throbber
Molecular Recognition of Ligands in Dipeptidyl Peptidase IV
`
`Current Topics in Medicinal Chemistry, 2007, 7, 609-619
`
`609
`
`Bernd Kuhn*, Michael Hennig, and Patrizio Mattei
`
`F. Hoffmann-La Roche Ltd, Discovery Research Basel, CH-4070 Basel, Switzerland
`
`Abstract: The serine protease dipeptidyl peptidase IV (DPP-IV) is a clinically validated target for the treatment of
`type II diabetes and has received considerable interest from the pharmaceutical industry over the last years.
`Concomitant with a large variety of published small molecule DPP-IV inhibitors almost twenty co-crystal
`structures have been released to the public as of May 2006. In this review, we discuss the structural characteristics
`of the DPP-IV binding site and use the available X-ray information together with published structure-activity
`relationship data to identify the molecular interactions that are most important for tight enzyme-inhibitor binding.
`Optimized interactions with the two key recognition motifs, i.e. the lipophilic S1 pocket and the negatively charged
`Glu 205/206 pair, result in large gains in binding free energy, which can be further improved by additional
`favorable contacts to side chains that flank the active site. First examples show that the lessons learned from the X-
`ray structures can be successfully incorporated into the design of novel DPP-IV inhibitors.
`
`INTRODUCTION
`Prolyl peptidases are a relatively small family of enzymes
`that are able to cleave peptide bonds after a proline residue
`[1]. One of the best characterized members of this class is the
`serine protease dipeptidyl peptidase IV (DPP-IV, EC
`3.4.14.5), which specifically removes N-terminal dipeptides
`from substrates containing proline, and to some extent
`alanine, at the penultimate position [2]. An important in vivo
`active substrate of DPP-IV is the incretin glucagon-like
`peptide 1 (GLP-1) which has a stimulating effect on insulin
`secretion in a meal-dependent manner [3]. As DPP-IV is
`responsible for rapid degradation of GLP-1 levels in plasma
`the concept of DPP-IV inhibition to increase the half-life of
`this hormone and prolong its beneficial effects has been
`pursued as a new potential therapeutic approach to the
`treatment of type 2 diabetes [4-7].
`As discussed in this issue and in several excellent
`reviews [6,7], there has been an explosion of patents and
`publications
`in
`recent years
`in particular from
`the
`pharmaceutical industry. The high interest in DPP-IV is on
`the one hand because it is a clinically validated target. On
`the other hand, the DPP-IV inhibitor binding site seems to
`be highly drugable in the sense that tight, specific binding to
`the enzyme can be achieved with small molecules without
`compromising favorable physico-chemical properties that are
`important for good pharmacokinetics. While no DPP-IV
`inhibitor is on the market yet, several molecules have
`progressed to phase II clinical trials and beyond: vildagliptin
`1 (Novartis, preregistered) [8], sitagliptin 2 (Merck, preregis-
`tered) [9], saxagliptin 3 (Bristol-Myers Squibb, phase III)
`[10], SYR322 4 (Takeda, phase III) [11], denagliptin 5
`(Glaxo SmithKline, phase II) [12], 815541 (GlaxoSmith
`Kline/Tanabe, phase II), PSN9301 (Osi-Prosidion, phase II),
`NVP-DPP728 6 (Novartis, phase II discontinued) [13], and
`P3298 7 (Merck/Probiodrug, phase II discontinued) [14].
`
`*Address correspondence to this author at the F. Hoffmann-La Roche Ltd,
`Discovery Research Basel, CH-4070 Basel, Switzerland; Tel: +41 61
`6889773; Fax: +41 61 6886459; E-mail: bernd.kuhn@roche.com
`
`The overview of the advanced inhibitors 1-7 in Fig. (1)
`shows that structurally diverse molecules are able to interact
`strongly with DPP-IV. For the design of novel DPP-IV
`inhibitors, it is of interest to identify those protein-ligand
`interactions that are crucial for achieving tight binding. More
`generally, a structural understanding why DPP-IV binds so
`many drug-like molecules might be useful for a better
`assessment of protein drugability in the future [15]. These
`questions can be addressed to some extent by X-ray
`crystallography. The first crystal structure of DPP-IV was
`published in 2003 by Rasmussen et al. [16] revealing its
`complex with the inhibitor valine-pyrrolidide 8. As can be
`seen from Fig. (2), the number of structures deposited to the
`Protein Data Bank (PDB) [17] has increased steadily since
`then, most of them have been solved as protein-inhibitor
`complexes [9,18-24]. The abundance of publicly available
`DPP-IV X-ray information provides a good basis to analyze
`the molecular recognition processes in this enzyme.
