throbber
Crystal structure of human dipeptidyl peptidase
`IV in complex with a decapeptide reveals
`details on substrate specificity and tetrahedral
`intermediate formation
`
`KATHLEEN AERTGEERTS, SHENG YE, MIKE G. TENNANT, MICHELLE L. KRAUS,
`JOE ROGERS, BI-CHING SANG, ROBERT J. SKENE, DAVID R. WEBB,1 AND
`G. SRIDHAR PRASAD
`Syrrx Inc., San Diego, California 92121, USA
`(RECEIVED September 26, 2003; FINAL REVISION October 25, 2003; ACCEPTED October 27, 2003)
`
`Abstract
`Dipeptidyl peptidase IV (DPPIV) is a member of the prolyl oligopeptidase family of serine proteases.
`DPPIV removes dipeptides from the N terminus of substrates, including many chemokines, neuropeptides,
`and peptide hormones. Specific inhibition of DPPIV is being investigated in human trials for the treatment
`of type II diabetes. To understand better the molecular determinants that underlie enzyme catalysis and
`substrate specificity, we report the crystal structures of DPPIV in the free form and in complex with the first
`10 residues of the physiological substrate, Neuropeptide Y (residues 1–10; tNPY). The crystal structure of
`the free form of the enzyme reveals two potential channels through which substrates could access the active
`site—a so-called propeller opening, and side opening. The crystal structure of the DPPIV/tNPY complex
`suggests that bioactive peptides utilize the side opening unique to DPPIV to access the active site. Other
`structural features in the active site such as the presence of a Glu motif, a well-defined hydrophobic S1
`subsite, and minimal long-range interactions explain the substrate recognition and binding properties of
`DPPIV. Moreover, in the DPPIV/tNPY complex structure, the peptide is not cleaved but trapped in a
`tetrahedral intermediate that occurs during catalysis. Conformational changes of S630 and H740 between
`DPPIV in its free form and in complex with tNPY were observed and contribute to the stabilization of the
`tetrahedral intermediate. Our results facilitate the design of potent, selective small molecule inhibitors of
`DPPIV that may yield compounds for the development of novel drugs to treat type II diabetes.
`Keywords: Dipeptidyl peptidase IV; DPPIV; CD26; crystal structure; adenosine deaminase binding pro-
`tein; serine protease; tetrahedral intermediate
`
`The type II transmembrane serine protease, DPPIV, also
`known as CD26, or adenosine deaminase binding protein
`(ADAbp), is highly expressed on endothelial cells, differ-
`entiated epithelial cells and lymphocytes (Hegen et al. 1997;
`
`Reprint requests to: G. Sridhar Prasad, Syrrx Inc., 10410 Science Center
`Drive, San Diego, CA 92121, USA; e-mail: Sridhar.Prasad@syrrx.com;
`fax: (858) 550-0526.
`1Present address: Celgene Corp., San Diego, CA 92121, USA.
`Abbreviations: DPPIV, dipeptidyl peptidase IV; NPY, Neuropeptide Y;
`tNPY, N-terminal decapeptide (residues 1–10) of Neuropeptide Y.
`Article published online ahead of print. Article and publication date are at
`http://www.proteinscience.org/cgi/doi/10.1110/ps.03460604.
`
`De Meester et al. 1999; Kahne et al. 1999). A soluble form
`of the enzyme was also found in plasma (Iwaki-Egawa et al.
`1998; Durinx et al. 2000). As a dipeptidyl peptidase, DPPIV
`plays a major role in the regulation of physiological pro-
`cesses including immune, inflammatory, CNS, and endo-
`crine functions. For example, DPPIV plays an important
`role in maintaining glucose homeostasis (Deacon et al.
`1998; Balkan et al. 1999; Pauly et al. 1999; Drucker 2003).
`These studies reveal that DPPIV helps regulate plasma glu-
`cose levels by controlling the activity of the incretins glu-
`cagon-like peptide 1 (GLP-1) and glucose-dependent insu-
`linotropic polypeptide (GIP). Inhibition of DPPIV in wild-
`
`412
`
`Protein Science (2004), 13:412–421. Published by Cold Spring Harbor Laboratory Press. Copyright © 2004 The Protein Society
`
`Page 1 of 10
`
`AstraZeneca Exhibit 2006
` Mylan v. AstraZeneca
` IPR2015-01340
`
`

`
`type and diabetic mice leads to increased levels of
`unprocessed GLP-1 and GIP in the circulation, enhanced
`insulin secretion, and improved glucose tolerance. Selective
`inhibitors of DPPIV improve plasma glucose levels in hu-
`man type II diabetics (Ahren et al. 2002). Independent of its
`dipeptidyl peptidase activity, DPPIV also binds adenosine
`deaminase (ADA; Morrison et al. 1993). This interaction
`has been shown to modulate immune function (Franco et al.
