Crystal structures of HIV-2 protease in complex with inhibitors containing the hydroxyethylamine dipeptide isostere

Similar documents
Chapter 6. X-ray structure analysis of D30N tethered HIV-1 protease. dimer/saquinavir complex

NIH Public Access Author Manuscript J Am Chem Soc. Author manuscript; available in PMC 2008 September 29.

Docking Analysis of Darunavir as HIV Protease Inhibitors

CS612 - Algorithms in Bioinformatics

Supplementary Material

Supplementary Table 1. Data collection and refinement statistics (molecular replacement).

Domain Flexibility in Retroviral Proteases: Structural Implications for Drug Resistant Mutations,

Chymotrypsin Lecture. Aims: to understand (1) the catalytic strategies used by enzymes and (2) the mechanism of chymotrypsin

Peptide hydrolysis uncatalyzed half-life = ~450 years HIV protease-catalyzed half-life = ~3 seconds

Amprenavir complexes with HIV-1 protease and its drug-resistant mutants altering hydrophobic clusters

Chemistry 135, First Exam. September 23, Chem 135, Exam 1 SID:

Lecture 10 More about proteins

Introduction to proteins and protein structure

Biochemistry - I. Prof. S. Dasgupta Department of Chemistry Indian Institute of Technology, Kharagpur Lecture 1 Amino Acids I

(B D) Three views of the final refined 2Fo-Fc electron density map of the Vpr (red)-ung2 (green) interacting region, contoured at 1.4σ.

Statin inhibition of HMG-CoA reductase: a 3-dimensional view

Excerpt from J. Mol. Biol. (2002) 320, :

SUPPLEMENTARY INFORMATION FOR. (R)-Profens Are Substrate-Selective Inhibitors of Endocannabinoid Oxygenation. by COX-2

Supplementary Materials for

Crystal structure of fructose-1,6-bisphosphatase complexed with

Calpain Assays, Nerve injury and Repair

Supporting Information

Amino Acids. Review I: Protein Structure. Amino Acids: Structures. Amino Acids (contd.) Rajan Munshi

Detergent solubilised 5 TMD binds pregnanolone at the Q245 neurosteroid potentiation site.

Supplementary Information Janssen et al.

THE RIGAKU JOURNAL VOL. 23 / 2006, A1-A10 CONSIDERATIONS REGARDING THE ALIGNMENT OF DIFFRACTOMETERS FOR RESIDUAL STRESS ANALYSIS

Transient β-hairpin Formation in α-synuclein Monomer Revealed by Coarse-grained Molecular Dynamics Simulation

This exam consists of two parts. Part I is multiple choice. Each of these 25 questions is worth 2 points.

CHAPTER 9: CATALYTIC STRATEGIES. Chess vs Enzymes King vs Substrate

2. Which of the following amino acids is most likely to be found on the outer surface of a properly folded protein?

Chemical Nature of the Amino Acids. Table of a-amino Acids Found in Proteins

Arginine side chain interactions and the role of arginine as a mobile charge carrier in voltage sensitive ion channels. Supplementary Information

Copyright 2008 Pearson Education, Inc., publishing as Pearson Benjamin Cummings

Bio 100 Serine Proteases 9/26/11

Ionization of amino acids

Catalysis & specificity: Proteins at work

Introduction to Proteomics Dr. Sanjeeva Srivastava Department of Biosciences and Bioengineering Indian Institute of Technology - Bombay

BIO 311C Spring Lecture 15 Friday 26 Feb. 1

PHARMACOPHORE MODELING

Introduction to Protein Structure Collection

Data collection. Refinement. R[F 2 >2(F 2 )] = wr(f 2 ) = S = reflections 326 parameters 2 restraints

SUPPORTING INFORMATION FOR. A Computational Approach to Enzyme Design: Using Docking and MM- GBSA Scoring

Biomolecules: amino acids

Chapter 3: Amino Acids and Peptides

Lecture 15. Membrane Proteins I

Protein Secondary Structure

KDM2A. Reactions. containing. Reactions

Previous Class. Today. Term test I discussions. Detection of enzymatic intermediates: chymotrypsin mechanism

CHAPTER 21: Amino Acids, Proteins, & Enzymes. General, Organic, & Biological Chemistry Janice Gorzynski Smith

!"#$%&' (#%) /&'(2+"( /&3&4,, ! " #$% - &'()!% *-sheet -(!-Helix - &'(&') +,(-. - &'()&+) /&%.(0&+(! - &'(1&2%( Basic amino acids

Lipids: diverse group of hydrophobic molecules

Chemical Mechanism of Enzymes

Secondary Structure North 72nd Street, Wauwatosa, WI Phone: (414) Fax: (414) dmoleculardesigns.com

PAPER No. : 16, Bioorganic and biophysical chemistry MODULE No. : 22, Mechanism of enzyme catalyst reaction (I) Chymotrypsin

SDS-Assisted Protein Transport Through Solid-State Nanopores

Prerequisite Knowledge: Students should have already been introduced to the inputs and outputs of photosynthesis.