`This review focuses on three aspects of DPP-IV inhibitor
`interactions. Firstly, important structural information of the
`DPP-IV binding site gleaned from the existing X-ray studies
`is summarized. Secondly, we highlight the central molecular
`recognition interactions as they have emerged from analyses
`of complex crystal structures and chemical probing. Lastly,
`first attempts to translate the lessons learned from the X-ray
`structures into the design of novel chemotypes are reviewed.
`
`GENERAL STRUCTURAL ASPECTS
`Human DPP-IV is a 766 amino acid transmembrane
`glycoprotein consisting of a cytoplasmic tail (residues 1-6), a
`transmembrane region (residues 7-28), and an extracellular
`part (29-766) [16]. The extracellular region can be further
`subdivided into two domains: a) the catalytic domain
`(residues 508-766) which shows an a/b
` hydrolase fold and
`contains the catalytic triad Ser630 – Asp708 – His740 and b)
`an eight-bladed b propeller domain (residues 56-497) which
`also contributes to the inhibitor binding site [19]. DPP-IV is
`enzymatically active as a homodimer and this is also the
`assembly predominantly
`found
`in
`the
`asymmetric
`
` 1568-0266/07 $50.00+.00
`
`© 2007 Bentham Science Publishers Ltd.
`
`Boehringer Ex. 2009
`Mylan v. Boehringer Ingelheim
`IPR2016-01563
`Page 1
`
`

`
`610 Current Topics in Medicinal Chemistry, 2007, Vol. 7, No. 6
`
`Kuhn et al.
`
`O
`
`N
`
`H2N
`
`N
`
`N
`
`F3C
`
`N
`
`F
`
`F
`
`F
`
`N
`
`O
`
`CN
`
`NH
`
`1, Vi ldagli ptin
`IC 50 = 3.5 nM
`Novartis
`
`HO
`
`O
`
`N
`
`O
`
`F
`
`N
`
`N
`
`N
`
`4, S YR -322
`IC50 = 4 nM
`Takeda
`
`N
`
`S
`
`H2N
`
`N
`
`O
`
`7, P3298
`Ki = 123 nM
`M erck/P robiodrug
`
`H2N
`
`N
`
`O
`
`O
`
`NH
`
`HOOC
`
`10, Diprotin A
`apparent IC 50 = 1.1 m M
`
`HO
`
`H2N
`
`N
`
`O
`
`C N
`
`3, Saxagliptin
`IC50 = 0.5 nM
`B ri stol-Myers S quibb
`
`N
`
`O
`
`C N
`
`NH
`
`HN
`
`N
`
`NC
`
`6, NVP-DPP 728
`IC50 = 22 nM
`Novarti s
`
`O
`
`N
`
`N
`
`O
`
`N
`
`N
`
`N
`
`NH
`
`9, B DPX
`Ki = 5.4 m M
`Novo Nordisk
`
`2, S itagliti pin
`IC50 = 18 nM
`Merck
`
`F
`
`F
`
`N
`
`H2 N
`
`O
`
`CN
`
`5, Denaglipt in
`IC50 = 22 nM
`GlaxoS mithKli ne
`
`H2 N
`
`N
`
`O
`
`8, Val-pyrrol idide
`IC5 0 = 2 m M
`
`N
`
`S
`
`N
`
`O
`
`O
`
`NH
`
`12
`IC50 = 39 nM
`Guilford P harma
`
`H2N
`
`11
`IC50 = 40 m M
`S anthera
`
`Boehringer Ex. 2009
`Mylan v. Boehringer Ingelheim
`IPR2016-01563
`Page 2
`
`

`
`Molecular Recognition of Ligands in Dipeptidyl Peptidase IV
`
`Current Topics in Medicinal Chemistry, 2007, Vol. 7, No. 6 611
`
`Fig. (1) Contd….
`
`I
`
`N
`
`H2N
`
`O
`
`CN
`
`13
`Ki = 25 nM
`
`F
`
`F
`
`H2N
`
`N
`
`O
`
`16
`Ki = 61 nM
`P fi zer
`
`N
`
`N
`
`S
`
`N
`
`O
`
`O
`
`O
`
`N
`
`21
`IC5 0 = 0.35 nM
`Eis ai
`
`N
`
`S
`
`F
`
`F
`
`N
`
`H2N
`
`O
`
`C N
`
`14
`IC5 0 = 0.6 nM
`Taisho
`
`H2 N
`
`N
`
`O
`
`15
`IC 50 = 43 nM
`M erck
`
`OMe
`
`C l
`
`MeO
`
`N
`
`N
`
`C l
`
`NH
`
`NH2
`
`17
`IC50 = 0.1 nM
`Roche
`
`R
`
`N
`
`H2 N
`
`O
`
`CN
`
`22a, R=H
`Ki = 13 nM
`22b, R=F
`Ki = 3 nM
`J ohnson & Johnson
`
`N
`
`H
`
`NH2
`
`M eO
`
`OMe
`
`19
`IC5 0 = 500 nM
`R oche
`
`O
`
`N
`
`N
`
`N
`
`N
`
`H2N
`
`23
`IC 50 = 10 nM
`Takeda
`
`Fig. (1). Selected inhibitors of DPP-IV. Ligands for which crystal structure information is available have underlined numbers. For
`these ligands, the location of the S1 pocket is indicated by a curved line, the atoms interacting with the Glu 205/206 dyad are in a
`rectangular box, and the atoms forming a covalent bond to Ser 630 are encircled.