`1998; Morimoto and Schlossman 1998).
`The catalytic triad of DPPIV is composed of residues
`S630, D708, and H740, which are located within the last
`140 residues of the C-terminal region (Ogata et al. 1992).
`The enzyme specifically removes dipeptides from the N
`terminus of peptide substrates that contain on average 30
`residues and have a Pro or Ala in the penultimate position.
`In addition, a slow release has been observed for dipeptides
`composed of X-Ser or X-Gly (Bongers et al. 1992; De
`Meester et al. 1999; Hinke et al. 2000; Lambeir et al. 2002).
`Physiological peptides recognized by DPPIV that contain
`this specificity profile at their cleavage site include neuro-
`peptides like neuropeptide Y, circulating peptide hormones
`like peptide YY, glucagon-like peptides (GLP)-1 and -2,
`gastric inhibitory peptides, as well as paracrine chemokines
`like RANTES (De Meester et al. 1999; Mentlein 1999).
`Catalytic efficiencies for the cleavage by DPPIV of differ-
`ent physiological substrates were determined by mass spec-
`trometry-based protease assays (Lambeir et al. 2001a,b; Zhu
`et al. 2003). These studies demonstrated that residues sur-
`rounding the scissile bond mainly determine the substrate
`selectivity of DPPIV. However, there is supporting kinetic
`evidence that nonconserved residues along the entire length
`of the peptide are involved in long-range interactions that
`play a role in substrate binding and catalysis (Lambeir et al.
`2001a,b, 2002; Zhu et al. 2003).
`Crystal structures of DPPIV in complex with several
`small molecule inhibitors and substrates have been pub-
`lished (Engel et al. 2003; Hiramatsu et al. 2003; Oefner et
`al. 2003; Rasmussen et al. 2003; Thoma et al. 2003). How-
`ever, the exact molecular determinants that contribute to the
`substrate specificity of DPPIV and how substrate peptides
`access the active site remains unclear. To help understand
`the function of DPPIV, we crystallized and solved the X-ray
`crystal structure of the enzyme in both its free form and in
`the presence of the first 10 residues of Neuropeptide Y.
`Neuropeptide Y is a physiological substrate of DPPIV
`widely distributed in the nervous system (Mentlein 1999),
`and involved in cardiovascular homeostasis and the regula-
`tion of insulin release (Ahren 2000; Ghersi et al. 2001). The
`catalytic efficiciency for N-terminal dipeptide cleavage of
`Neuropeptide Y by DPPIV is 3.0 × 106 M−1sec−1 (Mentlein
`et al. 1993). The DPPIV/tNPY structure provides direct evi-
`dence that the decapeptide accesses the active site through a
`side opening, unique to DPPIV, and not through the ␤-pro-
`peller opening. The latter mechanism was suggested for the
`
`DPPIV/decapeptide crystal structure
`
`closely related enzyme prolyl oligopeptidase (POP; Fulop et
`al. 1998, 2000). Our work also provides a detailed under-
`standing of the molecular determinants that contribute to the
`substrate specificity of DPPIV. Moreover, in the DPPIV/
`tNPY crystal structure the peptide was trapped in a tetrahe-
`dral intermediate, and gives new insight into DPPIV en-
`zyme catalysis. Earlier studies provided evidence for the
`existence of a tetrahedral intermediate, which was based on
`structural studies on complexes with small molecule “tran-
`sition-state analog” inhibitors; ab initio quantum mechanics
`(QM), molecular mechanics (MM), and molecular dynam-
`ics (MD) simulations or combined time-resolved/pH jump
`crystallographic studies (Wilmouth et al. 2001; Topf et al.
`2002a,b). Until now, no direct structural evidence of a
`single discrete intermediate formed between a physiological
`substrate and a serine protease has been published.