Chem 135: First Midterm

SUPPLEMENTARY INFORMATION

Macromolecules of Life -3 Amino Acids & Proteins

Secondary Structure. by hydrogen bonds

Chapter 3. Table of Contents. Section 1 Carbon Compounds. Section 2 Molecules of Life. Biochemistry

Experimental. Crystal data. C 22 H 26 N 2 O 3 M r = Triclinic, P1 a = (3) Å b = (4) Å c = (4) Å = (6) = 100.

Review of Biochemistry

Probing the Potential Glycoprotein Binding Site of Sindbis Virus Capsid Protein With Dioxane and Model Building

Molecular Graphics Perspective of Protein Structure and Function

UNIVERSITY OF GUELPH CHEM 4540 ENZYMOLOGY Winter 2005 Quiz #2: March 24, 2005, 11:30 12:50 Instructor: Prof R. Merrill ANSWERS

Supporting Information

BIOCHEMISTRY I HOMEWORK III DUE 10/15/03 66 points total + 2 bonus points = 68 points possible Swiss-PDB Viewer Exercise Attached

Proteins are sometimes only produced in one cell type or cell compartment (brain has 15,000 expressed proteins, gut has 2,000).

Chem 4B Spring 2018 Exam 3

PEPTIDES and PROTEINS

Proteins and their structure

Workshop on Analysis and prediction of contacts in proteins

Molecular Biology. general transfer: occurs normally in cells. special transfer: occurs only in the laboratory in specific conditions.

Amino acids. (Foundation Block) Dr. Essa Sabi

Objective: You will be able to explain how the subcomponents of

6. The catalytic mechanism of arylsulfatase A and its theoretical investigation

List of Figures. List of Tables

AP Bio. Protiens Chapter 5 1

The role of Ca² + ions in the complex assembling of protein Z and Z-dependent protease inhibitor: A structure and dynamics investigation

Polarization and Circular Dichroism (Notes 17)

A Chemical Look at Proteins: Workhorses of the Cell

Experimental. Crystal data. C 12 H 20 N 4 O M r = Orthorhombic, P a = (3) Å b = (4) Å c = 14.

Proteins consist of joined amino acids They are joined by a Also called an Amide Bond

HOMEWORK II and Swiss-PDB Viewer Tutorial DUE 9/26/03 62 points total. The ph at which a peptide has no net charge is its isoelectric point.

Supporting Information:

BIRKBECK COLLEGE (University of London)

SUPPLEMENTARY INFORMATION

Supplementary Figures

Analysis of Cancer Proteins: Pattern Identification on Escherichia coli Species

b = (17) Å c = (18) Å = (4) = (4) = (4) V = (3) Å 3 Data collection Refinement

Dr. Nafith Abu Tarboush

Supplementary Materials for

Lab 5: Proteins and the small molecules that love them (AKA Computer Modeling with PyMol #2)

For questions 1-4, match the carbohydrate with its size/functional group name:

Enzyme Catalytic Mechanisms. Dr. Kevin Ahern

Chapter 5: Structure and Function of Macromolecules AP Biology 2011

Self-association of α-chymotrypsin: Effect of amino acids

The Binding Mode of by Electron Crystallography

بسم هللا الرحمن الرحيم

Transcription:

