. Therefore, we decided to investigate by X-ray diffraction the enzyme – substrate complex as a function of Figure 1 (legend opposite) Chondroitinase AC Crystal Structure and Mechanism 371 Figure 2. Stereo drawing of the ribbon representation of ArthroAC showing the bound tetrasaccharide. The individual a-helical hairpins of the N-terminal (a/a)5 toroid are in different colors. The individual b-sheets of the C-terminal domain are in discrete colors. Insertions in ArthroAC blocking the cleft are in gray. The substrate is shown in stick representation and colored in magenta. The Figure was prepared with the programs MOLSCRIPT52 and Raster3d.53 soaking time. The soaking time ranged from 30 seconds to ten hours (Tables 2 and 3). At the end of each soak, the crystal was immediately flashfrozen and diffraction data collected (resolution varying between 1.25 Å and 1.6 Å, Table 3). Hyaluronan tetrasaccharide was also used as a substrate with a soaking time of two minutes and diffraction data were collected to 1.9 Å resolution. The structures were refined independently. In each case, the ArthroAC molecule was refined first, then the difference electron density map was inspected and interpreted appropriately. The modeled substrate/product was included in the refinement. Several sugar units were clearly visible in each refined model. The relative occupancy of the 2 and þ sites along the timed snapshots were evaluated as described in Materials and Methods. The reaction in the present crystals occurs on the minute timescale, indicated by well-defined electron density for the entire tetrasaccharide substrate after 30 seconds and even longer soaks (Figure 3(a) and (b); and Table 4). Indeed, the substrate is best defined in this dataset, with assigned occupancies of , 0.6. Somewhat higher occupancy for the (2 1, 2 2) subsites suggests a partial presence of the disaccharide product in the 2 sites. For the ten hour soak, the sugar units in subsites (2 1, 2 2) are clearly visible in the difference electron density map and refine with an occupancy of 1.0, while the derived occupancy for the sugars at the (þ 1, þ 2) subsites were 0.25. In the complex of ArthroAC with hyaluronan tetrasaccharide, only the sugar units in subsites (2 1, 2 2) were visible in the electron density, corresponding to a disaccharide reaction product and in accord with higher activity of ArthroAC toward hyaluronan.13 The location and orientation of the hyaluronan sugars is the same as the corresponding sugars of the chondroitin sulfate tetrasaccharide substrate. In the following discussion, we refer to the ArthroAC– tetrasaccharide complex after a 30 seconds soak. The oligosaccharide is bound within the groove in the N-terminal domain (Figure 2) and makes contacts with residues Asn124, Trp125, Trp126, Arg134, Gln169, Arg174, Asn183, His233, Tyr242, Arg296, Arg300, Asn303, Asn410 and Trp465 (Figure 3(c)). Tryptophan residues play an essential role in substrate binding. Two of these residues, Figure 1. Structure-based sequence alignment for ArthroAC, FlavoAC (1CB8), S. pneumoniae hyaluronate lyase (1EGU), S. alagalactiae hyaluronate lyase (1F1S), Bacillus sp. xanthan lyase (1J0M) and P. vulgaris chondroitin ABC lyase I (1HN0). Insertions in the ArthroAC sequence that close off one end of the substrate-binding site are boxed, a-helices are marked in white letters on black background and b-stands are marked by black letters on gray background. The secondary structure assignments follow sPDBv.51 Residues conserved in all six proteins are marked by an asterisk (*) above the sequence. Arrows mark residues essential for catalysis, Asn183, His233, Tyr242 and Arg296 and Glu407. Several residues that were assigned on the basis of the MS/MS data alone (disordered side-chain) are shown in small letters. Letter ‘x’ indicates (Glu/Gln) or (Asp/Asn), which we cannot distinguish, and question marks (?) indicate residues for which we are less certain of their amino acid type. 372 Chondroitinase AC Crystal Structure and Mechanism Table 2. Data collection statistics for various substrate soaking times Substrate Soaking time CStetra Wavelength (Å) a (Å b (Å) c (Å) b (deg.) Resolution range (last shell) Rsym (last shell) Completeness (%) (last shell) I=sðIÞ (last shell) Total reflections Unique reflections Redundancy 30 seconds Two minutes Ten minutes 35 minutes Two hours Four hours Ten hours HAtetra Two minutes 0.9798 57.9 86.9 81.5 107.0 50–1.41 (1.46–1.41) 0.081 (0.794) 99.9 (99.9) 0.9798 57.6 86.5 80.7 106.8 50–1.6 (1.66–1.6) 0.077 (0.739) 100 (100) 0.9798 57.7 86.4 80.6 106.9 50–1.5 (1.55–1.5) 0.077 (0.483) 99.9 (99.9) 0.9798 57.6 86.4 80.5 106.9 50–1.35 (1.4–1.35) 0.055 (0.567) 98.3 (96.3) 0.9798 57.6 86.5 80.6 106.9 50–1.3 (1.35–1.3) 0.057 (0.517) 96.5 (89.1) 0.9798 57.6 86.5 80.6 106.9 50 –1.35 (1.4– 1.35) 0.059 (0.530) 96.1 (93.5) 0.9798 57.6 86.3 80.5 107.0 50 –1.25 (1.29 –1.25) 0.058 (0.430) 90.8 (56.4) 0.9798 57.6 86.3 80.6 106.9 50 –1.9 (1.97 –1.9) 0.075 (0.652) 100 (99.9) 8.1 (2.0) 706,542 146,507 4.8 8.3 (2.5) 501,855 100,158 5.0 9.2 (3.4) 460,480 120,769 3.8 8.4 (2.0) 452,905 162,044 2.8 11.1 (3.5) 1,338,199 178,855 7.5 8.9 (2.9) 610,947 158,853 3.8 9.8 (2.8) 939,404 188,554 5.0 7.7 (2.8) 235,444 60,174 3.9 Table 3. Refinement statistics Model Resolution range R-factor (Rfree) No. non-hydrogen protein atoms No. of water molecules Average B-factor (Å2) Protein main-chain atoms Side-chain atoms Water molecules Substrate atoms r.m.s.d. bond length (Å) r.m.s.d bond angle (deg.) Ramachandran plot. Residues in: Most favorable region (%) Disallowed regions (%) tetra Hg derivative Native CS 30 seconds CStetra 10 minutes CStetra 10 hours HAtetra Two minutes 50–1.3 0.134 (0.155) 5687 1103 50 –1.35 0.130 (0.175) 5623 1025 50–1.45 0.138 (0.177) 5617 1049 50– 1.5 0.136 (0.170) 5627 1061 50–1.25 0.113 (0.142) 5646 1107 50–1.9 0.190 (0.251) 5629 837 14.9 16.5 28.7 14.1 16.3 29.4 0.019 1.84 0.022 1.97 15.5 17.9 31.4 24.8 0.023 1.84 15.4 16.7 30.5 21.2 0.021 1.86 13.6 15.4 29.6 15.5 0.021 2.01 21.8 22.4 31.6 22.5 0.018 1.72 90.8 0.3 90.7 0.3 90.5 0.3 89.7 0.3 89.9 0.3 88.3 0.5 Trp126 and Trp465, provide stacking interactions with the sugar units occupying positions 2 1 and þ 2, respectively, while Trp125 is aligned edge-on and forms a hydrogen bond with the bridging oxygen atom between the þ 2 and þ 1 units. The 4-O-sulfo groups of the substrate form several interactions with the protein. The 4-O-sulfo group of the þ 2 sugar makes H-bonds to Gln169 and, Table 4. Relative occupancies of sugars in positions 2 2, 2 1, þ1, þ2 for different soaking times Site occupancy Soak time Native CStetra HAtetra 30 seconds Two minutes Ten minutes 35 minutes Two hours Four hours Ten hours Two minutes 22 21 þ1 þ2 Phosphate 0.7 0.7 0.7 0.7 1.0 1.0 1.0 1.0 0.7 0.7 0.7 0.7 1.0 1.0 1.0 1.0 0.6 0.5 0.4 0.4 0.4 0.3 0.25 – 0.6 0.5 0.4 0.4 0.4 0.3 0.25 – 1.0 0.4 0.5 0.6 0.6 0.6 0.7 0.7 1.0 through a bridging water molecule, to Asp222 and Gln232. The 2 1 sugar 4-O-sulfo group is positioned just above the guanidinium group of Arg300 and, in addition, forms H-bonds to Glu412 and Asn598 through a bridging water molecule. Both 4-O-sulfo groups contribute to substrate binding but do not add significantly to the specificity of substrate recognition. The chondroitin sulfate tetrasaccharide used in these studies was obtained by the action of GAG lyases and contained an unsaturated ring at the non-reducing end, with a C4vC5 double bond.10 The electron density for the 2 2 sugar corresponds very well to this unsaturated ring in E1 conformation, with the C5 having sp2 hybridization (Figure 3a). A list of hydrogen bonds between the tetrasaccharide and the protein side-chains is given in Table 5. Most interesting from the viewpoint of the mechanism of catalysis is the conformation and interactions with the enzyme of the þ 