`
`Results
`
`Structure and domain organization of DPPIV
`
`The crystal structure of the extracellular domain (residues
`39–766) of DPPIV was solved to a resolution of 2.1 Å. The
`structure consists of two domains: an N-terminal 8-bladed
`␤-propeller domain (residues 61–495) and a C-terminal ␣/␤
`hydrolase domain (Nardini and Dijkstra 1999; residues 39–
`55 and 497–766; Fig. 1). The propeller domain packs
`against the hydrolase domain, and the catalytic triad (S630,
`H740, and D708) is at the interface of the two domains. In
`vitro catalytic activity of recombinant DPPIV was mea-
`sured. The catalytic efficiency for the cleavage of the fluo-
`rogenic substrate H-Ala-Pro-7-amido-4-trifluromethylcou-
`marin (Ala-Pro-AFC) by DPPIV is 5.2 × 106 M−1sec−1.
`The asymmetric unit is composed of two homodimers,
`the monomers of which are related by a twofold dyad axis
`(Fig. 1). This dimeric structure correlates with the biologi-
`cally active form of DPPIV (Bednarczyk et al. 1991; De
`Meester et al. 1992; Gorrell et al. 2001; Ajami et al. 2003).
`The overall structures of the monomers are similar with
`root-mean-square deviations (RMSDs) from 0.64 Å to 0.98
`Å for all heavy atoms and from 0.28 Å to 0.56 Å for the C␣
`atoms. The dimer interface buries a total of 2188 Å2 acces-
`sible surface area per monomer and comprises: (1) the last
`␤-strand (␤8) of the peptidase central ␤-sheet, (2) the last
`two ␣-helices (␣G and ␣H), (3) the loop between ␤6 and
`␣E, and (4) the antiparallel ␤-strand subdomain (␤1* and
`␤2*; Fig. 1). ␤8 mainly contains hydrophobic residues
`forming hydrophobic interactions at the center of the dimer
`interface. ␣-Helix H forms hydrogen bonds with the loop
`between ␣G and ␤8 in the other monomer. The antiparallel
`␤-strand arm formed by ␤1* and ␤2* interacts with its
`related arm, ␣G, and the loop between ␤6 and ␣E in the
`other monomer.
`
`www.proteinscience.org
`
`413
`
`Page 2 of 10
`
`

`
`Aertgeerts et al.
`
`Figure 1. Ribbon diagram showing overall structure of the DPPIV homodimer, viewed perpendicular to the twofold dyad axis.
`Secondary structural elements that are involved in dimer formation are represented in red and in blue. The active site residues are shown
`as ball-and-stick representations. The ␣-helix comprising residues E205 and E206 is indicated in gold. The figure was made using the
`programs MOLSCRIPT (Kraulis 1991) and Raster3D (Merrit and Bacon 1997).
`
`The primary structure of DPPIV contains nine potential
`N-linked glycosylation sites: N85, N92, N150, N219,
`N229, N281, N321, N520, and N685. The first N-acetyl-
`glucosamine (NAG) sugar moiety is observed with clear
`electron density in all the nine predicted sites. Detailed
`structural and biochemical analysis revealed that the glyco-
`sylation of DPPIV is not important for catalytic activity,
`homodimer formation and ADA binding (Aertgeerts et al.
`2004).
`
`Substrate access to the active site
`
`Bioactive peptides recognized by DPPIV could theoretically
`access the active site in two possible ways: through an open-
`ing in the propeller domain or via a side opening formed at
`the interface of the ␤-propeller and hydrolase domains (Fig.
`2). The propeller opening is formed by the ␤-propeller do-
`main, which is composed of an unusual eightfold repeat of
`blades. Each blade is composed of a four-strand antiparallel
`␤-sheet. The ␤-propeller domain defines a funnel shaped,
`solvent-filled tunnel that extends from the ␤-propeller’s
`lower face to the active site. The lower face of the funnel,
`distal to the hydrolase domain, has a diameter of approxi-
`mately 15 Å. The closing of the circle between the first and
`the last blade of propeller proteins has been termed “Vel-
`cro” (Neer and Smith 1996), and unlike most of the other
`known propeller proteins, the “Velcro” is not closed be-
`tween the first and the last blades in the DPPIV structure.