Crystal structures of HIV-2 protease in complex with inhibitors containing the hydroxyethylamine dipeptide isostere Liang Tong 1 *, Susan Pav 2, Suet Mui', Daniel Lamarre 3, Christiane Yoakim 4, Pierre Beaulieu 4 and Paul C Anderson 4 Departments of 1 Medicinal Chemistry and 2 Biochemistry, Boehringer Ingelheim Pharmaceuticals Inc., 900 Ridgebury Road, PO Box 368, Ridgefield, CT 06877, USA and Departments of 3 Biochemistry and 4 Chemistry, Bio-Mega/Boehringer Ingelheim Research Inc., 2100 rue Cunard, Laval, Quebec, Canada H7S 2G5 Background: The HIV protease is essential for the life cycle of the virus and is an important target for the development of therapeutic treatments against AIDS. The structures of HIV protease in complex with different inhibitors have helped in understanding the interactions between inhibitors and the protease and in the design and optimization of HIV protease inhibitors. Results: We report here crystal structures at up to 1.7 A resolution of the homodimeric HIV-2 protease in complex with seven inhibitors containing the hydroxyethylamine dipeptide isostere. A novel dimethylphenoxyacetyl group that is present in some of these inhibitors is inserted between residues 48' and 49' in the flap of the protease and residues 29' and 30' (where a prime indicates a residue in the second monomer), which undergo a conformational change to accommodate the phenyl ring of the inhibitor. Conclusions: This study shows that besides the residues in the flap and residues 79-81 in the S 1 substrate-binding pocket which undergo conformational changes upon inhibitor binding, residues 29 and 30 can also adapt their conformation to fit certain inhibitors. Conformational flexibility of the HIV protease plays an important role in inhibitor binding. Structure 15 January 1995, 3:33-40 Key words: AIDS, aspartic proteases, peptidomimetic inhibitors, retrovirus Introduction The protease of HIV-1, the etiological agent of AIDS in America, Europe and Asia, plays an essential role in the life cycle of the virus [1], and has been an important target for pharmaceutical intervention in the treatment of AIDS [2]. Many peptidomimetic inhibitors have been designed based on the transition-state analog concept, replacing the scissile amide bond with, for example, the reduced amide, hydroxyethylene or hydroxyethylamine nonhydrolyzable dipeptide isosteres. The understanding of the interactions between inhibitors and the protease, and the optimization of these inhibitors, have been aided by the vast amount of knowledge available for the structures of HIV-1 protease-inhibitor complexes [3]. This structural information has also helped in the design of a novel series of non-peptide HIV protease inhibitors [4]. HIV-2 is found mostly in western Africa. The HIV-2 protease is a homodimer with 99 amino acid residues in each monomer and shares 50% sequence identity with that of HIV-1 [5]. Like the protease of HIV-1 [1], the HIV-2 protease should be involved in the processing of viral polyproteins and its inhibition should prevent the maturation of infectious virions. Crystal structures of HIV-2 protease-inhibitor complexes have recently become available [6,7]. They show binding modes for the inhibitor molecules that are similar to those of HIV-1 protease. The two proteases show structural differences at residues 15-20, 34-40 and 65-73, regions away from the substrate-binding sites and where the homology between the two proteases is low [71. In this paper we report crystal structures at up to 1.7 A resolution of HIV-2 protease in complex with seven inhibitors containing the hydroxyethylamine dipeptide isostere. The HIV protease can catalyze the hydrolysis of the Phe-Pro, Tyr-Pro and other peptide bonds in the gag and gag-pol polyproteins. The inhibitors used in the current study contain a phenylalanine analog in the P 1 position (as defined in [8]) and a proline homolog, pipecolinic acid, in the Pl' position (Fig. 1) (PC Anderson, et al., & C Yoakim, unpublished data). The inhibitors exhibit differences in the group occupying the P2 and P 3 positions. The group is (quinolinecarbonyl)valyl in inhibitors 1-3 (referred to as the long inhibitors in the following discussions), but this group is replaced by a novel dimethylphenoxyacetyl moiety in the short inhibitors (4-8) (Fig. 1). All of the compounds are potent inhibitors of both HIV-1 and HIV-2 proteases (Fig. 1) (PC Anderson, et al., & C Yoakim, unpublished data). Different modifying groups can be placed on the pipecolinic ring; such groups modulate the bioavailability of these compounds but have little apparent effect on their potency (Fig. 1) (PC Anderson, et al., & C Yoakim, unpublished data). Inhibitor 7, an expected metabolic product of inhibitor 4, is also highly potent against both proteases. *Corresponding author. Current Biology Ltd ISSN 0969-2126 33

34 Structure 1995, Vol 3 No 1 IC 5 0 (nm) Inhibitor R HIV-1 PR HIV-2 PR Chemical structure 1 J <3.2 2.0 N H - 0 R 0 OH 3 o <4.0 10 H 4 <3.3 <2.3 e. <1 <4? 6 _ 1 < 5.7 9.0 S N K 0 1I*~ I H 7 % "j <5.7 11 s\\ 8 O IN 11 7 13 Fig. 1. HIV protease inhibitors. The long inhibitors (1-3), contain a (quinolinecarbonyl)valyl group at the P 2 and P 3 positions and the short inhibitors (4-8), contain a novel dimethylphenoxyacetyl group. Results and discussion The crystal structures In the discussion that follows, the amino acid residues in the first monomer of the protease dimer are numbered 1-99, and those in the second monomer are numbered 1'-99'. A total of seven crystal structures are presented here of HIV-2 protease in complex with inhibitors containing the hydroxyethylamine dipeptide isostere. The resolution of the structures varies between 2.4 A and 1.7 A, and the R- factor varies between 17.9% and 19.8% (Table 1). The structure models have low deviations from ideality in bond lengths and bond angles (Table 1). All the amino acid residues assume energetically favorable main-chain conformations (Fig. 2). The expected error in atomic coordinates [9] for the complex with inhibitor 1, which is refined at 1.7 A resolution, is 0.2 A (data not shown). Having a collection of independently determined structures makes it possible to estimate the coordinate errors by comparing the different structures. For example, the root mean square (rms) deviation in main-chain atoms (N, Ca, C and 0) between the complexes of inhibitors 1 and 2 is 0.22 A. Moreover, such comparisons are important in assessing the significance of any observed differences among the structures. A genuine structural difference should be observed consistently in homologous structures. The two-fold axis relating the protease monomers lies mostly parallel to the ab plane in this tetragonal crystal form. The angle of rotation around this axis is 179.50 for all seven structures studied here. The translation element along the axis is <0.1 A. The orientation of the HIV-2 protease dimer in complex with the long inhibitors differs from that in complex with the short'inhibitors by a rotation of -0.5 around an axis mostly parallel to the c direction of the crystal (Table 2). The crystal structures of the different inhibitor complexes are superimposed onto that of the complex with inhibitor 1 at 1.7 A resolution for comparison (Table 2). The rms differences in the positions of the C. atoms used in this superposition vary from 0.14 A to 0.26 A (Table 2), consistent with the Luzzati analysis [9]. The distribution of the average temperature-factor values for the main-chain atoms of the protease is shown in Fig. 3 for the complex with inhibitor 1. Residues around 25 and 25' have the lowest temperature-factor values, suggesting rather rigid conformations for the catalytic aspartate residues. The temperature-factor values for residues 38-80 are consistently lower for the second monomer than the first monomer. This is probably attributable to crystal packing interactions of some of these residues in the second molecule (39'-41', 45'-56' and 63'-72') as well as the proximity of the quinoline group of the inhibitor to residues 48'-49' (see below). The binding orientation of the inhibitor For some inhibitor complexes of HIV-1 and HIV-2 proteases the inhibitors are present in two orientations in the crystal, related by the local two-fold of the protease dimer [3,6,7]. In the current structures only one orientation of the inhibitors is observed. The 2Fo-F c electron-density map at 1.7 A resolution for the complex