`
`unit of the known X-ray structures. The catalytic site lies in
`a large cavity between the two extracellular domains and can
`be accessed through two active site openings. The crystal
`structure of DPP-IV with a decapeptide suggests
`that
`substrates access the catalytic site through a cleft between the
`two domains [20]. The second opening which is located in
`the b propeller domain might be used for the dipeptidic
`product to leave the enzyme [19].
`The shape and hydrophobic/hydrogen bonding properties
`of the inhibitor binding site of DPP-IV are illustrated in Fig.
`
`(3a,b left) for the bound cyanopyrrolidine 6 and xan-thine
`derivative 9 [22]. We use color coding to distinguish
`hydrogen bond acceptor/donor functionalities and aromatic/
`non-aromatic hydrophobic surface patches. The binding site
`is open in the front for solvent access and a considerable
`number of ligand atoms are in contact with surrounding
`water. Hydrophobic and hydrophilic patches in the environ-
`ment of the ligands are roughly equal in size with hydrogen-
`bond acceptor groups dominating the hydrophilic part. As
`can be seen from the protein-ligand interactions in Fig. (3a,b
`
`Boehringer Ex. 2009
`Mylan v. Boehringer Ingelheim
`IPR2016-01563
`Page 3
`
`

`
`612 Current Topics in Medicinal Chemistry, 2007, Vol. 7, No. 6
`
`Kuhn et al.
`
`Fig. (2). Annual number of structures deposited to the Protein Data Bank. The grey bar indicates the structures deposited in the first
`quarter of 2006.
`
`Fig. (3). Illustration of solvent-accessible surface of the DPP-IV binding site (left) as well as of important protein-ligand interactions
`(right) for two inhibitors. (a) X-ray complex structure of cyanopyrrolidine NVP-DPP728, 6, with human DPP-IV, (b) crystal structure
`of xanthine derivative BDPX, 9, with porcine kidney DPP-IV (PDB-id: 2aj8). Surfaces are colored by hydrophobic and hydrogen
`bonding (HB) properties: HB acceptor (red), HB donor (blue), HB acceptor/donor (magenta), hydrophobic (grey), aromatic
`hydrophobic (green). Dashed red lines indicate protein-ligand hydrogen bonds and the dashed blue line shows the covalent linkage
`between NVP-DPP728 and Ser 630. Residues Tyr 631, Val 656, Trp 659, Tyr 666, and Val 711 lining the S1 pocket in the back are
`removed for the sake of clarity.
`
`right), this is due to backbone carbonyl groups pointing into
`the binding site at the bottom of the pocket and the
`negatively charged side chains of the Glu 205/206 dyad
`which strongly bind to basic groups such as the secondary
`amines of the two ligands. Crystal structures with the low-
`turnover substrate diprotin A (Ile-Pro-Ile, 10) confirmed that
`
`the two carboxylates of the Glu dyad make short hydrogen
`bonds to the N-terminus of the peptide (d = 2.6Å) providing
`a strong ligand recognition motif [19,25]. For
`ligands
`containing primary amines, the third hydrogen bond is
`typically made to the hydroxyl group of Tyr 662.
`
`Boehringer Ex. 2009
`Mylan v. Boehringer Ingelheim
`IPR2016-01563
`Page 4
`
`

`
`Molecular Recognition of Ligands in Dipeptidyl Peptidase IV
`
`Current Topics in Medicinal Chemistry, 2007, Vol. 7, No. 6 613
`
`The two binding modes displayed in Fig. (3) provide a
`good illustration of the interactions made by inhibitors to
`achieve tight binding to DPP-IV. Apart from the ionic
`interactions with the Glu dyad both ligands fill the proline-
`specific S1 pocket in the back with hydrophobic fragments.