1 glucuronic acid. The electron density shows that this ring assumes a distorted boat conformation O,3B with 373 Chondroitinase AC Crystal Structure and Mechanism Table 5. Close contacts between the substrate and the protein in CStetra complex for the 30 second soak experiment Sugar number Atom Protein atom Distance (Å) 22 O2 O2 O3 O6A O6B OD2 Asn303 NH1 Arg300 OD1 Asn303 NH1 Arg134 NH2 Arg134 2.85 3.07 3.03 3.01 2.90 21 O7 O7 SO43 SO44 NH2 Arg300 NH1 Arg300 NE Arg300 NH2 Arg300 2.92 3.01 3.27 3.46 þ1 O2 O2 O3 O4 O4 C5 O6A O6A O6B ND1 Asn410 OD1 Asn124 ND2 Asn124 NH2 Arg296 OH Tyr242 OH Tyr242 NE2 His233 ND2 Asn183 OD1 Asn183 3.20 2.77 2.91 3.02 2.88 2.75 2.76 3.11 2.62 þ2 O1 O3 O7 SO41 SO43 NH2 Arg174 NE1 Trp125 NH1 Arg174 NE2 His233 NE2 Gln169 3.22 3.32 3.07 2.85 2.34 the hydroxyl groups equatorial and the C5 carboxylate group pseudoaxial (Figure 3(b)). This carboxylate group is placed exactly opposite the conserved side-chain of Asn183, whose OD1 and ND2 atoms are clearly distinguished by their peak height in the electron density map (Supplementary Material). The distance between the amide ND2 atom and the carboxyl O6A is 3.1 Å, and that between carbonyl OD1 and carboxyl O6B is 2.6 Å. This short O6B· · ·OD1 distance suggests the existence of a hydrogen bond between them,26 which in turn would indicate that the glucuronic acid carboxylate group is protonated and therefore in a neutral state. The O6A, in addition to accepting a hydrogen bond from ND2 of Asn183, also forms a second, 2.8 Å long hydrogen bond with the NE2 atom of His233. The geometry of hydrogen bonds involving the carboxylate group is very close to ideal, the C6 –O6A/B-donor angles are in the range 115 – 1248 and the –COO group and the three hydrogen bond donor atoms are nearly coplanar (Figure 3(d)). His233 donates, in addition, a second hydrogen bond from the ND1 atom to the side-chain of the conserved Glu407; therefore, this histidine residue must be protonated. The hydroxyl groups at C2 and C3 of the þ 1 unit are held firmly through several hydrogen bonds. Atoms O2 and O3 are H-bonded to OD1 and ND2 of Asn124, respectively, while O2 is H-bonded also to ND2 of Asn410 (Figure 3(c)). Two other residues make crucial contacts with the substrate. The hydroxyl O atom of Tyr242 is within H-bonding distance of the bridging oxygen atom between þ 1 and 2 1 sugars (2.9 Å), and is 3.3 Å from O5 of the þ 1 sugar ring. This hydroxyl group is also only 2.8 Å from the C5 of the glucuronic acid, with the O atom nearly along the derived direction of the C5 – H5 bond (Figure 3(d)). The Arg296 side-chain also forms a hydrogen bond to the bridging oxygen atom between þ 1 and 2 1 sugars and is 3.0 Å from the Tyr242 hydroxyl group. On the other hand, the distance from the potential base His233 NE2 to the C5 of the þ 1 glucuronic acid is relatively long at 4.0 Å. The electron density corresponding to the carboxylate group of the glucuronic acid in the þ 1 site showed not two but three bulges, with the third one having somewhat lower density. This shape was common to the density observed in all datasets and its position coincided with the phosphate group in the native structure. We have modeled a phosphate group with partial occupancy in the same location (Table 4), assuming that it is present there when the glucuronic acid does not occupy this site. The total occupancy of the þ 1 site, that is glucuronic acid plus phosphate, equals 1. A strong peak in the electron density map was found in the proximity of the tetrasaccharide. It is surrounded by six oxygen atoms in tetragonal bipyramidal coordination, with distances to the equatorial oxygen atoms of 2.2– 2.4 Å and to the axial oxygen atoms of 2.8 Å. This peak was interpreted as a sodium ion. The equatorial ligands are the carbonyl groups of His233 and Trp465 and two water molecules, while the axial ligands are the OG1 of Thr235 and a water molecule. Substrate specificity The initial characterization of ArthroAC showed that the enzyme degrades chondroitin 4-sulfate, chondroitin 6-sulfate and hyaluronan.13 Our results with soaking various oligosaccharides indicated that ArthroAC displays higher activity toward hyaluronan than to chondroitin sulfate. We have determined the kinetic parameters of ArthroAC with GAG obtained from whale cartilage (CS-A, predominantly chondroitin 4-sulfate), shark cartilage (CS-C, predominantly chondroitin 6-sulfate), shark fin (CS-D, a mixture of chondroitin 4-sulfate, 6-sulfate and -4,6-disulfate), low molecular mass hyaluronan and high molecular mass hyaluronan (Table 6). Indeed, the Vmax for hyaluronan is twice as high as that for chondroitin sulfate, while the KM calculated on a per monomer basis is about three times lower. These values point to a rather small contribution made by the sulfate groups to the total binding energy of the substrate. Comparison with other GAG lyases ArthroAC is very similar in overall structure to other GAG lyases with the (a/a)5 fold. The closest similarity is with S. pneumonia hyaluronate lyase (SpHL, PDB code 1EGU). These two structures superimpose with root-mean-squares (rms) deviation of 1.3 Å for 632 Ca atoms out of 750 residues 374 Chondroitinase AC Crystal Structure and Mechanism Figure 3 (legend opposite) (Figure 4). The superposition with FlavoAC results in an rms deviation of 1.4 Å for 468 Ca atoms. The similarity extends as well to the general mode of substrate binding. Comparison with the SpHL Y408F mutant (inactive, equivalent to Y242 of ArthroAC) complexed with hyaluronan oligosaccharide,27 FlavoAC complexed with the dermatan sulfate hexasaccharide, and its inactive Y234F Chondroitinase AC Crystal Structure and Mechanism 375 Figure 3. (a) Electron density for the tetrasaccharide of chondroitin 4-sulfate substrate in the omit map calculated without the substrate present with the data for the 30 seconds soak of ArthroAC in 5 mM tetrasaccharide solution and contoured at the 3s level. The phosphate group with partial occupancy is shown. An arrow marks the bond that is cleaved by the enzyme. (b) Close-up of the same map for the þ 1 glucuronic acid showing the distorted boat conformation; (c) stereo view of the substrate-binding site. The tetrasaccharide is shown in thick semitransparent lines, hydrogen bonds are shown with broken lines; (d) close-up of the active site showing residues involved in catalysis and the hydrogen bonding network connecting these residues. Magenta dotted line marks the close contact between the Tyr hydroxyl group and the C5 atom of glucuronic acid. mutant (ArthroAC Y242 equivalent) complexed with chondroitin tetrasaccharide10 shows that the mode of oligosaccharide binding is nearly identical, with the largest difference restricted to the glucuronic acid at the þ 1 site (Figure 4(b)). The side-chains making crucial contacts with the oligosaccharide are conserved in their type and position. In the previously observed complexes with Tyr-to-Phe mutant enzymes, the þ 1 glucuronic acid ring is in a chair conformation, with the carboxylic group at C5 in an equatorial position. That differs from the conformation of the glucuronate sugar observed in the wild-type ArthroAC – substrate complex, where the ring forms a distorted boat with a pseudoaxial carboxylate group. A detailed comparison of these structures reveals subtle differences in the hydrogen bonding network involving active-site residues. In ArthroAC, the hydrogen bonds between the C5 carboxylic group of the þ 1 sugar and Asn183 and His233 have nearly ideal geometry. The corresponding H-bonds in the FlavoAC(Y234F) and SpHL(Y408F) 376 Chondroitinase AC Crystal Structure and Mechanism Table 6. Kinetic parameters of ArthroAC CS-A CS-C CS-D V (DABS/s) Km (mg/ml) Ma (Da) Km (mM) Mb (Da) Km (nM) 1.42 £ 10 2 3 0.196 505.2313 0.387 19,000 10.304 1.74 £ 10 2 3 6.10 £ 10 2 4 0.196 