Crystal structures of HIV-2 protease Tong et al. 35 Table 1. Summary of crystallographic data. Inhibitor number 1 2 3 4 5 6 7 Resolution (A) 1.7 2.2 2.4 2.2 2.1 2.1 2.2 Data source (C = CHESS, R = R-Axis) C+R C R C R R R No. of crystals 2 2 1 2 1 2 1 No. of data frames 32 54 45 37 39 48 39 No. of observations 87499 69222 59 005 48 854 72 889 85 326 67 701 No. of unique reflections 22 846 11 300 8499 10 529 11 909 11 303 11 053 Rmerge (o/,)a 7.1 7.8 6.4 8.4 6.4 6.6 5.4 Completeness (%) 80 87 89 80 85 80 87 No. of protein and inhibitor atoms 1562 1561 1562 1554 1555 1556 1556 No. of water molecules 115 53 27 31 51 42 44 No. of reflections in refinement 22017 10158 7810 9870 11288 10610 10200 Final R-factor (%)b 19.6 18.3 17.9 18.5 19.8 19.3 18.0 Rms deviation in bond lengths (A) 0.017 0.017 0.017 0.021 0.019 0.019 0.017 Rms deviation in bond angles () 2.6 2.6 2.9 2.9 2.9 2,8 2.6 Average protein B-values (A 2 ) 14 13 18 11 20 18 19 armerge = Ilhi- - <Ih >/l25lhi, where <Ih> is the average intensity of the different observations for reflection h. br-factor = lif - F! F with inhibitor 1, calculated before the inhibitor was included in the atomic model, is shown for the two alternative orientations of the inhibitor (Fig. 4). There is clear electron density for the quinoline group in one orientation of the inhibitor and little density for the alternative orientation. The pipecolinic ring would be lying mostly outside electron density in the alternative orientation. The electron density for the t-butyl group has clear tri-dentate features whereas in the alternative orientation the electron density can only fit a valine side chain (Fig. 4). The crystal structure of the unliganded HIV-1 protease reveals a dimer with perfect two-fold symmetry [3]; therefore, there is only one orientation in which the inhibitor can bind because the two monomers cannot be distinguished in the unliganded state. The binding of an asymmetric inhibitor will introduce asymmetry in the HIV protease, which is likely to be localized around the inhibitor-binding region. If the region of asymmetry is used for packing contacts in forming the crystal, the protease-inhibitor complexes may be able to pack only in one orientation. Consequently, the inhibitor may be observed only in one orientation in the crystallographic analysis. If the region of asymmetry is not used for packing contacts, the protease molecules can pack in both orientations, and the inhibitor will then appear to be present in both orientations. Table 2. Superposition of HIV-2 protease structures. Inhibitor number 1 2 3 4 5 6 7 No. of C,s in superposition 198 186 190 163 177 166 181 Rotation angle ( ) 0 0.25 0.30 0.84 079 0.78 0.92 Rms deviation (A) 0 0.14 0.20 0.21 0.26 0.26 0.19 The crystal packing contacts of the protease-inhibitor complexes under study here involve residues in the flap (45'-56') close to the P 3 position of the inhibitor. The asymmetry of the P 3 and P 3 ' positions of the inhibitor, and consequently the asymmetry of the S 3 and S 3 ' binding sites, probably lead to the preferential packing of only one orientation of the protease dimers in this crystal form. The 0.5 rotation of the long inhibitor complexes relative to the short inhibitor complexes (see above) is another indication of the effect of the inhibitor structures on the packing of the crystals. Interaction of the long inhibitors with HIV-2 protease Three structures are reported here of HIV-2 protease in complex with the long inhibitors. The three long -180-120 -60 0 60 10 18 0 180.8-80 1 120-120 60 60 x X 'x r 11 X x --120-180 -180-180 -120-60 0 60 120 180 Fig. 2. Ramachandran plot for the complex with inhibitor 1 at 1.7 A resolution. Glycine residues are shown as crosses, prolines as pentagons, and other residues as dots.