`This is similar to the arrangement observed in POP (Fulop
`
`et al. 1998). The propeller opening connects to a larger side
`opening (∼21 Å) formed at the interface of the ␤-propeller
`domain and the hydrolase domain. This oval-shaped cavity
`creates a second entrance to the active site (Fig. 2). To
`understand which entrance/exit pathway substrate peptides
`use to access the active site of DPPIV, we cocrystallized the
`enzyme with YPSKPDNPGE (tNPY), corresponding to the
`first 10 residues of the physiological substrate, Neuropep-
`tide Y. (DPPIV used in the experiment contains a single
`mutation S716A. The catalytic efficiency of this mutant for
`cleavage of Ala-Pro-AFC is 41 × 106 M−1 sec−1, which is
`similar to the value measured for wild-type DPPIV. We also
`obtained crystals of wild-type DPPIV in complex with
`tNPY, but
`the crystals using DPPIV-S716A/tNPY dif-
`fracted to a higher resolution.) Clear continuous electron
`density was observed for the first six of the 10 residues of
`the peptide. Four of the six residues make molecular inter-
`actions (see below), with residues lining the side opening of
`DPPIV. No clear electron density was observed for the last
`four residues because they are solvent exposed and there-
`fore not ordered in the structure. In conclusion, the crystal
`structure of the DPPIV/tNPY complex suggests that physi-
`ological substrates may employ the side opening of DPPIV
`to access the active site.
`
`Substrate specificity
`
`DPPIV cleaves the amide bond after the penultimate N-
`terminal residue (P1, according to Berger and Schechter
`
`414
`
`Protein Science, vol. 13
`
`Page 3 of 10
`
`

`
`DPPIV/decapeptide crystal structure
`
`Figure 2. (A) Surface representation of the ␤-propeller domain only, showing the propeller opening to the active site. The view was
`taken from the interface with the ␣/␤-hydrolase domain and down the pseudo-eightfold axis. The four-strand antiparallel ␤-sheets of
`the eight blades are indicated (␤1–␤8). (B) Surface representation of whole DPPIV molecule, showing the side opening to the active
`site. Residues of DPPIV that make direct molecular interactions with tNPY are colored in both panels. Hydrophobic negatively charged
`and positively charged residues are shown in green, in red, and in blue, respectively. The figures were made with the program MOE
`(MOE, Chemical Computing Group).
`
`1970) of physiological peptides. Oligopeptide N termini are
`recognized by the negatively charged active site residues
`E205 and E206, and are anchored by hydrogen bond for-
`mation with the side chains of the two glutamates (Fig. 3).
`E205 and E206 reside on a short ␣-helix insertion (residues
`200–206) protruding from the ␤-propeller domain and
`pointing toward the active site. The two glutamic acid resi-
`dues are conformationally restrained by salt bridge forma-
`
`tion and hydrogen bond interactions with residues R125,
`Y662, D663, and N710.
`The best catalytic efficiencies for dipeptide cleavage by
`DPPIV was measured for peptides with a Pro or Ala at P1
`(Lambeir et al. 2003; Leiting 2003). The well-defined hy-
`drophobic S1 pocket lined by residues V656, Y631, Y662,
`W659, Y666, and V711 determines this specificity (Figs. 3,
`4). The S2 pocket is hydrophobic and determined by the
`
`Figure 3. Stereo drawing of first six residues of Neuropeptide Y (magenta) and the underlying active site residues of DPPIV (pink)
`that make direct molecular interactions with the peptide. The peptide and selected DPPIV residues are shown as ball-and-stick
`representations. The peptide is not cleaved and trapped in a tetrahedral intermediate by which the carbonyl carbon is covalently linked
`to the active site S630. Hydrogen bonds are indicated as green dashed lines. The figure is made using the programs MOLSCRIPT
`(Kraulis 1991) and Raster3D (Merritt and Bacon 1997).
`
`www.proteinscience.org
`
`415
`
`Page 4 of 10
`
`

`
`Aertgeerts et al.
`
`side chains of residues R125, F357, Y547, P550, Y631, and
`Y666. In this structure we observe two water molecules
`occupying this S2 pocket, and the P2 tyrosine is interacting
`with these waters and partially occupying the S2 site. The
`S1⬘ pocket is flat and not well defined, and the only inter-
`actions observed between the P1⬘ serine and the S1⬘ resi-
`dues are nonspecific van der Waals interactions. The car-
`bonyl oxygen of the P1⬘ serine makes a hydrogen bond with
`R125. The side chain of the P2⬘ lysine packs against the face
`of W629, completely occluding the tryptophane from sol-
`vent, and forms a hydrogen bond with the hydroxyl oxygen
`of Y752 (Fig. 3). Beyond P2⬘, no specific interactions are
`observed between the peptide and the underlying DPPIV
`residues.
`Ser 630 is located on “the nucleophilic elbow” formed
`by residues Gly-Trp-Ser630-Tyr-Gly. This sequence is
`essential for DPPIV activity (Ogata et al. 1992) and con-
`served in the ␣/␤ hydrolase family (Gly-X-Ser-X-Gly). The
`orientation of S630 is maintained by hydrogen bonds
`between the carbonyl oxygen of S630 and the amide of
`Y634, and the amide of S630 and the carbonyl oxygen of
`V653.