`The low micromolar affinity of the small Val-pyrrolidide 8
`and of the b -phenethylamine fragment 11 from Santhera
`Pharmaceuticals [24], shows that substantial binding affinity
`can be gained by optimized interactions with the S1 and Glu
`dyad anchor sites. The
`importance of
`these
`two
`pharmacophores is also underpinned by the fact that almost
`all inhibitor classes share the presence of a lipophilic moiety
`in close proximity to a primary or secondary amine.
`Additional ligand-protein interactions that are seen in
`Fig. (3) are a covalent bond to the catalytic Ser 630 by a
`cyano electrophile mimicking the transition state of peptide
`cleavage, as well as hydrogen bonds to the backbone NH of
`Tyr 631, which forms the oxyanion hole, and to
`the
`backbone carbonyl of Glu 205. The aromatic rings of Phe
`357 and Tyr 547 are exposed to the binding site and provide
`opportunities for additional lipophilic interactions. Each of
`the two ligands in Fig. (3) uses one of these two aromatic
`residues for p -p
` stacking. Finally, the polar Arg 125 and
`Asn 710 side chains are located close to the carbonyl amide
`linking the P1 fragment with the N-terminus of substrates
`(termed P2 amide recognition region throughout this text).
`As observed for the classical serine proteases, few DPP-IV
`inhibitors explore the C-terminal direction of the binding
`
`the
`site (S1’-S2’-…). One noteworthy exception are
`ketopyrrolidines 12, or related ketoazetidines, in which the
`benzothiazole ring is an optimized substituent of the S1’ site
`[26]. Replacing the benzothiazole moiety with a methyl or
`phenyl group renders the molecules inactive, indicating that
`a significant amount of additional binding free energy can be
`gained on the C-terminal end of the scissile bond. In the
`next paragraph, we will use the structural information
`together with examples of published structure-activity
`relationship (SAR) data to investigate the importance of the
`different interaction regions in more detail.
`The large number of X-ray structures allows an assess-
`ment of the flexibility of the DPP-IV binding site, which is
`of relevance for computational drug design. A superposition
`of the available complex crystal structures reveals very little
`movement in the active site (Fig. (4)). Notable exceptions
`are Ser 630 and Tyr 547, which both can adopt two
`conformations, and especially the flexible side chain of Arg
`358 at the end of the binding pocket. The crystallographic
`temperature factors of atoms of Arg 358 are typically high
`and the average position of its side chain is strongly
`influenced by the bound ligand. In contrast, the other
`arginine
`residue, Arg 125,
`shows
`little movement,
`presumably due to the stabilizing effect of its salt bridge
`with Glu 205. The two binding modes in Fig. (3) reveal the
`induced fit effect of the xanthine core which triggers a
`rotation of the aromatic ring of Tyr 547 from its usual
`orientation for better p -p
` interaction.
`
`Fig. (4). Illustration of flexibility in the active site of DPP-IV. Shown are the residues for which different conformations are seen in
`the X-ray complex structures. The cyanopyrrolidine inhibitor 13 is displayed to indicate the positions of the flexible residues
`relative to a bound ligand (PDB-id: 1orw).
`
`Boehringer Ex. 2009
`Mylan v. Boehringer Ingelheim
`IPR2016-01563
`Page 5
`
`

`
`614 Current Topics in Medicinal Chemistry, 2007, Vol. 7, No. 6
`
`Kuhn et al.
`
`MOLECULAR RECOGNITION
`This section is divided into several subparagraphs that
`separately discuss the different interaction motifs used by
`DPP-IV ligands.
`
`Catalytic Ser 630 and Oxyanion Hole
`The catalytic machinery of DPP-IV involves a serine
`nucleophile within the catalytic triad Ser-Asp-His, whose
`sequential order, however, is inverse to
`that found in
`classical serine proteases (His-Asp-Ser) [27]. Several early
`inhibitors have been developed that use an electrophilic
`group, mainly a nitrile, to interact covalently with Ser 630
`[6,7]. The concept of covalent binding to nitriles is well
`known from cysteine protease inhibitors [28]. X-ray studies
`confirmed that the nitrile carbon atom changes its hybridi-
`zation state and is in covalent bond distance from the oxygen
`atom of the Ser 630 side chain (see Fig. (3a)) [23]. The
`increase in binding affinity with the additional -CN group is
`substantial, leading to an up to 1000-fold tighter binding to
`the enzyme [5]. For example, cyanopyrrolidine 6 has an IC50
`of 22 nM while the des-cyano analogue binds only with
`16m M to DPP-IV [29]. Enzymatic and biophy-sical studies
`revealed that the covalent interaction of comp-ound 6 is
`reversible and that the activity of the enzyme is regenerated
`upon release of
`the
`inhibitor [29,30]. Compared
`to
`equipotent
`non-covalent DPP-IV
`inhibitors,
`cyano-
`pyrrolidines show both slower association and dissociation
`rates (compound 6: k on = 2.2 x 106 M - 1s- 1, koff = 7.3 x 10- 3
`s- 1, K D = 3.4 x 10- 9 M- 1) [30].