0.188 511.1607 532.166 0.383 0.353 43,000 30,000 4.554 6.256 Disaccharide composition (%)c GAG source DDi-0S DDi-6S DDi-4S DDi-diSD DDi-diSE DDi-triS Whale cartilage 1.6 19.3 76.2 2.7 0.3 – Shark cartilage 1.7 72.9 15.4 9.3 0.6 – a b c LMHA (50K) HA (1000K) 2.57 £ 10 2 3 0.052 401.3 0.130 50,000 1.046 2.78 £ 10 2 3 0.082 401.3 0.205 1,000,000 0.082 Shark fin 0.6 43.9 26.9 21.3 7 0.3 Disaccharides composition. GPC-HPLC (CS STD). Analysis of unsaturated disaccharides from glycosaminoglycan by HPLC. complexes have less favorable geometry (Figure 4(b)). This asparagine residue in SpHL is modeled with the opposite orientation of the amide group to that in the other two enzymes. In the FlavoAC (Y234F)-chondroitin sulfate tetrasaccharide complex, the glucuronate sugar remains in a chair conformation with all substituents to the ring being equatorial, but rotates so that the carboxylic group occupies the space vacated by the missing hydroxyl group of Tyr234.10 Thus, even small changes in the active site affect the mode of binding of the substrate and may result in a nonproductive binding. Catalytic mechanism The enzymatic reaction carried out by GAG lyases is thought to proceed via abstraction of the C5 proton by a general base followed by proton donation by a general acid or a water molecule to the bridging O4, with concomitant b-elimination of the leaving group.7 Recent kinetic analysis of the FlavoAC using a well defined synthetic substrate agrees with the predicted stepwise, as opposed to concerted, mechanism.28 A proposal for the mechanism of polysaccharide lyases formulated by Gacesa included the neutralization of the acidic group by a positively charged group to shift the equilibrium toward the enolate tautomeric form.7 Unlike polysaccharide lyases that adopt the b-helix fold, the structures of FlavoAC and hyaluronate lyases showed an absence of such a positively charged group in the vicinity of the acidic group of uronate. Instead, an asparagine side-chain was found to face and form a H-bond with the acidic group, leaving the issue of neutralization of the acidic group an open question. The structures of the complexes presented here suggest that Asn183 OD1 forms a strong hydrogen bond with this carboxylic group, substantially increasing its pKa and promoting its protonation.29,30 This asparagine residue is aided by His233, which also forms a H-bond to the carboxylate group of uronic acid and is protonated in the complex, as judged from it being hydrogen bonded to two acidic groups (Figure 3(d)). On the basis of structural evidence from GAG lyase –oligosaccharide complexes, several proposals have been put forward as to the identity of the general base and general acid participating in the reaction. Jedrzejas and co-workers proposed that the proton is abstracted by a nearby histidine residue (His233 in the present structure), and that another proton is donated to O4 by a tyrosine residue (Tyr242 here).11,27,31 They support this proposed role of the histidine residue as the general base by the close distance between the histidine NE1 and C5 in their structures. Huang et al. considered three possible mechanistic scenarios, ultimately favoring one in which a tyrosine residue initially functioned as a general base and subsequently as a general acid.10 In a structurally related but sequence-distant alginate lyase also, the central catalytic role was assigned to a tyrosine residue.32,33 The kinetic characterization of FlavoAC with a synthetic substrate strongly favors tyrosine as the proton acceptor.28 We have carefully re-analyzed the available structural data to assess the role of histidine (His233 in ArthroAC and its equivalent) in catalysis. The first suggestion for the role of histidine as a general base was derived from the structure of hyaluronan lyase and its complexes with a disaccharide product,11,34 in which the þ 1 site was not occupied. These authors modeled the þ 1 sugar and estimated the NE1· · ·C5 distance to be , 4 Å. This value corresponds to the N· · ·C van der Waals distance and seems to be too long for the proposed proton-abstracting role of the histidine. A direct observation of the enzyme – substrate complex was accomplished for FlavoAC(Y234F)10 and for SpHL(Y408F).27 In the case of FlavoAC (Y234F), it was concluded that the mode of substrate binding in the þ 1 site is influenced by the Chondroitinase AC Crystal Structure and Mechanism 377 Figure 4. (a) Stereo view of the superposition of the Ca traces of ArthroAC (blue) and SpHL (1EGU) (red); (b) overlay of oligosaccharide substrates from ArthroAC (blue), FlavoAC(Y234F) (1HMW, magenta) and SpHL(Y408F) (1LXK, green) based on the superposition of the backbone of active site Asn, His, Tyr (Phe) and Arg residues. The Figure was prepared with programs sPDBv51 and POV-Raye (http://www.povray.org/). mutation and does not reflect the reaction intermediate (the C5 carboxylic group occupies the volume vacated by the missing Tyr hydroxyl group in the Phe mutant) therefore precluding the distinction between the possible mechanisms. In the case of SpHL(Y408F), the distance between His399 and the C5 of the uronic acid is 3.73 Å, but the C5 proton lies almost along the line of CD2 – NE2 bond, very poor geometry indeed for proton abstraction. If the Phe408 is replaced in this model by the original Tyr, its hydroxyl O atom would be only , 3.0 Å from C5 and would have a much better geometry for interacting with the C5 proton. Hence, we find these results equally inconclusive concerning the assignment of His as the general base. A structure of the crystals of S. agalactiae hyaluronate lyase soaked for several days in 10 – 50 mM hexasaccharide substrate was reported recently.31 In the model, the enzyme assumes a more open conformation and Tyr408 and His399 are further away from the carbohydrate. Specifically, the distance of 5.7 Å between the NE1 atom of His399 and the C5 atom of the þ 1 sugar does not allow the conclusion to be drawn that the His plays the role of general base. As well, the temperature factors in this structure (PDB code 1LXM) for most of the atoms of the modeled substrate are equal to 100 Å2, significantly higher than the average value of , 35 Å2 for the surrounding atoms, suggesting poor order of the substrate. The difference electron density map (not reported in the 378 Chondroitinase AC Crystal Structure and Mechanism Figure 5. Proposed catalytic mechanism of (a/a)5 GAG lyases. Panel 1, Tetrasaccharide bound in the substrate-binding site. The hydrogen bonding network involving the active site and substrate-binding residues is shown schematically. The tryptophan residues stack against sugars in 21 and þ2 subsites. His233 is protonated and OD1 of Asn183 forms a strong hydrogen bond with the protonated carboxylic group of glucuronic acid. Deprotonated Tyr242 abstracts the C5 proton. Panel 2, Tyr242 accepts the proton from C5 atom leading to carbanion formation. Now Tyr242 forms hydrogen bond with bridging O4. Panel 3, The proton from Tyr242 is transferred to the O1 of galactosamine in the 2 1 subsite with concomitant break of C4 –O4 bond and formation of unsaturated ring in þ 1 subsite. Panel 4, Movement of Trp465 triggers the release of disaccharide product from (þ1, þ 2) subsites, reorganization of the active site and release of product from (22, 2 1) subsites. The enzyme is ready for the next catalytic cycle. original paper) calculated from the deposited structure factors and coordinates, with the substrate excluded from calculations, showed weak and scattered density that in our view cannot be modeled reliably as a single carbohydrate molecule (Supplementary Material). The role of His399 as a general base is further put in doubt by the fact that His399Ala mutant of SpHL retains a significant fraction (6%) of the wild-type enzyme activity.11 The