`
`Tetrahedral intermediate
`
`In the crystal structure of the DPPIV/tNPY complex, we
`observed that the peptide was not cleaved, but trapped in a
`tetrahedral intermediate (Fig. 5). As expected for tetrahedral
`intermediate formation, the O␥ atom of S630 was found in
`close contact (between 1.6–1.8 Å) with the carbonyl carbon
`
`of the scissile bond. The electron density map contoured at
`3␴ was continuous between the two atoms, and the electron
`density map contoured at 1␴ was discontinuous between the
`O␥ atom of S630 and N␦2 of H740. Comparison of this
`structure with a 2.1 Å structure of the free form of DPPIV
`shows that the hydroxyl group of the active site serine
`(S630) has moved significantly to optimally interact with
`the carbonyl carbon of the scissile bond (Fig. 5B). In addi-
`tion, the imidazole ring of H740 rotates by about 15° along
`the ␹2 torsion (Fig. 5B). The hydrogen bond distance be-
`tween S630 and H740 in the native enzyme is 2.8 Å,
`whereas this distance changes to 3.2 Å in the transition state
`structure. The oxyanion is stabilized by hydrogen bond for-
`mation with the main chain amide of Y631 (∼3.1 Å) and
`with the hydroxyl group of Y547 (∼2.2 Å; Fig. 5B). For-
`mation of such a short, very strong, low-barrier hydrogen
`bond is expected in transition states, which stabilizes inter-
`mediates in enzymatic reactions and lowers the energy of
`transition states.
`To verify our conclusion that the decapeptide was trapped
`in a tetrahedral intermediate, the peptide was omitted from
`the model, the active site S630 was changed to an alanine,
`and the structure was again refined using REFMAC (CCP
`1994). The resultant electron density maps showed unam-
`biguous density for the decapeptide, O␥ atom of S630 and
`the continuous electron density between the O␥ atom of
`S630 and the carbonyl carbon of the scissile bond. The
`asymmetric unit is composed of four independent DPPIV/
`tNPY complexes, and in all four structures, the peptide is
`trapped in the tetrahedral intermediate.
`
`Figure 4. Molecular surface representations showing the interaction of tNPY with DPPIV. Residues of the peptide are shown in ball-and-stick represen-
`tations and DPPIV is shown as a solid surface. (A) Colors represent positive and negative electrostatic potential from blue (electropositive; white, neutral)
`to red (electronegative). (B) Colors represent hydrophobicity (green, polar; yellow, hydrophobic; white, exposed). The figures were made with the program
`MOE (MOE, Chemical Computing Group).
`
`416
`
`Protein Science, vol. 13
`
`Page 5 of 10
`
`

`
`DPPIV/decapeptide crystal structure
`
`Figure 5. Schematic representations showing tetrahedral intermediate formation. (A) 2Fo−Fc electron density map contoured at 1␴ of
`the first six residues of tNPY and of the active site serine (S630). The peptide and the side chain of S630 are shown as ball-and-stick
`representations, and part of the DPPIV molecule is represented as a pink ribbon diagram. (B) Schematic representation showing the
`difference in conformation and hydrogen bond formation of active site residues S630 and H740 between the free form (green) of the
`enzyme and the tetrahedral intermediate (pink). The first three residues of the peptide are shown in gold as ball-and-stick represen-
`tations. Hydrogen bonds are represented as green dotted lines, and measured distances are indicated in angstroms. Part of the DPPIV
`molecule around the active site is represented in pink as a ribbon diagram. The figures were made using the programs MOLSCRIPT
`(Kraulis 1991) and Raster3D (Merrit and Bacon 1997) and XtalView (McRee 1999).
`
`Discussion
`
`DPPIV is an ectopeptidase implicated in the degradation of
`various peptides and hormones including glucagon family
`peptides, neuropeptides, and chemokines (Mentlein 1999).
`The enzyme selectively cleaves dipeptides at the N terminus
`when preferably Pro or Ala is present in the penultimate
`position. Several structural features that were observed from
`the crystal structure of the free form of the enzyme and the
`DPPIV/tNPY complex explain the substrate specificity of
`DPPIV and its mode of interaction with substrate peptides.