`Covalent coordination from the catalytic serine residue to
`nitriles or other electrophiles has several potential pitfalls.
`Phosphonates are known as irreversible inhibitors of DPP-IV
`[4]. Inhibitors with nitrile or boronic acid electrophiles have
`a limited chemical stability due to intramolecular cyclization
`with the free amino group [6,7]. To prevent the confor-
`mational change required for intramolecular reaction bulky
`substituents were introduced in proximity to the amino
`
`and/or cyano group (compounds 1, 3, 5). These second-
`generation cyanopyrrolidines show significantly higher
`chemical stability.
`Apart from the transition state mimetics that covalently
`bind to Ser 630, few inhibitors use the oxyanion hole, which
`is composed of the backbone NH of Tyr 631 and the side
`chain OH of Tyr 547, for binding. The only ligands for
`which this interaction is confirmed by crystal structures are
`the xanthines 9 (Fig. (3b)) and the related pyrimidine-2,4-
`dione 4 [31], in which a carbonyl group accepts a hydrogen
`bond from the amide NH of Tyr 631. As hydrogen bonding
`is very sensitive to a correct geometry few chemotypes are
`apparently able to interact both with the hydrogen bond
`donor arrangement of the oxyanion hole and the S1 and Glu
`dyad anchors that are mandatory for tight binding.
`
`S1 Pocket
`The specificity pocket S1 is composed of the side chains
`of Tyr 631, Val 656, Trp 659, Tyr 662, Tyr 666, and Val
`711. It is highly hydrophobic as illustrated in Fig. (5a).
`Overlays of the existing X-ray structures reveal very little
`changes in size and shape of the pocket demonstrating its
`high specificity for proline residues. The rigidity of this
`pocket was probed by several groups through modification of
`the ring size of P1 fragments. In isoleucine analogues of the
`non-covalent inhibitor 8 an order of magnitude lower affinity
`was observed when the pyrrolidine moiety was replaced by a
`piperidine or azetidine [32]. The low tolerance for larger
`rings was confirmed for the covalent 2-ketopyrrolidine 12
`while the 2-ketoazetidine analogues were equipotent [26].
`The close-up view of the S1 pocket in Fig. (5a and b)
`reveals a small hydrophobic niche in the back and suggests
`to introduce some asymmetry into the P1 fragment to mimic
`the shape of this pocket. This higher asymmetry can be
`achieved by introducing a sulfur atom into a 5-membered
`ring, as illustrated by the thiazolidine 7, which is approxi-
`mately 2-fold more active than the corresponding pyrrolidine
`[32]. Other favorable modifications of the pyrrolidine ring
`
`Fig. (5). Close-up view of the S1 pocket illustrating the hydrophobic niche that can be optimally filled with small substituents. (a)
`front view of S1 pocket with bound Val-pyrrolidide 8 (PDB-id: 1n1m). See Fig. (3) for the color coding of the surface. (b) top view of
`S1 pocket showing the 2,4-dichloro phenyl ring of aminopyrimidine 17 (cyan, PDB-id: 1rwq) and the phenyl ring of
`aminobenzoquinolizine 20a (blue).
`
`Boehringer Ex. 2009
`Mylan v. Boehringer Ingelheim
`IPR2016-01563
`Page 6
`
`

`
`Molecular Recognition of Ligands in Dipeptidyl Peptidase IV
`
`Current Topics in Medicinal Chemistry, 2007, Vol. 7, No. 6 615
`
`are cyclopropanation as in 3 or asymmetric substitution with
`fluorine atoms. Highly potent fluoro-substituted pyrrolidines
`have been published by GlaxoSmithKline 5 [12], Taisho 14
`[33], Merck 15 [34], and Pfizer 16.[35] Hulin et al.
`performed a fluorine scan around the pyrrolidine ring in the
`cyclohexylglycine amide series 16 and found that the activity
`highly depends on the position and stereo-configuration of
`the fluorine substitution [35]. A maximal gain in Ki of ~4-
`fold compared to the unsubstituted pyrrolidine could be
`achieved. As fluorine occupies little more space than
`hydrogen this high sensitivity underlines the stringent shape
`constraints of the S1 pocket.