present series of timed snapshots of structures of complexes of the wild-type ArthroAC with chondroitin 4-sulfate tetrasaccharide substrate provides very strong support for Tyr242, rather than His233, playing the role of a general base in the first step of catalysis. The proposed reaction mechanism is shown in Figure 5. Substrate binding to ArthroAC is associated with a deformation of the glucuronate sugar ring at the þ 1 site from a chair to a distorted boat and protonation of its acidic group. Such a change to a higher-energy conformation of the sugar ring is not unusual and has been observed in other carbohydrate-processing enzymes.35 – 40 The distortion of the ring brings the pseudoaxial acidic group coplanar with the amide group of Asn183, and the presence of an additional proton leads to the formation of two hydrogen bonds, one of them being a strong O· · ·H· · ·O 379 Chondroitinase AC Crystal Structure and Mechanism hydrogen bond. The high-resolution structure of the complex shows that the Oh atom of Tyr242 is positioned 2.8 Å from the C5 atom of the þ 1 glucuronic acid, along the direction of the C5 – H bond. This Oh atom is at the same time within hydrogen bonding distance of the O atom bridging þ 1 and 2 1 sugars and of Arg296. We postulate that Tyr242 becomes deprotonated upon substrate binding and that the Oh takes up the proton from the C5 atom of glucuronic acid (Figure 5, panel 2), which is then transferred to the bridging O atom 2.9 Å away, with a concomitant break of the O4 – C4 bond (Figure 5, panel 3). A similar mechanism utilizing a deprotonated tyrosine residue was proposed for alginate lyase A1-III,33 which is structurally similar to ArthroAC despite little sequence similarity. What then is the role of His233? This side-chain, and its equivalents in other lyases, is in the proximity of the þ 1 glucuronic acid, but with the NE2 atom at a distance of , 4 Å or more from C5 of the þ 1 sugar, typical for van der Waals interactions. The geometry of the H· · ·NE2 is also such that the direction of the expected lone pair on NE2 is , 608 away from the direction to the C5 proton. This geometry is observed consistently in the structures of the FlavoAC and SpH carbohydrate complexes. Finally, NE2 of His233 is hydrogen bonded to the carboxylate group of the þ 1 glucuronic acid and, thus, must be protonated. Considering these facts collectively, it is unlikely that His233 plays the role of general base to remove the C5 proton. We postulate that this side-chain has two functions: (1) it helps properly orient the carboxylate group of the glucuronic acid through the NE2· · ·O6A hydrogen bond; (2) being protonated in the complex and in conjunction with Arg296, it lowers the pKa of Tyr242 leading to the deprotonation of its hydroxyl group and priming it for the role as a general base (Figure 5, panel 1). This second function of His233 is directly related to the nearby presence of Glu407, which engages the histidine ND2 proton in a hydrogen bond and through electrostatic interactions aids in histidine protonation. Glu407 forms at the same time a salt-bridge with Arg296, completing a tetrad of hydrogen bonded residues (Tyr242, His233, Arg296, Glu407) forming the active site. The conservation in FlavoAC and hyaluronate lyases of all Asn183, His233, Tyr242, Arg300 and Glu407, residues critical for the above described mechanism supports the view that this is a common catalytic mechanism for this class of enzymes. A possible mechanism of product release was illuminated by the structure of thimerosal-soaked ArthroAC solved here. Of the three bound heavyatoms, two bound to surface-exposed cysteine residues. The third bound to Cys408, which in the native structure is covered by the 460 –469 loop and inaccessible to the solvent, suggesting that this loop is flexible enough to allow access for a relatively large thimerosal molecule. At the tip of this loop is Trp465, which provides stacking inter- actions to the sugar unit of the substrate bound in the þ 2 site (Figure 3(c)). This loop in ArthroACHg assumes an open conformation, indicating its intrinsic mobility even in the crystal environment. When Trp465 is sequestered from the substratebinding cleft, the side-chains of Arg296 and His233 extend into the volume previously occupied by Trp465 (Figure 6). His233 moves away from Glu407, forming hydrogen bonds with water molecules, and is most likely neutral. The saltbridge between Glu412 and Arg296 is broken, but the side-chain of Glu407 follows Arg296 and forms a stronger salt-bridge, with two H-bonds. Movement of this loop after the reaction is completed would substantially decrease the binding of the product in (þ 1, þ 2) sites and aid in its release from the enzyme. A rearrangement of residues in the active site together with their hydrogen bonding network and neutralization of His233 would lead to reprotonation of Tyr242 making it ready for the next catalytic cycle (Figure 5, panel 4). We propose that the deprotonation of Tyr242 and protonation of His233 is concomitant with substrate binding and triggered by the interaction of the carboxylate group on the þ 1 sugar with Asn183 and His233. The access of the substrate to the active site necessitates either local movements of two to three loops closing off the binding site,10 or a hinge movement between the domains leading to a global movement of the N and C-terminal domains creating an opening to the active site.27,33 Structural rationale for the exolytic versus endolytic mode of action Enzymatic characterization of ArthroAC showed that it acts as an exolyase, releasing disaccharides from the glycosaminoglycan substrate,41 while FlavoAC acts as an endolyase.42 Comparison of the structures of these two enzymes provides a rationale for the observed differences in the mode of action. ArthroAC has two insertions in the N-terminal domain relative to the FlavoAC: ,15 residues (Arg23– Ser38) and 25 residues (Thr343 – Gly366). These two segments form a-helicescontaining loops near the N and C termini of the domain, which come together, closing off a large part of the cleft running along the side of the (a/a)5 toroidal N-terminal domain. These loops form a wall that converts the open cleft into a deep cavity and precludes binding of an extended oligosaccharide. The size of the cavity restricts the binding of the carbohydrate on the non-reducing end and can accommodate approximately, two to three sugar molecules (Figure 7(a)), which correlates with the exolytic activity of this enzyme. Binding of a longer carbohydrate would require substantial rearrangement of these two loops and apparently does not occur frequently. The open cleft of the FlavoAC does not impose such a constraint and allows binding to the middle of a long carbohydrate (Figure 7(b)). 