`Two channels give access to the active site: a propeller
`opening, and a side opening. Until now, no direct evidence
`has been described on which entrance/exit pathway is uti-
`lized by bioactive peptide substrates. No side opening was
`observed in the structure of POP, and a gating filter mecha-
`nism utilizing the propeller opening was proposed to ex-
`plain its narrow substrate specificity (Fulop et al. 1998).
`Because the first and the last blade of the ␤-propeller do-
`main are not closed, they might partially separate to facili-
`tate substrate access of peptides to the active site. This
`hypothesis was subsequently proven by mutagenesis and
`kinetic data on POP (Fulop et al. 2000). DPPIV and POP
`have the same structural organization of the ␤-propeller
`domain, and DPPIV substrates could therefore theoretically
`utilize the central tunnel formed by this domain to access
`the active site. However, in contrast to the structure of POP,
`a second and larger side opening that facilitates access to the
`active site was observed in the structure of DPPIV. This
`opening is characterized by an oval shaped groove and is
`sterically the most favorable way to enter to and exit the
`
`active site. The DPPIV/tNPY structure provides direct evi-
`dence for peptides to employ this opening to access the
`active site of the enzyme. Six of the 10 residues of the
`substrate were ordered in the crystal structure, and the first
`four residues interact with underlying amino acids present
`in the side opening of DPPIV. The last four residues of the
`decapeptide are solvent exposed, and this observation pro-
`vides further evidence that the propeller opening was not
`utilized, because based on the length and diameter of the
`tunnel, extensive interactions for all 10 residues would have
`been predicted.
`The DPPIV/tNPY structure gives a detailed understand-
`ing of the molecular mechanisms that determine the inter-
`action of the peptide with the residues present in the active
`site of DPPIV. The presence of two glutamates (E205 and
`E206) at the end of an ␣-helical segment that protrudes from
`the ␤-propeller domain into the active site of the enzyme
`determines the aminopeptidase function of DPPIV. Both
`residues are essential for enzyme activity (Abbott et al.
`1999). The Glu motif is conserved in the DPPIV-like gene
`family, and was not found in the structure of POP. The Glu
`motif functions as a recognition site for the N terminus of
`peptide substrates, and anchors the substrate so that only
`dipeptides can be cleaved off. The hydrophobic S1 groove
`is shaped to optimally accommodate and interact with a Pro
`or Ala residue, and explains the strong preference of DPPIV
`for peptides with these amino acids in the penultimate po-
`sition. The S2 subsite preferentially recognizes large hydro-
`phobic and aromatic side chains. With respect to the shape
`and chemical composition of the S1⬘ subsite, we observed
`that most side chains can be modeled into this pocket; how-
`
`www.proteinscience.org
`
`417
`
`Page 6 of 10
`
`

`
`Aertgeerts et al.
`
`ever, charged residues are not preferred due to possible
`unfavorable electrostatic interactions. Because of the ␲
`electron system of W629, the S2⬘ groove is well suited to
`accept large aliphatic side chains. We observed that residues
`P2–P2⬘ of the decapeptide are mainly recognized by DPPIV,
`and substrate recognition does not extend beyond P2⬘. This
`is in agreement with kinetic data measured in vitro on
`chemokines that are substrates for DPPIV (Lambeir et al.
`2001a). However, the adenylyl cyclase-activating peptides,
`pituitary adenylate cyclase-activating polypeptide (PACAP)-
`27, and PACAP38 contain the same 27 amino acids at the N
`terminus, but PACAP38 was processed 15-fold more effi-
`ciently than PACAP27 (Lambeir et al. 2001b; Zhu et al. 2003).
`The only difference between the two peptides is a basic C-
`terminal extension present in PACAP38. These data suggest
`that secondary sites on DPPIV remote from the active site are
`important for substrate binding and catalysis.
`In the crystal structure of the DPPIV/tNPY complex, the
`peptide is trapped in a tetrahedral intermediate that occurs
`during enzyme catalysis. From the electron density maps,
`we observed electron transfer between O␥ of S630 and the
`carbonyl carbon of the scissile bond; the measured distance
`was 1.75 ± 0.15 Å. The theoretical expected distance of this
`bond during the tetrahedral intermediate is ∼1.4 Å (Topf et
`al. 2002a). This suggests that the intermediate is forming,
`but has not yet proceeded to completion. The oxyanion is
`stabilized by hydrogen bond formation with the main chain
`amide group of Y631 and the hydroxyl group of Y547.