36 Structure 1995, Vol 3 No 1 S 30 t 20 I I0 10,,,i,,,,i,T,,,T,,,,T,i,,T,,,.I,,,,F,,I., T,, phi,,,,i... 10 20 30 40 50 60 70 80 90 Residue Number Fig. 3. Plot of the average main-chain (N, C, C and 0) temperature factor values (A 2 ) for the two HIV-2 protease monomers in the complex with inhibitor 1. Thick line, residues 1-99; thin line, residues 1'-99'. Fig. 4. Electron density for inhibitor 1 at 1.7 A resolution. The 2 Fo-FC electron density is shown, calculated before the inhibitor was included in the atomic model. The atomic models for the observed orientation of the inhibitor (thick lines) and the alternative orientation (thin lines), obtained by applying the local twofold axis of the protease, are also shown. The contour level is at l. Electron density for the pyridyl group becomes visible when the contour level is reduced to 0.5u. inhibitors differ only in the chemical modifying group on the pipecolinic ring. Except for the differences in the conformations of these modifying groups (see below), the three structures are generally in excellent agreement with each other. The differences among the structures are mostly within the expected errors of the atomic coordinates. Several residues on the surface of the protein have weak electron density, possibly attributable to flexibility, and can assume different conformations among the structures. The structure of the complex with inhibitor 1 will be used in the following discussions. As has been observed with other peptidomimetic inhibitors [3], the long inhibitors are bound in the extended conformation and occupy the S 3 -S 3 ' pockets (Fig. 5). The amide groups of the inhibitor are hydrogen bonded to the protease dimer either directly or through water molecules (Fig. 6a). Besides water molecule W301 [3], another water molecule (W302) is observed in these structures and makes hydrogen bonds to the t-butyl amido nitrogen of the inhibitor and the main-chain amido nitrogen of residue 29. The hydroxyl of the isostere is located between the catalytic Asp25 and Asp25' residues (Fig. 6b). The planes of the two side-chain carboxylic acid groups are not parallel to each other. One face of the quinoline group of the inhibitor, in the S 3 pocket, is exposed to the solvent. The other face lies against the phenyl group at the P 1 position of the inhibitor and the side chains of Pro81 and Ile82. The edge of the quinoline ring is 4 A away from and perpendicular to the amide bond between residues 48' and 49', and is 5 A from the side chain of Phe53'. The S2 and S2' pockets are highly hydrophobic and can accommodate the valine side chain or the slightly larger t-butyl group of the inhibitor. The pyridyl ring of the modifying group has weak electron density in all seven structures. It is mostly exposed to solvent and generally has few interactions with the protease, consistent with the observation that these varied modifying groups have little apparent effect on the potency of the inhibitors. The pyridyl group can assume several conformations based on the current structure analysis. In one of the conformations, it is parallel to and 4 A away from the 48-49 amide bond. A comparison of the conformations of the two monomers in the complex with inhibitor 1 reveals the asymmetry in the HIV-2 protease dimer. The largest main-chain differences occur at residues 46-55 and 79-82. The former can be ascribed to differences in crystal-packing contacts (see above) and the proximity of the quinoline ring of the inhibitor. The latter involves residues in the S pocket and therefore could be a reflection of the asymmetry between the phenyl and the pipecolinic side chains. Conformational differences are apparent at several side chains on the surface of the protease, possibly representing differences in flexibility or crystal packing. The asymmetry between the monomers is also manifested by the difference in the distribution of the temperature factors (Fig. 3). Conformational flexibility of the HIV-2 protease for binding the short inhibitors The bound conformations of the four short inhibitors are essentially identical except in the regions of the pyridine or pyrimidine groups (Fig. 5). A comparison of the amino acid residues of the protease of the four structures showed no significant differences among them. The structure of the complex with inhibitor 7 will be used in the following discussions. The conformations of the short inhibitors from the P 1 to the P 2 ' positions are similar to those of the long inhibitors (Fig. 5). The hydroxyl of the isostere occupies the same position as that of the long inhibitors. The second water molecule (W302) is also observed in these structures. The dimethylphenoxyacetyl group is bound in a groove between the flap residues 48' and 49' and residues 27'-30' (Fig. 7). Residues 29' and 30' move away from their positions in the complexes with the long inhibitors to

Crystal structures of HIV-2 protease Tong et al. 37 Fig. 5. The binding modes of the inhibitors. (a) Overlap of the bound conformation of the long inhibitors 1 (red), 2 (cyan) and 3 (black). (b) Overlap of the bound conformation of the short inhibitors 4 (black), 5 (cyan), 6 (pink) and 7 (green). (c) Overlap of the bound conformation of the long inhibitor 1 (red) and the short inhibitor 7 (green). accommodate the dimethylphenoxy group. The mainchain atoms of residue Asp29' shift by as much as 1 A and the side-chain atoms of Asp30' shift by as much as 4 A relative to the structure of the inhibitor 1 complex. A smaller conformational change is observed for residues 29 and 30 in the other monomer. Conformational flexibility in residues 29 and 30 has not been observed previously in HIV-1 or HIV-2 proteases. A conformational change at residue 27 was recently reported for an HIV-1 protease-inhibitor complex [10]. However, that conformational change was accompanied by a 10-fold loss in potency of the inhibitor [10], suggesting that residue 27 may not be very flexible. The conformational changes at residues 29-30 occurred without any apparent loss in the potency of the inhibitors (Fig. 1). The residues in the flap region (43'-48') also undergo a slight conformational change, possibly caused by the removal of the quinoline group. The phenyl side chain in the P 1 position rotates slightly (roughly 20 in X2) in the absence of the quinoline ring. The oxygen atom of the phenoxy group is located close to the Cy atom of the valine residue in the long inhibitors. One of the methyl groups of the dimethylphenoxyacetyl moiety probes deeper into the S 2 pocket than the valine or the t-butyl groups, whereas the other methyl group is exposed to solvent. Conformational flexibility of the hydroxyethylamine isostere The inhibitors used in this study do not extend beyond the P2' position, though the pyridyl group on the pipecolinic ring partly occupies the S 3 ' substrate-binding pocket. If a P3' group is present, the opposite stereoisomer is preferred at the hydroxyl position of the isostere for the inhibition of HIV-1 protease [11]. Crystal structure comparisons between the two types of inhibitors show that the hydroxyethylamine dipeptide isostere adopts different conformations and that the proline or pipecolinic ring at the P' position assumes opposite configurations at the nitrogen atom [11]. The superposition of inhibitor 1 and the hydroxyethylamine inhibitor JG-365 bound to HIV-1 protease [12], which contains a P 3 ' residue, is shown in Fig. 8. The torsion angle across the C,-C bond of the Pi residue changes by 1200 such that the hydroxyl groups of the isostere occupy the same position, where they can interact with both catalytic aspartic acid residues. Further torsion-angle changes in the next two bonds of the isostere (1200 and 1700 respectively), and a change in the configuration of the nitrogen atom, bring the carbonyl oxygens of the P' residues to within 1 A of each other. The position of W301 also shifts between the two inhibitor complexes. The t-butyl group of inhibitor 1 is inserted into the S' pocket rather than following the backbone of the P' residue of JG-365. Interestingly, the second water molecule (W302) that bridges the