380 Chondroitinase AC Crystal Structure and Mechanism Figure 6. Stereo view of the conformation of the 460– 469 loop in native and complexed ArthroAC (blue) and in ArthroAC-Hg (magenta). The location of the thimerosal Hg atom near the Cys408 in the open conformation is shown as a magenta ball. The side-chains of His233, Arg296, Glu407 and Trp465 are shown explicitly with the hydrogen bonds marked in broken lines. The þ2 and þ 1 sugars are also shown. Materials and Methods Protein purification The protein was purified from its natural host at Seikagaku Corporation. To stimulate expression of chondroitin AC lyase II, A. aurescens was grown under the following culture conditions: medium (32.5 l) containing 0.4% (w/v) peptone (Kyokuto Pharmaceutical Industry Co. Ltd, Tokyo), 0.4% (w/v) Ehrlich’s fish extract (Kyokuto Pharmaceutical Industry Co. Ltd, Tokyo) and 0.75% (w/v) chondroitin sulfate C (Seikagaku Co., Tokyo); initial pH 6.2; aeration rate 1 vessel volume per minute; agitation speed 220 rpm; cultivation time 24 hours. The enzyme was purified essentially as described13 but with some modifications. Briefly, after bacterial cells were pelleted by centrifugation at 15,000g for 15 minutes, solid ammonium sulfate was added to the supernatant fluid up to 75% saturation. At 4 8C, 75 g of the protein precipitate (50,000 units) was dialyzed against 20 mM sodium acetate (pH 5.2) and loaded onto an SP-Sepharose (Amersham Bioscience Corp, Piscataway, NJ) column (2.6 cm £ 70 cm) pre-equilibrated with the same buffer. The enzyme was eluted with a linear gradient from 20 mM to 300 mM sodium acetate buffer (pH 5.2). Enzyme activity and protein amounts were monitored as described.13 This chromatography procedure was repeated three times and the enzymecontaining fractions collected and concentrated by ultrafiltration (Ultrafilter Type P0200, cut-off 20,000 Da, Advantec, Tokyo Japan) under N2 : Finally, the enzyme was loaded onto a Sephacryl S-200 HR gel-filtration column (Amersham Bioscience Corp, Piscataway, NJ) equilibrated with 0.01 M sodium acetate buffer (pH 5.6) and eluted with the same buffer. The fractions showing the highest specific activity and high purity by SDSPAGE were pooled and used in crystallization experiments. Oligosaccharide preparation Chondrotin 4-sulfate tetrasaccharide (CStetra) and hyaluronan tetrasaccharide were prepared and characterized as described.10 Briefly, chondroitin 4-sulfate from bovine trachea and dermatan sulfate from porcine intestinal mucosa were subjected to controlled depolymerization using chondroitin ABC lyase and the reactions were terminated prior to completion by boiling for five minutes. Each oligosaccharide mixture was separated on a BioGel P6 column and fractions consisting of tetrasaccharides and hexasaccharides were collected. These mixtures were further fractionated by strong anion-exchange HPLC, single oligosaccharides were obtained, and their purity confirmed by capillary electrophoresis and their structures by MS and NMR analyses. The structure of CStetra was DUAp (1 ! 3)-b-D -GalpNAc4S (1 ! 4)-b-D GlcAp (1 ! 3)-a,b-D -GalpNAc4S, where DUAp is the unsaturated sugar residue, 4-deoxy-a-L -threo-hex-4-is Chondroitinase AC Crystal Structure and Mechanism 381 Figure 7. Stereo view of the molecular surface within the extended substrate binding site of (a) ArthroAC– tetrasaccharide complex. The insertions in the sequence capping the substrate binding site are shown in green. (b) FlavoAC, with bound tetrasaccharides. The Figure was prepared with the program GRASP.54 enopyranosyluronic acid; GlcAp is glucopyranosyluronic acid; GalpN is 2-deoxy-2-aminogalactopyranose; S is sulfate; and Ac is acetate. Protein crystallization and data collection Needle-shaped crystals of ArthroAC were reported many years ago;13 however, they were not characterized. We obtained crystals by the hanging-drop, vapor-diffusion method in drops containing 2 ml of protein (10 mg/ ml) and 2 ml of reservoir solution (23% (w/v) PEG8000, 0.1 M sodium phosphate buffer (pH 6.4), 0.4 M ammonium acetate, 10% (v/v) glycerol) suspended over 1 ml of reservoir solution. Small crystals appeared within a few days. Larger crystals were obtained by macroseeding with precipitant concentration lowered to 21% (w/v). 382 Chondroitinase AC Crystal Structure and Mechanism The crystals belong to the monoclinic space group P21 with cell dimensions a ¼ 57:6 Å, b ¼ 85:5 Å, c ¼ 80:5 Å, b ¼ 106.98 and contain one molecule in the asymmetric unit. Prior to data collection the crystals were immersed for ten seconds in a cryoprotectant solution containing 22.5% (w/v) PEG8000, 0.1 M sodium phosphate buffer (pH 6.4), 0.4 M ammonium acetate and 20% (v/v) glycerol, mounted in a nylon loop and flash-cooled in a cold stream of N2 gas to 100 K. These crystals diffracted to 1.3 Å resolution at the X8C beamline at Brookhaven National Laboratory. Crystals of the protein complexed with thimerosal were obtained by an overnight soak of native crystals in the cryoprotectant solution containing 2 mM Hg salt. These crystals were non-isomorphous with the native crystals and had cell dimensions: a ¼ 57:4 Å, b ¼ 85:3 Å, c ¼ 82:2 Å, b ¼ 105.88. Multiwavelength anomalous diffraction (MAD) data at three wavelengths were collected (Table 7). Enzyme –substrate complexes To obtain complexes of ArthroAC with its substrates we resorted to a series of soaks of the native crystals in cryoprotectant solution containing the substrate for times ranging from 30 seconds to ten hours (Table 2). Native crystals were soaked in the cryo-protecting solution containing 5 mM CStetra or HAtetra for a specified length of time, then flash-frozen in a cold N2 gas stream (100 K) on the detector and used immediately for data collection. There was no significant change in cell dimensions as compared to the crystals of native ArthroAC (Table 2). Diffraction data were collected at the X8C beamline, NSLS, Brookhaven National Laboratory, using the Quantum-4 CCD detector. The highest-resolution data, 1.25 Å, were obtained from the native crystal soaked for ten hours in 5 mM CStetra. Data processing and scaling was performed with HKL2000.43 Data collection statistics are shown in Tables 2 and 7. Structure determination and refinement The structure of ArthroAC was solved from the Hg- Table 7. MAD data collection statistics Wavelength (Å) a (Å) b (Å) c (Å) b (deg.) Resolution range (Å) (last shell) Rsym (last shell) Completeness (%) (last shell) I=sðIÞ (last shell) Total reflections Unique reflections Redundancy Hg— peak Hg— inflection Hg— remote 1.005133 1.009078 57.3 85.2 82.1 105.8 40–1.3 (1.35– 1.30) 0.067 (0.325) 98.8 (89.0) 14.7 (6.8) 0.997068 40– 1.3 (1.35– 1.30) 0.066 (0.274) 98.4 (85.0) 14.9 (5.3) 876,240 182,880 4.8 Native 0.950000 57.6 86.5 80.5 106.9 50–1.3 (1.35– 1.30) 0.086 (0.540) 96.4 (93.4) 8.6 (3.4) 884,015 183,424 50–1.4 (1.45– 1.40) 0.067 (0.221) 99.0 (90.7) 13.7 (4.5) 721,975 147,941 1,294,307 178,066 4.8 4.9 7.3 containing thimerosal derivative of ArthroAC. Analysis of the MAD datasets using the program SOLVE44 revealed three Hg sites in the asymmetric unit. These sites were used to calculate experimental phases to a resolution of 1.3 Å and resulted in an overall figure of merit (FOM) of 0.33 – 1.3 Å. Electron density modification performed with the program RESOLVE45 assuming a solvent content of 0.4 led to a significant increase of the FOM (0.45 at 1.3 Å resolution) and substantially improved the electron density map. Approximately, 80% of the protein main chain was built automatically using the program RESOLVE. Additional fragments of the main chain and many side-chains were built manually using the program O.46 Since the primary sequence of the protein was unknown, the amino acid type for each residue was selected to fit the experimental electron density map. At 1.3 Å resolution, most of the assignments were unambiguous and nearly the entire chain was traced in the initial map. This initial sequence assignment was adjusted during the progress of refinement as the electron density features improved. Initial refinement was performed with the program CNS, version 1.147 and the model was rebuilt manually using the program O. Subsequently, refinement was continued using the program REFMAC5, version 5.1.08.48 During refinement of this model, 1% of the reflections were set aside for the calculation of Rfree : Water molecules were initially added automatically with the program CNS and subsequently updated and corrected by visual inspection of the difference map. The final model of the thimerosal-soaked crystal has been refined to an R-factor of 0.134 and Rfree of 0.156 at 1.3 Å resolution (Table 3). The current model contains 754 residues (Pro4 – Arg757). The final amino acid sequence of the model was derived from the combination of electron density maps and mass spectrometry data. Refinement of native