`A considerably larger gain in binding affinity compared
`to the pyrrolidines could be achieved by small substituents
`on aromatic rings in the S1 pocket [21,36,37]. This is
`illustrated by the aminopyrimidines 17, for which the
`location of the 2,4-dichloro-phenyl ring in the S1 pocket is
`depicted in Fig. (5b) [21]. Table (1) shows the SAR for
`small substitutions around this phenyl ring. Slightly bigger
`substitutents than fluorine, such as chlorine or methyl, are
`best -
` in the optimal para position -
` and improve the IC50
`compared to the unsubstituted phenyl by a factor 30-40.
`Substitutions in
`the meta position
`lead
`to
`repulsive
`interactions with the enzyme and are less favorable.
`
`Table 1. Structure-Activity Relationship of 6-Phenyl Subs-
`tituents of Aminopyrimidines 18 [21]. The Numbers
`are IC50 Data for DPP-IV Inhibition [m M]
`
`N
`
`N
`
`6
`
`R
`
`NH
`
`NH2
`
`18 (R = H: 42)
`
`Cl
`
`2.5
`
`31
`
`1.4
`
`OMe
`
`1.5
`
`80
`
`47
`
`F
`
`14
`
`40
`
`18
`
`R =
`
`ortho
`
`meta
`
`para
`
`Me
`
`1.5
`
`20
`
`1.0
`
`The affinity enhancements observed in the aminopyri-
`midines are even surpassed by the substituent effects in the
`aminobenzoquinolizine (amino-BZQ) series (19 and 20a-j).
`The binding mode of the HTS hit 19 in Fig. (6) reveals the
`potential to substitute the flexible n-butyl side chain with
`better fitting P1 fragments [37]. As can be seen in Table (2),
`amino-BZQ’s with different unsubstituted aromatic and
`lactam moieties (20a-d) inhibit DPP-IV in the high nM to
`low m M range. The X-ray structure with compound 20a
`(Fig. (5b)) reveals the position of the phenyl substituent in
`the S1 pocket. A 40- to 50-fold increase is observed through
`methyl substitution at the correct position, as illustrated by
`20e, 20g, and 20h. The most dramatic improvement
`(>1000-fold) is achieved when the only weakly active pyrrole
`20b is converted into 3-methylpyrrole 20f. The importance
`
`of optimally filling small voids in buried cavities has been
`noted previously [38,39]. This optimi-zation concept is
`nicely exemplified on DPP-IV, with a steep SAR for S1
`pocket substituents.
`It is noteworthy that also some polar groups such as
`pyridine 20c and lactam 20d are tolerated in the S1 pocket
`leading to compounds with an overall more balanced polarity
`pattern. The data in Table (2) underscore the importance of
`the correct orientation of the substituents. Inhibitor 20i has a
`methyl group capable of filling the niche in the S1 pocket in
`the same manner as 20g, however, the polar pyridine
`nitrogen is oriented towards the lipophilic environment in
`the back of this pocket. As a consequence, 20i is only
`slightly more active than the unsubstituted 20c. The penalty
`for unmatched polarities is even more drastic in the lactam
`subseries (20j vs. 20h).
`
`P2 Amide Recognition: Arg 125, Asn 710
`The carbonyl of the amide bond connecting the N-
`terminus with the P1 residue in DPP-IV substrates is located
`in a polar, “electrophilic” environment consisting of the side
`chains of Arg 125 and Asn 710. As shown in Fig. (7, green)
`for the peptide mimetic 13, a hydrogen bond with the amido
`NH2 group of Asn 710 is formed. Conversion of the amide
`to a thioamide leads to a reduction of affinity by 2 and
`replacing the amide by a methylene unit makes the molecule
`inactive, confirming the favorable electrostatic interaction of
`the carbonyl dipole with the protein environment [32].
`Attempts in
`the cyanopyrrolidine series to mimic
`the
`geometry and dipole effect of an amide linker by a trans-
`fluoroolefin lead to a reduction of potency [5]. Interestingly,
`the position of the electronegative oxygen atom can be
`reached by ortho-substitution from aromatic P1 fragments,
`as illustrated in Fig. (7) for the compounds 2 and 17. Merck
`reported a 3-4 fold tighter binding to DPP-IV substituting
`the phenyl in various b -phenethylamine series with a fluorine
`at the ortho-position [36,40]. While the terminal amide
`group of Asn 710 is slightly rotated and not involved in a
`hydrogen bond with the ligand, a favorable electrostatic
`
`Fig. (6). Binding mode and important interactions of the
`aminobenzoquinolizine inhibitor 19 with human DPP-IV [37].
`Dashed red lines indicate protein-ligand hydrogen bonds.
`
`Boehringer Ex. 2009
`Mylan v. Boehringer Ingelheim
`IPR2016-01563
`Page 7
`
`

`
`616 Current Topics in Medicinal Chemistry, 2007, Vol. 7, No. 6
`
`Kuhn et al.