38 Structure 1995, Vol 3 No 1 (a) and HIV-2 proteases at nanomolar concentrations. The crystal structures of these inhibitors in complex with HIV-2 protease show that they are bound in the extended conformation, similar to what has been observed with other peptidomimetic HIV protease inhibitors. (b) D 29' G 27' D 291 \k2 3.0 W 3 0 2 The flap residues (43-58) of the protease undergo a large conformational change to come into contact with the inhibitor (although the flaps in the unliganded simian immunodeficiency virus (SIV) protease can assume the closed conformation [13]). Residues 79-81, in the S substrate-binding pocket, have also been shown to adopt different conformations upon binding of different inhibitors. The crystal structures reported in this study suggest another region of conformational flexibility in the HIV protease. Shifts in the main-chain and side-chain atoms of residues 29 and 30 enable the binding of the novel, non-peptidic, dimethylphenoxyacetyl group that is present in some of the inhibitors in the current study. Understanding and taking advantage of such conformational flexibility will be important in the design and optimization of inhibitors of HIV protease. Fig. 6. (a) Hydrogen-bonding interactions between inhibitor 1 and the protease. (b) The interactions between the hydroxyl of the isostere and the catalytic aspartic acid residues. interaction between inhibitor 1 and the protease occupies the same position as the carbonyl oxygen of the P,' residue of JG-365. The conformational flexibility of the hydroxyethylamine isostere is important for the tight binding of these two types of inhibitors to HIV protease. Biological implications The protease of HIV plays an essential role in the processing of viral polyproteins, and the assembly and maturation of infectious virions. The inhibition of HIV protease therefore represents an important therapeutic area in the treatment of AIDS. Many highly potent inhibitors have been designed based on the transition-state analog concept, which replaces the scissile P-P 1 ' amide bond with a nonhydrolyzable isostere with tetrahedral geometry. The inhibitors used in the current study contain the hydroxyethylamine dipeptide isostere, and are active against HIV-1 Materials and methods Crystallization The cloning, expression, purification and crystallization of HIV-2 protease have been described earlier [14]. Briefly, purified HIV-2 protease was concentrated to 6 mg ml - i in 10% (v/v) aqueous Dulbecco's phosphate-buffered saline (GIBCO/BRL, 310-4200), (ph 7.4), containing 0.02% (w/v) NaN 3. A stock solution of the inhibitor (20 mg ml - 1) in dimethylsulfoxide (DMSO) was further diluted with 10% (v/v) DMSO in water to 40-50-fold molar excess of the protein. Final dilution of the inhibitor into the enzyme solution resulted in a 4:1 to 6:1 molar excess of the inhibitor, with a final DMSO concentration of 4-6%. The enzyme/inhibitor mixture was incubated for 1 h at room temperature. Crystals were grown by the hanging-drop vapor-diffusion method. The precipitant solution contained 0.3-0.5 M NaCl, 0.1 M sodium acetate ph 5.4, and 0.02% NaN 3. Crystals appeared within a week but could take several months to reach a usable size. Except for the complex with inhibitor 8 (see below), crystals grew as thin tetragonal prisms. The largest crystal had dimensions of 0.1 mmxo.1 mmxl.5 mm, though generally crystals of size 0.06 mmx0.06 mmxl.0 mm were used for data collection. The crystals belong to space group P4 1 2 1 2 or P4 3 2 1 2, with a=b=62.6 A and c=115.8 A. The crystals contain one HIV-2 protease dimer per asymmetric unit. Crystals of the complex with inhibitor 8 grew under the same condition as a cluster of five prisms arranged like a flower. They belong to space group P2 1 3 with a=b=c=115.4 A. The crystals contain two HIV-2 protease dimers per asymmetric unit. These crystals diffract X-rays poorly and a reflection data set to 3.2 A resolution was collected with an R-Axis system. The structure was determined by molecular replacement and the inhibitor molecules could be located in the 2F-F c electron-density maps.