protein and enzyme – substrate complexes The model of Hg-bound ArthroAC was taken as a starting point for refinement of the crystal structure of the native protein. Refinement was performed with the program REFMAC5. For monitoring of Rfree during refinement, 1% of reflections were set aside. The electron density map showed that one loop had a substantially different conformation and was rebuilt manually. This model was refined at 1.35 Å resolution to a final R-factor of 0.130 and Rfree of 0.175. The model contains residues 4 – 757, 1025 water molecules, one sodium ion and one phosphate ion (present at 0.1 M concentration in the mother liquor). This model, in turn, was used to determine the structures of all the complexes of chondroitin AC lyase with chondroitin 4-sulfate tetrasaccharide (CStetra) with different soaking times (30 seconds, two minutes, ten minutes, 35 minutes, two hours, four hours, ten hours) and of the HAtetra complex (two minutes soaking time). It was clear from the height of peaks in the difference electron density map that the occupancies of sugars in sites (22, 2 1) and (þ 1, þ 2) differed systematically from structure to structure, in a manner consistent with the enzyme being active in the crystal. In such a case, the initial concentration of tetrasaccharide would decrease with time, while that of a disaccharide product increases. At any time, active sites of some enzyme molecules might be temporarily empty. Previous structural studies on chondroitin lyase showed that disaccharides were found 383 Chondroitinase AC Crystal Structure and Mechanism predominantly in subsites (2 2, 21), and not in (þ 1, þ2).10 Therefore, at any given time there is a mixture of tetrasaccharides bound to sites (2 2, 2 1, þ1, þ2) and disaccharides bound predominantly to sites (22, 2 1). For that reason, the occupancies of sites (22, 2 1) are expected to be higher than for sites (þ1, þ 2). To obtain an estimate of the relative occupancies at the 2 and þ sites, we assumed that, since the substrate is bound tightly to the enzyme through numerous hydrogen bonds, stacking and van der Waals interactions, the temperature factors of the substrate are similar to those of the residues with which it interacts. We have adjusted independently the occupancies of sites (22, 21) and (þ1, þ2) in steps of 0.1 until the average temperature factor of the substrate was close to that of the surrounding atoms (Table 4). The derived occupancy values provide information about the relative rather than absolute occupancies. In the HAtetra complex, only two sugar rings were visible in the difference electron density map, in positions (22, 2 1) (reaction product). Full statistics of refinements are summarized in Table 3. PROCHECK49 showed that all models have good geometry with no outliers. Protein Data Bank accession numbers Coordinates of the native chondroitinase AC (ArthroACnat), the mercury derivative (ArthroACHg), the two minutes soaked complex with hyaluronan (ArthroACHA), and 30 seconds, ten minutes and ten hours soaked complexes with CStetra (ArthroACCStetra) are deposited in the Protein Data Bank, RCSB, with accession codes 1RW9, 1RWA, 1RWC, 1RWF, 1RWG, 1RWH. Amino acid sequence assignment As mentioned above, the amino acid sequence of the protein has not been previously determined. The high resolution of the diffraction data and the parallel independent refinement of several structures (native, thimerosal derivative and several enzyme – substrate complexes) allowed us to deduce the sequence directly from the electron density maps. This sequence was confirmed and corrected in several places by mass spectrometry analysis. For each of the five best datasets at resolutions 1.25– 1.5 Å, the refinements converged at very low Rfactors of 0.114 – 0.138 (R-free of 0.142– 0.178), indicating highly refined and reliable structures. Electron density-based assignments At 1.3 Å resolution the residue types Trp, Tyr, His, Arg, Lys, Pro, Ile, Leu are defined unambiguously by their shape (Supplementary Material). The three highest peaks in the MAD-derived experimental map corresponded to Hg atoms and each was associated with a cysteine residue. The subsequent 12 highest peaks occur in the middle of the side-chain electron densities and were associated either with a cysteine or a methionine residue, based on the shape of the electron density. The next group of peaks corresponded to main chain or side-chain oxygen atoms. During the refinement, three additional residues on the protein surface were assigned as methionine based on the refined shape of the electron density. A total of seven cysteine and 11 methionine residues were identified. The distinction between Val and Thr could be made based on the height of the electron density for CG2 versus CG1/OG1, often supported by the presence of hydrogen bond donors or acceptors within 3 Å. The assignment of Phe was rather straightforward. The remaining amino acid types, Ser, Ala and Gly, were assigned based on the shape of the electron density. The electron density was convincing for Asx and Glx. Initially, all Asx were refined as Asp and all Glx as Glu. When the R-factor dropped below 0.2 we reassigned these residues to either Gln (Asn) or Glu (Asp) based on: (1) the height of peaks corresponding to the sidechain atoms; (2) the hydrogen bonding network and the directionality of hydrogen bonds; (3) sequence conservation in related protein sequences. All these assignments were re-evaluated during each rebuilding cycle and crosschecked between all independently refined structures. Mass spectrometry-based assignments Mass spectrometry was applied to determine the consistency between the measured molecular mass of the entire molecule and trypsin-derived peptides and the predicted molecular mass based on the established amino acid sequence. The molecular mass of the entire protein was determined on an LC/MS Agilent 1100 mass spectrometer. For MS/MS analysis the protein was in-gel digested with trypsin. Peptides were separated on a CapLC HPLC (Waters, Milford, USA) with 0.1% (v/v) aqueous formic acid and 0.21% (v/v) formic acid in acetonitrile used for the gradient composition. A volume of 20 ml of in-gel digest sample was injected onto the trapping column at a flow-rate of 15 ml/minute and then washed for five minutes with 0.1% (v/v) aqueous formic acid. The peptides from the trapping column were directed to the PicoFrite analytical column (New Objective, MA) filled with 10 cm of C18 BioBasice packing (5 mm, 300 Å, 75 mm ID £ 10 cm). The spraying tip of the PicoFrite column was positioned near the sampling cone of the mass spectrometer and the capillary voltage adjusted to achieve the best plume possible. The analysis was done on a QTOF-2 mass spectrometer (Micromass, UK) upgraded with EPCAS electronics. The instrument was set in data directed analysis (DDA) mode, where an MS survey scan from 350 to 1600 m/z was recorded in one second, then the strongest ion was selected for MS/MS (50 –2000 m/z) for duration of one second. The DDA was set to select doubly and triply charged ions. The instrument then switched back to the MS survey to select the second most intense ion for the next MS/MS spectrum. The mass spectrometric analysis was designed to collect a maximum of different ions to get a maximum of protein coverage. The interscan time was set to 0.1 second. The optimized lyase sequence of 757 amino acid residues when matched with the experimental 317 tandem mass spectra resulted in the assignment of 157 of the 317 tandem MS spectra that identified 37 tryptic fragments (645 residues) that covered 85.2% of the sequence. Mascot search parameters were restricted to oxidation of methionine, histidine and/or tryptophan, alkylation of cysteine by iodoacetamide or in-gel during electrophoresis by acrylamide, and the deamidation