`
`Table 2. Structure-Activity Relationship of P1 Substituents of
`Aminobenzoquinolizines [37]. The Numbers are IC50
`Data for DPP-IV Inhibition [nM]
`
`MeO
`
`OMe
`
`N
`
`20b
`9300
`
`N
`
`20f
`5.4
`
`20a
`200
`
`20e
`4.6
`
`R
`
`NH2
`
`N
`
`H
`
`N
`
`20c
`880
`
`20g
`19
`
`N
`
`N
`
`20i
`320
`
`N
`
`O
`20d
`510
`
`N
`
`O
`
`20h
`9.3
`
`N
`
`O
`
`20j
`10000
`
`interaction between the positively charged Arg 125 and the
`Cd+ -Fd-
` dipole moment remains. Using another ortho-
`substituent with high dipole moment, Takeda reported a
`favorable interaction of their 2-cyano group with Arg 125 in
`the pyrimidine-2,4-dione series 4 [31]. For the aminopyri-
`midines (18, Table (1)), small ortho-substitutions (-Me, -Cl,
`-OMe) on the 6-phenyl moiety lead to 17-28 fold lower IC50
`values compared to the unsubstituted ring. The consistent
`gain in affinity with these three substituents indicates that
`the change of torsional angle between the pyrimidine and 6-
`phenyl rings due
`to ortho-substitution might be a
`contributing factor for better protein-ligand fit in this series.
`From the overlays in Fig. (7) and the assembled SAR it can
`be concluded that placing hydrophobic and/or electronegative
`ligand atoms at a very precise location in the vicinity of Arg
`125 and Asn 710 is rewarded with substantial affinity gains.
`
`N-Terminal Recognition: Glu 205, Glu 206, Tyr 662
`Hydrogen bonding interactions with the side chains of
`the two glutamate residues 205 and 206 are, besides the
`filling of the S1 pocket, the second most important anchor
`point for inhibitor binding. Secondary and primary amines
`are recognized by DPP-IV in this region, and for the latter
`ones a third hydrogen bond is formed, typically to Tyr 662.
`This interaction substitutes the binding of the N-terminus of
`
`Fig. (7). Interactions between inhibitors 2 (orange, PDB-id:
`1x70), 13 (green, PDB-id: 1orw), and 17 (blue, PDB-id: 1rwq)
`with the DPP-IV residues Arg 125 and Asn 710 at the P2 amide
`recognition site. The dashed red line indicates the hydrogen
`bond between the carbonyl group of inhibitor 13 and Asn 710.
`
`peptide substrates. The positions of the basic nitrogen atoms
`in the different crystal structures overlap within a sphere of
`only 1.2 Å, making this a very tight pharmacophore
`constraint. Destruction of the hydrogen bond network
`through alkylation of the amino group abolishes activity
`[41]. Furthermore, variation of the basicity of the amine in a
`Roche cyanopyrrolidine series revealed a sharp drop in
`binding affinity by 2 oders of magnitude when the pKa was
`lowered from 7.3 to 6.0 [42]. An interesting class of
`compounds without basic nitrogen are the carbamoyltriazoles
`from Eisai, 21 [43].
`
`Additional Interactions: Phe 357, Tyr 547 and Arg 358
`High nanomolar affinity can be achieved by interactions
`with the S1 pocket, the Glu dyad, and the P2 amide
`recognition site (for ex. 7: Ki = 123nM). However, further
`affinity gains require either covalent binding to the Ser 630
`residue or the exploitation of additional protein-ligand
`interactions. Prime candidates which are used in almost all
`low nanomolar DPP-IV inhibitors are the phenyl rings of the
`two residues Phe 357 and Tyr 547. These are 6-10 Å away
`from S1 pocket and Glu dyad and exposed to the ligand
`binding site. Figures (3a and b) illustrate p -p
` stacking
`interactions between each of the two residues with different
`aromatic ligand fragments. The other energetically favorable
`arrangement, namely aromatic edge-to-face interactions, are
`also seen in crystal structures, one example being the
`interaction between the biphenyl of 22a and Phe 357 [23].
`There are several possibilities to optimize the aromatic
`interactions in both arrangements, as reviewed previously
`[44]. The inhibitor denagliptin 5 is
`interesting as
`it
`presumably makes interactions to both aromatic side chains
`simultaneously. Alternative to aromatic-aromatic interac-
`tions, hydrophobic contacts between Phe 357, Tyr 547 and
`
`Boehringer Ex. 2009
`Mylan v. Boehringer Ingelheim
`IPR2016-01563
`Page 8
`
`

`
`Molecular Recognition of Ligands in Dipeptidyl Peptidase IV
`
`Current Topics in Medicinal Chemistry, 2007, Vol. 7, No. 6 617
`
`large aliphatic groups, such as adamantyl in 1 or 3, can be
`used to achieve low nanomolar IC50 values.