Crystal structures of HIV-2 protease Tong et al. 39 Fig. 7. Stereo plot of conformational changes in the HIV-2 protease in complexes with the short inhibitors. Residues 27'-31' and the (quinolinecarbonyl)valyl group of the complex with inhibitor 1 (red) are shown superimposed on the corresponding residues and the dimethylphenoxyacetyl group of the complex with inhibitor 7 (green). '1',' Fig. 8. Stereo plot of the superposition of inhibitor 1 (thin lines) and JG-365 (thick lines). The water molecules (W301 and W302) are shown as crosses. However, because of the relatively low resolution of the structure, this complex will not be discussed further in this paper. Data collection and processing X-ray diffraction data for the complexes with inhibitors 1, 3, 5, 6 and 7 were collected at 4C on an R-Axis imaging plate detector mounted on a Rigaku RU-200 X-ray generator operated at 50 kv and 100 ma. The crystal-to-detector distance was 100-115 mm. An oscillation range of 1.5 per frame was used and the exposure time per frame was 75-110 min. The reflection data were processed, scaled and merged with the software provided with the R-Axis system. The data processing statistics are summarized in Table 1. X-ray diffraction data for the complexes with inhibitors 1, 2 and 4 were collected at 4C on the F-1 beamline at the Cornell High Energy Synchrotron Source (CHESS). The X-ray wavelength was 0.91 A. The diffraction pattern was recorded on Fuji imaging plates, which were scanned at a raster of 0.1 mm. The crystal-to-detector distance was 160-220 mm and the exposure time per frame of 2 oscillation varied between 25 s and 60 s. The data frames were integrated with the Rossmann processing package [15]. The full observations of the frames were scaled and merged with the program PRO- TEIN [16]. The data-processing statistics are given in Table 1. For the complex with inhibitor 1, the 1.7 A data set from CHESS was scaled and merged with the 2 A data set from the R-Axis to improve the completeness of the reflection data. Structure determination by molecular replacement The crystal structure of the complex with inhibitor 2 was determined first and was used as the starting model for the other inhibitor complexes. The model for the HIV-2 protease dimer (Protein Data Bank [17] entry 2PHV [5]) was placed in a P1 cell with a=b=c=100 A and ot=3='y = 9 0. Reflection data