of asparagine/ glutamine or the carbamidomethyl derivative of cysteine, with a parent ion mass tolerance of 0.5 Da and allowing for one missed cleavage. Relaxing the stringency of the search to allow for non-specific proteolysis by trypsin (and maintaining the parameters above) resulted in the identification of a further 28 (non-tryptic) fragments defined by 35 tandem MS spectra with a 384 concomitant increase in coverage by 25 amino acid residues to 88.5%, while further decreasing the stringency of the matching by increasing the tolerance on the parent peptide mass to 2 Da resulted in the assignment of another ten spectra to the sequence. In total, 202 of the tandem MS spectra were matched to the best sequence in this manner. The quality of the remaining MS/MS spectra was insufficient for confirmation of sequence although sequence tags50 of low confidence could be generated and about 42 of the spectra were weak, sparse and possibly correspond to background noise. An accounting of the peptides as predicted from the X-ray data but not observed by tandem MS indicated that several short tryptic peptides (residue numbers 44 – 50, 51 – 57, 175– 176, 295– 296, 297– 300, 342– 343, 356– 360, 385– 388, 496– 501, 502, 586– 588, 639, 640– 645, 646– 650, 662– 664, 752– 754, 755– 756, 757) that account for 64 residues were most likely lost during the loading/washing of the reversed-phase LC column, whereas the rather large tryptic fragment (residues 503–525) that is well defined by the X-ray data may have remained in the gel during the robotic in-gel digestion/extraction and/or remained bound to the reversed-phase column throughout the LC-QToF MS analysis. Results of Mascot analysis and samples of MS/MS spectra for several peptides are deposited as Supplementary Material. An example of the experimental and the refined electron density map for the region around Asn183 is shown in Supplementary Material. Finally, the sequence of ArthroAC derived by combining information from electron density maps and mass spectrometry was aligned with related enzymes based on their structural superposition (Figure 1). At completion of this process we were confident of the identity of 746 out of 754 modeled residues (98.9%). For three additional residues (Asn/Asp and Gln/Glu) we are less certain of the assignment and it is based primarily on differences in the B-factors and potential formation of hydrogen bonds. Finally, five residues located in flexible loops exposed to solvent and with B-factors of more than 30 Å2 (average B for the protein is , 15 Å2) and for which no fragment was found in MS/MS data could not be identified unambiguously. These residues are distant from the substrate binding site (marked in Figure 1). At the end of the structure determination process we have determined the N-terminal amino acid sequence of 21 residues of a dissolved crystal by Edman degradation. The first three N-terminal residues are disordered in the structure, the identity of the subsequent 18 residues agrees with the electron-density-based assignments. Kinetic analysis Glycosaminoglycans CS-A (19,000 Da), CS-C (43,000 Da), CS-D (30,000 Da), low molecular mass hyaluronic acid (50,000 Da) and hyaluronic acid (1,000,000 Da) were obtained from Seikagaku Corporation (Tokyo, Japan). Arthro AC was dissolved at concentration of 50 mU/ml in 0.1% (w/v) BSA (solution A). Each chondroitin sulfate and lower molecular mass of hyaluronic acid was dissolved in distilled water at concentration of 10 mg/ml, whereas hyaluronic acid was dissolved in it at concentration of 2 mg/ml, respectively. From 10 ml to 200 ml of solution containing each GAG thus prepared was added to 160 ml of 0.4 M sodium acetate buffer (pH 6.0) and the appropriate amount of distilled water was added to each solution to adjust the final volume to 700 ml (solution B). Each solution B was incubated at 37 8C. After 30 Chondroitinase AC Crystal Structure and Mechanism seconds, 100 ml of solution A was added to solution B and the increased absorbance (ABS) at 232 nm due to the generation of unsaturated bonds in disaccharides caused by Arthro AC was monitored at every 30 seconds by spectrophotometer (UV 160A, Shimadzu, Kyoto, Japan). Reaction velocities (v) at different concentration of the GAGs were plotted as DABS=Dt: Since all GAGs used in the experiment were catalyzed to disaccharide units, we calculated substrate concentrations of the GAGs as molar basis of disaccharide units contained in each GAG. For example, in the case of the CS-A, it was composed of 95.5% of mono-sulfated disaccharide unit (503.3 Da), 3.0% of di-sulfated disaccharide unit (605.3 Da) and 1.6% of non-sulfated saccharide unit (401.3 Da). Therefore, mean molecular mass of the CS-A as disaccharide unit can be calculated to be 505.2. Accordingly, mean molecular mass of the CS-C, the CS-D and both hyaluronic acid from the disaccharide-unit basis were calculated as 511.2 Da, 532.2 Da and 401.3 Da, respectively. By use of substrate concentration based on this calculation, a Lineweaver –Burke plot was obtained, and V and KM of every GAG were determined. Acknowledgements We thank Drs Allan Matte, Joseph D. Schrag and J. Sivaraman for help with data collection, Ms France Dumas for amino acid sequencing, Ms Christine Munger, Dr Daniel Boismenu, Mr Sajid Karsan and Montreal Proteomics Network for mass spectrometry analysis. This work was supported, in part, by CIHR grant 200009MOP84373-M-CFAA-26164 to M.C. and NIH grants GM38060 and HL62244 to R.J.L. References 1. Scott, J. E. (1995). Extracellular matrix, supramolecular organisation and shape. J. Anat. 187, 259– 269. 2. Iozzo, R. V. (1998). Matrix proteoglycans: from molecular design to cellular function. Annu. Rev. Biochem. 67, 609– 652. 3. Lidholt, K. (1997). Biosynthesis of glycosaminoglycans in mammalian cells and in bacteria. Biochem. Soc. Trans. 25, 866– 870. 4. Prydz, K. & Dalen, K. T. (2000). Synthesis and sorting of proteoglycans. J. Cell Sci. 113, 193– 205. 5. Ernst, S., Langer, R., Cooney, C. L. & Sasisekharan, R. (1995). Enzymatic degradation of glycosaminoglycans. Crit. Rev. Biochem. Mol. Biol. 30, 387 –444. 6. McCarter, J. D. & Withers, S. G. (1994). Mechanisms of enzymatic glycoside hydrolysis. Curr. Opin. Struct. Biol. 4, 885– 892. 7. Gacesa, P. (1987). Alginate-modifying enzymes. A proposed unified mechanism of action for the lyases and epimerases. FEBS Letters, 212, 199– 202. 8. Sutherland, I. W. (1995). Polysaccharide lyases. FEMS Microbiol. Rev. 16, 323– 347. 9. Scavetta, R. D., Herron, S. R., Hotchkiss, A. T., Kita, N., Keen, N. T., Benen, J. A. et al. (1999). Structure of a plant cell wall fragment complexed to pectate lyase C. Plant Cell, 11, 1081– 1092. 10. Huang, W., Boju, L., Tkalec, L., Su, H., Yang, H. O., Gunay, N. S. et al. (2001). Active site of chondroitin 385 Chondroitinase AC Crystal Structure and Mechanism 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. AC lyase revealed by the structure of enzyme – oligosaccharide complexes and mutagenesis. Biochemistry, 40, 2359– 2372. Li, S., Kelly, S. J., Lamani, E., Ferraroni, M. & Jedrzejas, M. J. (2000). Structural basis of hyaluronan degradation by Streptococcus pneumoniae hyaluronate lyase. EMBO J. 19, 1228– 1240. Linhardt, R. J., Galliher, P. M. & Cooney, C. L. (1986). Polysaccharide lyases. Appl. Biochem. Biotechnol. 12, 135–176. Hiyama, K. & Okada, S. (1975). Crystallization and some properties of chondroitinase from Arthrobacter aurescens. J. Biol. Chem. 250, 1824– 1828. Tkalec, A. L., Fink, D., Blain, F., Zhang-Sun, G., Laliberte, M., Bennett, D. C. et al. (2000). Isolation and expression in Escherichia coli of cslA and cslB, genes coding for the chondroitin sulfate-degrading enzymes chondroitinase AC and chondroitinase B, respectively, from Flavobacterium heparinum. Appl. Environ. Microbiol. 