`Lastly, the presence of Arg 358 in close proximity to Phe
`357 makes the positively charged side chain an additional
`interaction partner for substituents on ligand aromatic rings.
`While considerable flexibility of Arg 358 is seen in Fig. (4)
`the observed SAR for several series interacting with Phe 357
`indicates that additional binding free energy can be gained by
`optimizing the electrostatics in this region. For example,
`placing electronegative groups such as trifluoromethyl or
`fluorine next to the positive charge of Arg 358 led to a 4-
`fold increase in binding
`in sitagliptin 2 and in
`the
`cyanopyrrolidine 22b each [9,23].
`
`STRUCTURE-BASED DESIGN OF DPP-IV INHIBI-
`TORS
`The vast majority of known DPP-IV inhibitors have been
`identified by a) derivatization of Xaa-Pro dipeptides, which
`are the products of enzyme cleavage, and b) high-throughput
`screens of compound collections followed by optimization of
`the initial hits [6]. Given
`the
`increasing amount of
`information about the interactions between DPP-IV and
`bound inhibitors from X-ray crystallography during the last
`four years it is timely to review first structure-based design
`attempts for lead finding and optimization.
`A successful example of scaffold hopping from the
`xanthine core of 9 to the quinazolinones 23 through the help
`of molecular modeling and high
`throughput structural
`biology has been reported by Takeda [31]. As shown in Fig.
`(3b), the xanthine moiety of BDPX 9 is involved in a strong
`hydrogen bond with the backbone NH of Tyr 631 and p -
`stacking interactions with Tyr 547, further providing good
`exit vectors to reach the S1 pocket and the Glu 205/206
`dyad. As revealed by a crystal structure, the designed
`quinazolinone scaffold retains these same interactions while
`incorporating carbonyl acceptor function and substituent
`vectors in a single ring. From this series, the inhibitor SYR-
`322 4 with an IC50 of 4 nM emerged, which is currently in
`phase III clinical trials.
`Researchers at Johnson & Johnson identified a series of
`novel biaryl-based DPP-IV inhibitors using a structure-based
`approach [23]. Initial analysis of the X-ray structure of Val-
`pyrrolidide 8 revealed an unoccupied hydrophobic pocket at
`the S2 subsite formed by the side chains of Phe 357 and Arg
`358 (see Fig. (3a)). This cavity could be filled by biaryl
`substituents yielding the cyanopyrrolidine inhibitor 22a with
`a Ki of 13 nM. Further 4-fold improvement was achieved by
`a fluorine atom in the 4-position of the terminal phenyl ring,
`22b. From their X-ray studies, they attributed this increase
`in binding to a better fit to the targeted pocket and stronger
`edge-to-face p -p
` interaction between the biphenyl group and
`Phe 357.
`We have used a structure-based approach to optimize the
`aminobenzoquinolizine class starting from the HTS hit 19
`whose IC 50 is 500 nM [37]. Analysis of its X-ray complex
`structure with DPP-IV (Fig. (6)) revealed the potential for
`tighter binding to the S1 pocket by replacement of the
`flexible n-butyl side chain with more constrained cyclic
`fragments. Virtual screening of monocyclic rings using
`
`docking and similarity search techniques yielded new P1
`substituents for the amino-BZQ scaffold with different
`polarities (Table (2)). As discussed
`in
`the previous
`paragraph, small substitutions at specific positions in these
`rings drastically increased the binding affinity towards DPP-
`IV, the most interesting ones being in the low nanomolar
`potency range and showing both excellent drug-like
`properties as well as promising in vivo activity in rats.
`Finally, virtual screening of larger molecule selections to
`identify novel inhibitor chemotypes for DPP-IV has been
`attempted. After an unsuccessful HTS run yielding no
`suitable starting points for lead generation, a structure-based
`focused screen with emphasis on low molecular weight
`compounds was performed by Ward et al. [45]. Using a
`hierarchical approach of in silico filtering, 3D pharmaco-
`phore matching, docking, and visual inspection a library of
`4000 molecules was assembled and tested at a three times
`higher compound concentration (30 m M) than in the HTS.
`Several novel, but quite weak DPP-IV inhibitors, the most
`potent showing an inhibition of 82% at 30 m M, could be
`identified. Santhera Pharmaceuticals reported results from a
`biased fragment screen with the aim to identify novel S1
`pocket binders [46]. As known DPP-IV inhibitors typically
`have li

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