40 Structure 1995, Vol 3 No 1 between 10 A and 3.5 A resolution were used in the rotation function calculation with the program GLRF [18]. The radius of integration was 20 A and 23% of the reflections were saved as large terms. Two peaks were found in the rotation function map, at Eulerian angles of (48, 48, 316 ) and (39, 48, 316 ) and with heights of 999 (arbitrary scale, 5.8cr above the average) and 976 (5.7(T) respectively. The next peak in the rotation function had a height of 676 (3.6(r). The first peak would orient the local two-fold axis of the HIV-2 protease dimer along the [110] direction of the unit cell, but could not lead to a translation function solution in either space group. A clear translation function solution, at (0.900, 0.411, 0.033), was obtained with the second orientation in space group P4 3 2 1 2. Reflection data between 8 A and 3.5 A resolution were used in the Patterson correlation translation-function calculation [19]. The height of the second peak in the translation function was 64% of that of the top peak. Reasonable packing of the molecules in the unit cell provided further support for the correctness of the solution. Structure refinement For the complex with inhibitor 2, rigid-body refinement with the program X-PLOR [20] using reflections between 6 A and 4 A resolution lowered the R-factor from 45.5% to 41.4%. The angle of rotation relating the two monomers, modeled to be 178.5 based on the HIV-1 protease structures [5], was changed to 179.6 by this rigid-body refinement. A simulatedannealing slow-cooling refinement [20] reduced the R-factor to 21.0% for reflection data between 6.0 A and 3.0 A resolution. The atomic model was examined against the 2Fo-F c electron density map and rebuilt using the program FRODO [21]. The resolution of the data was extended to 2.5 A and then 2.2 A in subsequent rounds of simulated-annealing refinement and model rebuilding. Residues 37-41 in both monomers were refitted against omit maps after simulated-annealing refinement in their absence. The R factor is 25.6% for 10158 reflections between 6 A and 2.2 A resolution. The electron density for the inhibitor and the conserved W301 [3] was clearly visible in the 2Fo-F c map. A cycle of simulated-annealing refinement in the presence of the inhibitor, followed by restrained individual temperature-factor refinement, lowered the R-factor to 21.7%. Solvent water molecules were selected from difference electron-density maps and their positions and temperature factors were refined. The final atomic model was obtained after one cycle of positional refinement with TNT [22]. For the structure refinement of the other complexes, the structure of the complex with inhibitor 2 was used as the initial atomic model. The inhibitor and the water molecules were removed and the atomic temperature factors were set to an overall value. The inhibitor molecule and W301 were built into the 2Fo-F c electron-density map after one cycle of simulated-annealing slow-cooling refinement. After restrained individual temperature-factor refinement, solvent water molecules were identified from the difference electron-density map. A final round of refinement with TNT produced the current atomic model. The refinement statistics are summarized in Table 1. The atomic coordinates have been deposited at the Brookhaven Protein Data Bank [17] (entry names 11DA and 1DB). Acknowledgements: We thank Dr Karl Hargrave and Thomas Warren for their help with data collection at CHESS. We thank Professor Michael G Rossmann for providing some beam time on the F-1 station (0pm-lam on New Year's Eve, 1993) so that one of us (LT), with the help of Dr Hao Wu, could collect the data on the complex with inhibitor 2. We thank Michael Sintchak for his help with some of the initial crystallographic characterization of the crystals. We thank Drs Karl Hargrave, Christopher Pargellis, Peter Grob, Peter Farina, Julian Adams and Yvan Guindon for their interest and encouragement. References 1. Kohl, N.E., et al., & Sigal, I.S. (1988). Active human immunodeficiency virus protease is required for viral infectivity. Proc. Natl. Acad. Sci. USA 85, 4686-4690. 2. Mitsuya, H., Yarchoan, R. & Broder, S. (1990). Molecular targets for AIDS therapy. Science 249, 1533-1544. 3. Wlodawer, A. & Erickson, J. (1993). Structure-based inhibitors of HIV- I protease. Annu. Rev. Biochem. 62, 543-585. 4. Lam, P.Y.S., et al., & Erickson-Viitanen, S. (1994). Rational design of potent, bioavailable, nonpeptide cyclic ureas as HIV protease inhibitors. Science 263, 380-384. 5. Gustchina, A. & Weber, I.T. (1991). Comparative analysis of the sequences and structures of HIV-1 and HIV-2 proteases. Proteins 10, 325-339. 6. Mulichak, A.M., et al., & Watenpaugh, K.D. (1993). The crystallographic structure of the protease from human immunodeficiency virus type 2 with two synthetic peptidic transition state analog inhibitors. J. Biol. Chem. 268, 13103-13109. 7. Tong, L., Pav, S., Pargellis, C., Do, F., Lamarre, D. & Anderson, P.C. (1993). Crystal structure of human immunodeficiency virus (HIV) type 2 protease in complex with a reduced amide inhibitor and comparison with HIV-1 protease structures. Proc. Natl. Acad. Sci. USA 90, 8387-8391. 8. Schechter, I. & Berger, A. (1967). On the size of the active site in proteases.. Papain. Biochem. Biophys. Res. Commun. 27,157-162. 9. Luzzati, V. (1952). Traitement statistique des erreurs dans la determination des structures cristallines. Acta Crystallogr. 5, 802-810. 10. Hosur, M.V., et al., & Erickson, J.W. (1994). Influence of stereochemistry on activity and binding modes for C 2 symmetry-based diol inhibitors of HIV-1 protease. J. Am. Chem. Soc. 116, 847-855. 11. Krohn, A., Redshaw, S., Ritchie, J.C., Graves, B.J. & Hatada, M.H. (1991). Novel binding mode of highly potent HIV-proteinase inhibitors incorporating the (R)-hydroxyethylamine isostere. J. Med. Chem. 34, 3340-3342. 12. Swain, A.L., et al., & Wlodawer, A. (1990). X-ray crystallographic structure of a complex between a synthetic protease of human immunodeficiency virus 1 and a substrate-based hydroxyethylamine inhibitor. Proc. Nat. Acad. Sci. USA 87, 8805-8809. 13. Wilderspin, A.F. & Sugrue, R.J. (1994). Alternative native flap conformation revealed by 2.3 A resolution structure of SIV proteinase. J. Mo. Bio. 239, 97-103. 14. Pay, S., Lubbe, K., Do, F., Lamarre, D., Pargellis, C. & Tong, L. (1994). Microtube batch protein crystallization: application to human immunodeficiency virus type 2 (HIV-2) protease and human renin. Proteins 20, 98-102. 15. Rossmann, M.G. (1979). Processing oscillation diffraction data for very large unit cells with an automatic convolution technique and profile fitting.. Apple. Crystallogr. 12, 225-238. 16. Steigemann, W. (1974). Ph.D. Thesis. Technische University, Munich. 17. Bernstein, F.C., et a., & Tasumi, M. (1977). The Protein Data Bank: a computer based archival file for macromolecular structures. J. Mol. Bio. 112, 535-542. 18. Tong, L. & Rossmann, M.G. (1990). The locked rotation function. Acta Crystallogr. A 46, 783-792. 19. Tong, L. (1993). REPLACE, a suite of computer programs for molecular-replacement calculations. J. Apple. Crystallogr. 26, 748-751. 20. Bronger, A.T. (1992). X-PLOR Manual, Version 3.0. Yale University, New Haven, CT. 21. Jones, T.A. (1978). A graphics model building and refinement system for macromolecules. J. Apple. Crystallogr. 11, 268-272. 22. Tronrud, D.E., Ten Eyck, L.F. & Matthews, B.M. (1987). An efficient general-purpose least-squares refinement program for macromolecular structures. Acta Crystallogr. A 43, 489-501. Received: 28 Sep 1994; revisions requested: 17 Oct 1994; revisions received: 21 Oct 1994. Accepted: 24 Oct 1994.