66, 29 – 35. Pojasek, K., Shriver, Z., Kiley, P., Venkataraman, G. & Sasisekharan, R. (2001). Recombinant expression, purification, and kinetic characterization of chondroitinase AC and chondroitinase B from Flavobacterium heparinum. Biochem. Biophys. Res. Commun. 286, 343–351. Féthière, J., Eggimann, B. & Cygler, M. (1999). Crystal structure of chondroitin AC lyase, a representative of a family of glycosaminoglycan degrading enzymes. J. Mol. Biol. 288, 635 –647. Hiyama, K. & Okada, S. (1977). Action of chondroitinases. III. Ionic strength effects and kinetics in the action of chondroitinase AC. J. Biochem. (Tokyo), 82, 429–436. Hiyama, K. & Okada, S. (1975). Amino acid composition and physiochemical characterization of chondroitinase from Arthrobacter aurescens. J. Biochem. (Tokyo), 78, 1183– 1190. Davies, G. J., Wilson, K. S. & Henrissat, B. (1997). Nomenclature for sugar-binding subsites in glycosyl hydrolases. Biochem. J. 321, 557– 559. Perkins, D. N., Pappin, D. J., Creasy, D. M. & Cottrell, J. S. (1999). Probability-based protein identification by searching sequence databases using mass spectrometry data. Electrophoresis, 20, 3551– 3567. Altschul, S. F., Madden, T. L., Schaffer, A. A., Zhang, J., Zhang, Z., Miller, W. & Lipman, D. J. (1997). Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucl. Acids Res. 25, 3389– 3402. Li, S. & Jedrzejas, M. J. (2001). Hyaluronan binding and degradation by Streptococcus agalactiae hyaluronate lyase. J. Biol. Chem. 276, 41407 –41416. Hashimoto, W., Nankai, H., Mikami, B. & Murata, K. (2003). Crystal structure of Bacillus sp GL1 xanthan lyase, which acts on the side chains of xanthan. J. Biol. Chem. 278, 7663– 7673. Huang, W., Lunin, V. V., Li, Y., Suzuki, S., Sugiura, N., Miyazono, H. & Cygler, M. (2003). Crystal structure of Proteus vulgaris chondroitin sulfate ABC lyase I at 1.9 Å resolution. J. Mol. Biol. 328, 623– 634. Lo, C. L., Ailey, B., Hubbard, T. J., Brenner, S. E., Murzin, A. G. & Chothia, C. (2000). SCOP: a structural classification of proteins database. Nucl. Acids Res. 28, 257– 259. Cleland, W. W., Frey, P. A. & Gerlt, J. A. (1998). The low barrier hydrogen bond in enzymatic catalysis. J. Biol. Chem. 273, 25529– 25532. Jedrzejas, M. J., Mello, L. V., De Groot, B. L. & Li, S. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. (2002). Mechanism of hyaluronan degradation by Streptococcus pneumoniae hyaluronate lyase: structures of complexes with the substrate. J. Biol. Chem. 277, 28287– 28297. Rye, C. S. & Withers, S. G. (2002). Elucidation of the mechanism of polysaccharide cleavage by chondroitin AC lyase from Flavobacterium heparinum. J. Am. Chem. Soc. 124, 9756– 9767. Cleland, W. W. (2000). Low-barrier hydrogen bonds and enzymatic catalysis. Arch. Biochem. Biophys. 382, 1 – 5. Gerlt, J. A. & Gassman, P. G. (1993). Understanding the rates of certain enzyme-catalyzed reactions: proton abstraction from carbon acids, acyl-transfer reactions, and displacement reactions of phosphodiesters. Biochemistry, 32, 11943– 11952. Mello, L. V., De Groot, B. L., Li, S. & Jedrzejas, M. J. (2002). Structure and flexibility of Streptococcus agalactiae hyaluronate lyase complex with its substrate. Insights into the mechanism of processive degradation of hyaluronan. J. Biol. Chem. 277, 36678 – 36688. Yoon, H. J., Hashimoto, W., Miyake, O., Murata, K. & Mikami, B. (2001). Crystal structure of alginate lyase A1-III complexed with trisaccharide product at 2.0 Å resolution. J. Mol. Biol. 307, 9 – 16. Mikami, B., Suzuki, S., Yoon, H. J., Miyake, O., Hashimoto, W., Murata, K. (2002). X-ray structural analysis of algininatelyase A1-III mutants/substrate complexes: activation of a catalytic tyrosine residue by a flexible lid loop. Acta Crystallog. sect. A, 58 (Suppl), C271. Ponnuraj, K. & Jedrzejas, M. J. (2000). Mechanism of hyaluronan binding and degradation: structure of Streptococcus pneumoniae hyaluronate lyase in complex with hyaluronic acid disaccharide at 1.7 A resolution. J. Mol. Biol. 299, 885– 895. Sidhu, G., Withers, S. G., Nguyen, N. T., McIntosh, L. P., Ziser, L. & Brayer, G. D. (1999). Sugar ring distortion in the glycosyl-enzyme intermediate of a family G/11 xylanase. Biochemistry, 38, 5346– 5354. White, A. & Rose, D. R. (1997). Mechanism of catalysis by retaining beta-glycosyl hydrolases. Curr. Opin. Struct. Biol. 7, 645– 651. Sulzenbacher, G., Driguez, H., Henrissat, B., Schulein, M. & Davies, G. J. (1996). Structure of the Fusarium oxysporum endoglucanase I with a nonhydrolyzable substrate analogue: substrate distortion gives rise to the preferred axial orientation for the leaving group. Biochemistry, 35, 15280 –15287. Kuroki, R., Weaver, L. H. & Matthews, B. W. (1993). A covalent enzyme – substrate intermediate with saccharide distortion in a mutant T4 lysozyme. Science, 262, 2030– 2033. Strynadka, N. C. & James, M. N. (1991). Lysozyme revisited: crystallographic evidence for distortion of an N-acetylmuramic acid residue bound in site D. J. Mol. Biol. 220, 401– 424. Strynadka, N. C. & James, M. N. (1996). Lysozyme: a model enzyme in protein crystallography. EXS, 75, 185 –222. Jandik, K. A., Gu, K. & Linhardt, R. J. (1994). Action pattern of polysaccharide lyases on glycosaminoglycans. Glycobiology, 4, 289– 296. Capila, I., Wu, Y., Rethwisch, D. W., Matte, A., Cygler, M. & Linhardt, R. J. (2002). Role of arginine 292 in the catalytic activity of chondroitin AC lyase from Flavobacterium heparinum. Biochim. Biophys. Acta, 1597, 260– 270. 386 43. Otwinowski, Z. & Minor, W. (1997). Processing of X-ray diffraction data collected in oscillation mode. Methods Enzymol. 276, 307– 326. 44. Terwilliger, T. C. & Berendzen, J. (1999). Automated MAD and MIR structure solution. Acta Crystallog. sect. D, 55, 849– 861. 45. Terwilliger, T. C. (2000). Maximum-likelihood density modification. Acta Crystallog. sect. D, 56, 965– 972. 46. Jones, T. A., Zhou, J. Y., Cowan, S. W. & Kjeldgaard, M. (1991). Improved methods for building protein models in electron density maps and the location of errors in these models. Acta Crystallog. sect. A, 47, 110y119. 47. Brünger, A. T., Adams, P. D., Clore, G. M., DeLano, W. L., Gros, P., Grosse-Kunstleve, R. W. et al. (1998). Crystallography and NMR system: a new software suite for macromolecular structure determination. Acta Crystallog. sect. D, 54, 905– 921. 48. Murshudov, G. N., Vagin, A. A. & Dodson, E. J. (1997). Refinement of macromolecular structures by the maximum-likelihood method. Acta Crystallog. sect. D, 53, 240– 255. 49. Laskowski, R. A., MacArthur, M. W., Moss, D. S. & Thornton, J. M. (1993). PROCHECK: a program to check the stereochemical quality of protein structures. J. Appl. Crystallog. 26, 283– 291. 50. Mann, M. & Wilm, M. (1994). Error-tolerant identification of peptides in sequence databases by peptide sequence tags. Anal. Chem. 66, 4390– 4399. 51. Guex, N. & Peitsch, M. C. (1997). SWISS-MODEL and the Swiss-PdbViewer: an environment for comparative protein modeling. Electrophoresis, 18, 2714– 2723. Chondroitinase AC Crystal Structure and Mechanism 52. Kraulis, P. J. (1991). MOLSCRIPT: a program to produce both detailed and schematic plots of protein structures. J. Appl. Crystallog. 24, 946– 950. 53. Merritt, E. A. & Bacon, D. J. (1997). Raster3D: photorealistic molecular graphics. Methods Enzymol. 277, 505– 524. 54. Nicholls, A., Sharp, K. A. & Honig, B. (1991). Protein folding and association: insights from the interfacial and thermodynamic properties of hydrocarbons. Proteins: Struct. Funct. Genet. 11, 281– 296. Edited by I. Wilson (Received 23 September 2003; received in revised form 19 December 2003; accepted 29 December 2003) Supplementary Material comprising Figures showing electron density maps at various contour levels and Tables and plots displaying peptide identification by mass spectrometry is available on Science Direct