`While we were writing our manuscript, Thoma et al. pub-
`lished the crystal structure of DPPIV in complex with di-
`protin A (Ile-Pro-Ile). The tripeptide was also trapped in a
`tetrahedral intermediate and its conformation is comparable
`to the first three residues of tNPY in our structure (Thoma
`et al. 2003).
`The crystal structures of the free form of DPPIV and the
`DPPIV/tNPY complex suggest that physiological peptide
`substrates utilize the side opening, unique to DPPIV, to
`access the active site. The structures also provide a clear
`insight into the different molecular determinants that are
`responsible for the substrate specificity of DPPIV. Further-
`more, in the DPPIV/tNPY complex, the decapeptide was
`trapped in a tetrahedral intermediate and gives thereby di-
`rect structural evidence for its existence and provides a de-
`tailed understanding in the molecular mechanisms that are
`utilized to stabilize the intermediate. The availability of the
`DPPIV/tNPY structure will assist in the rational design of
`highly specific and potent inhibitors that can be used to
`better understand the role of DPPIV, and as potential treat-
`ments for diabetes and related disorders.
`
`Materials and methods
`Protein expression and purification
`The cDNA encoding human DPPIV was isolated by PCR from
`spleen cDNA (Clontech) and the extracellular domain (residues
`
`418
`
`Protein Science, vol. 13
`
`39–766) was cloned into the SmaI site of a modified pFastBacHTb
`vector (Invitrogen). The final construct contains a baculovirus
`gp67 signal peptide followed by a His6 tag fused to the coding
`sequence corresponding to residues 39–766 of DPPIV. Recombi-
`nant baculovirus was generated by transposition using the Bac-to-
`Bac system (Gibco-BRL). Large-scale production of recombinant
`protein was performed by infection of Trichoplusia ni (Hi5) insect
`cells (Gibco-BRL) for 48 h in 5-L Wave Bioreactors (Wave Bio-
`tech). The secreted glycosylated recombinant protein was isolated
`from the cell culture medium by diafiltration using crossflow ul-
`trafiltration followed by passage over a nickel chelate resin (Bind-
`ing buffer: 25 mM Tris [pH 7.9], 400 mM NaCl). The column was
`washed overnight (0.2 mL/min) with 50 mM K2HPO4 (pH 7.9);
`400 mM NaCl; 20 mM Imidazole-HCl, and 0.25 mM TCEP fol-
`lowed by five column volumes (1 mL/min) of 50 mM Tris HCl
`(pH 7.9), 400 mM NaCl and 0.25 mM TCEP. Protein bound was
`eluted with four column volumes of 50 mM Tris-HCl (pH 7.9),
`400 mM NaCl, 200 mM imidazole-HCl, and 0.25 mM TCEP. To
`remove oligomeric forms, the sample was further purified over a
`size-exclusion column (BioSep SEC S3000, 300 × 21.2 mm, Phe-
`nomenex) equilibrated with 25 mM Tris (pH 7.6); 150 mM NaCl;
`0.25 mM TCEP; and 1 mM EDTA. A yield of 16 mg/L culture
`(3 × 106 cells/mL) was obtained. DPPIV substituted with Se-Met
`was produced as described above with the following modifica-
`tions: 16 h following infection the Hi5-infected insect cells were
`centrifuged at 500 × g for 15 min and the pelleted cells resus-
`pended in an equal volume of protein-free methionine-free ESF-
`921 medium (Expression Systems). Following 4 h offurther
`growth Se-Met (Acros) was added to a final concentration of 50
`mg/L. The medium from the infected culture was harvested 64 h
`following viral infection. The Se-Met-substituted protein was pu-
`rified as described above. The incorporation of Se-Met was esti-
`mated to be in the region of 30%–40%. The complex of DPPIV
`with a decapeptide (YPSKPDNPGE; tNPY) (custom synthesized
`by Biopeptide Co., LLC) that corresponds to the first 10 amino
`acids of Neuropeptide Y was formed at pH 7.6 (25 mM Tris, 250
`mM NaCl, 0.25 mM TCEP, 1 mM EDTA) by incubation for 30
`min at room temperature using a 10-fold molar excess of tNPY
`(final concentration 1 mM) over DPPIV (final concentration
`0.1 mM).
`
`Determination of catalytic activity
`
`The determination of the catalytic constants of DPPIV and DPPIV-
`S716A for dipeptide cleavage was performed using a fluorescent
`assay. Enzyme (0.1 nM) was mixed with 0.4–400 ␮M of Ala-Pro-
`AFC (Bachem) in 20 mM Tris (pH 7.4) 20 mM KCl, 0.1 mg/mL
`BSA, and 1% DMSO in a 96-well half-area plate and monitored
`kinetically at Ex400 nm and Em505 nm using Molecular Devices
`SpectraMax Gemini. Assays were performed in duplicate for each
`sample. MDL data analysis toolbox was used for analysis of Mi-
`chaelis-Menten kinetics.
`
`Crystallization and data collection
`
`Wild-type DPPIV, Se-Met DPPIV, and the DPPIV/tNPY complex
`were crystallized at 4°C using Syrrx’s automated Nanovolume
`CrystallisationTM technology (Hosfield et al. 2003). In all cases,
`the reservoir solution was 20% PEG MME 2000, 100 mM Bicine
`(pH 8.0–8.5). Thick plate-shaped crystals appeared in about 5
`days, which grew to about 0.5 mm in longest dimension and vary-
`ing width and thickness. For X-ray data collection, crystals were
`flash-frozen at 100 K using 25% v/v ethylene glycol as a cryo-
`
`Page 7 of 10
`
`

`
`DPPIV/decapeptide crystal structure
`
`Table 1. Heavy atom and Se-Met data statistics for human DPPIV
`
`Unit cell parameters
`a (Å)
`b (Å)
`c (Å)
`␤ (°)
`Wavelength (Å)
`Resolution (Å)
`Total observations
`Unique reflections
`Completeness (%)
`Rsymm (I)
`l/␴ (I)
`Number of sites
`Phasing power (ano)
`Phasing power (iso)
`Figure of merit
`NCS correlation
`
`Native
`
`121.8
`124.1
`144.5
`114.7
`1.0
`2.1
`996,059
`218,087
`96.4 (95.0)
`0.062 (0.524)
`19.8 (2.3)
`
`Se-Met
`
`121.9
`123.0
`145.0
`114.9
`0.97913
`3.0
`656,502
`91,599
`95.8 (94.0)
`0.164 (697)
`10.4 (2.7)
`56
`0.46
`0.50
`
`0.91837
`3.0
`646,108
`91,218
`95.4 (93.0)
`0.170 (0.652)
`8.4 (2.2)
`
`0.61
`0.38
`
`EMTSa
`
`PIPb
`
`122.2
`123.1
`145.8
`114.9
`1.00720
`2.8
`914,648
`96,270
`99.8 (100)
`0.204 (0.716)
`12.4 (1.6)
`8
`1.01
`2.83
`
`121.4
`121.7
`144.2
`114.8
`1.0721
`3.0
`582,365
`76,748
`99.8 (98.3)
`0.119 (0.639)
`16.6 (2.5)
`16
`0.98
`1.35
`
`0.97893
`2.8
`1,323,896
`92,320
`96.9 (96.0)
`0.168 (0.701)
`15.0 (2.4)
`
`1.043
`
`0.385 (acentric), 0.345 (centric)
`0.44 (initial) 0.81 (final)
`
`a Ethylmercurithiosalicylate (EMTS).
`b Di-␮-iodobis(ethylenediamine)diplatinum (PIP).
`
`protectant. Data were collected at Advanced Light Source (ALS)
`and Stanford Synchrotron Laboratory (SSRL) beam lines and pro-
`cessed with both HKL2000 programs and MOSFLM (Otwinowski
`and Minor 1997; Leslie et al. 2002). For heavy atom derivatization,
`the native crystals were soaked in varying concentrations of heavy
`
`atom solutions made in synthetic mother liquor. Extensive screen-
`ing of a large number of heavy atom-soaked crystals resulted in
`isomorphous derivatives: di-␮-iodobis(ethylenedi-
`two useful
`amine)diplatinum (PIP), and ethylmercurithiosalicylate (EMTS).
`In addition, a three wavelength multiple wavelength anomalous
`
`Table 2. Data collection and refinement statistics for wild-type DPPIV and DPPIV/tNPY complex
`
`Crystals
`Space group
`Cell dimension
`
`Solvent content (%)
`Data processing statistics
`Wavelength (Å)
`Resolution range (Å)
`Total reflections
`Unique reflections
`I/␴(I)
`Completeness (%)
`Multiplicity
`Rmerge
`Refinement statistics
`Number of protein/sugar/water residues
`Reflection in working/free set
`Rfactor/Rfree
`Average B-factor (Å2)
`r.m.s. deviation of Bonds (Å)/angle (°) from ideality
`Ramachandran plot
`Residues in most favor region (%)
`Additionally allowed region (%)
`Generously allowed regio

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket