10.12.2012 Views

netLibrary - eBook Summary Structure-based Drug Design by ...

netLibrary - eBook Summary Structure-based Drug Design by ...

netLibrary - eBook Summary Structure-based Drug Design by ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>netLibrary</strong> - <strong>eBook</strong> <strong>Summary</strong><br />

University of North Florida <strong>eBook</strong> Collection<br />

<strong>Structure</strong>-<strong>based</strong> <strong>Drug</strong> <strong>Design</strong><br />

<strong>by</strong> Veerapandian, Pandi.<br />

New York Marcel Dekker, Inc., 1997.<br />

ISBN: 0824798694<br />

<strong>eBook</strong> ISBN: 0585157448<br />

Subject: <strong>Drug</strong>s--<strong>Design</strong>. <strong>Drug</strong>s--<strong>Structure</strong>-activity<br />

relationships. <strong>Drug</strong>s--Conformation. <strong>Drug</strong> <strong>Design</strong>.<br />

<strong>Structure</strong>-Activity Relationship.<br />

Language: English<br />

Add to your My Favorites list<br />

Recommend this <strong>eBook</strong> to a friend<br />

<strong>netLibrary</strong> <strong>eBook</strong> Warranty Disclaimer<br />

You are here: home > <strong>eBook</strong> summary<br />

Read this <strong>eBook</strong><br />

Browse this <strong>eBook</strong> online<br />

(borrow for a short time)<br />

Check out and read online<br />

(add to "<strong>eBook</strong>shelf")<br />

Log out.<br />

Hello, Tony.<br />

Go to your<br />

● <strong>eBook</strong>shelf - 0<br />

● My Favorites - 11<br />

● Bookmarks - 0<br />

● Notes - 0<br />

home | search tools | reading room | about us | help | log out<br />

© 2001 - 2004, <strong>netLibrary</strong>, a division of OCLC Online Computer Library Center, Inc. All rights reserved. privacy statement | terms of use<br />

http://legacy.netlibrary.com/ebook_info.asp?product_id=12640&piclist=19799,20772,39801,42375 [4/9/2004 12:08:55 AM]


<strong>Structure</strong>-<strong>based</strong> <strong>Drug</strong> <strong>Design</strong><br />

Return to:<br />

home | | <strong>eBook</strong> summary<br />

<strong>Structure</strong>-<strong>based</strong> <strong>Drug</strong> <strong>Design</strong><br />

Table of Contents<br />

<strong>Structure</strong>-Based <strong>Drug</strong> <strong>Design</strong><br />

Preface<br />

Contents<br />

Contributors<br />

1 Inhibitors of HIV-1 Protease<br />

Help<br />

2 Structural Studies of HIV-1 Reverse<br />

Transcriptase and Implications for<br />

<strong>Drug</strong> <strong>Design</strong><br />

3 Retroviral Integrase: <strong>Structure</strong> as a<br />

Foundation for <strong>Drug</strong> <strong>Design</strong><br />

4 Bradykinin Receptor Antagonists<br />

5 <strong>Design</strong> of Purine Nucleoside<br />

Phosphorylase Inhibitors<br />

6 Structural Implications in the <strong>Design</strong><br />

of Matrix-Metalloproteinase Inhibitors<br />

7 <strong>Structure</strong>—Function Relationships in<br />

Hydroxysteroid Dehydrogenases<br />

8 <strong>Design</strong> of ATP Competitive Specific<br />

Inhibitors of Protein Kinases Using<br />

Template Modeling<br />

9 Structural Studies of Aldose<br />

Reductase Inhibition<br />

10 <strong>Structure</strong>-Based <strong>Design</strong> of<br />

Thrombin Inhibitors<br />

11 <strong>Design</strong> of Antithrombotic Agents<br />

Directed at Factor Xa<br />

12 Polypeptide Modulators of Sodium<br />

Channel Function as a Basis for the<br />

Development of Novel Cardiac...<br />

13 Rational <strong>Design</strong> of Renin Inhibitors<br />

http://legacy.netlibrary.com/reader/reader.asp?product_id=12640 (1 of 2) [4/9/2004 12:09:15 AM]<br />

Page #


<strong>Structure</strong>-<strong>based</strong> <strong>Drug</strong> <strong>Design</strong><br />

14 Structural Aspects in the Inhibitor<br />

<strong>Design</strong> of Catechol O-<br />

Methyltransferase<br />

15 Antitrypanosomiasis <strong>Drug</strong><br />

Development Based on <strong>Structure</strong>s of<br />

Glycolytic Enzymes<br />

16 Progress in the <strong>Design</strong> of<br />

Immunomodulators Based on the<br />

<strong>Structure</strong> of Interleukin-1<br />

17 <strong>Structure</strong> and Functional Studies of<br />

Interferon: A Solid Foundation for<br />

Rational <strong>Drug</strong> <strong>Design</strong><br />

18 The <strong>Design</strong> of Anti-Influenza Virus<br />

<strong>Drug</strong>s from the X-ray Molecular<br />

<strong>Structure</strong> of Influenza Virus Ne...<br />

19 Rhinoviral Capsid-Binding<br />

Inhibitors: Structural Basis for<br />

Understanding Rhinoviral Biology and<br />

f...<br />

20 The Integration of <strong>Structure</strong>-Based<br />

<strong>Design</strong> and Directed Combinatorial<br />

Chemistry for New Pharmaceut...<br />

21 <strong>Structure</strong>-Based Combinatorial<br />

Ligand <strong>Design</strong><br />

22 Peptidomimetic and Nonpeptide<br />

<strong>Drug</strong> Discovery: Impact of <strong>Structure</strong>-<br />

Based <strong>Drug</strong> <strong>Design</strong><br />

Index<br />

http://legacy.netlibrary.com/reader/reader.asp?product_id=12640 (2 of 2) [4/9/2004 12:09:15 AM]


Document<br />

1<br />

Inhibitors of HIV-1 Protease<br />

Krzysztof Appelt<br />

Agouron Pharmaceuticals, Inc., San Diego, California<br />

I. Introduction<br />

Page 1<br />

Since the discovery of human immunodeficiency virus (HIV) as the causative agent of acquired<br />

immunodeficiency syndrome (AIDS), perhaps the largest and most powerful consortium of scientists<br />

ever assembled to tackle a single disease has been brought to bear on the problem of AIDS and its<br />

treatment. From an unprecedented wealth of information regarding the molecular biology and virology<br />

of HIV collected in recent years, it became possible to identify numerous intervention points in the viral<br />

life cycle that could be exploited in the development of drugs for AIDS therapy (for reviews see<br />

Reference 1, 2, and 3). Among these, the virally-encoded enzymes, in particular reverse transcriptase<br />

and protease, have emerged as the most popular targets. A separate chapter of this book is dedicated to<br />

the description of reverse transcriptase and its inhibitors [4]. For the purpose of introduction only, it<br />

should be noted that nucleoside inhibitors of reverse transcriptase (AZT, ddI, ddC, d4T, and 3TC) have<br />

been widely used in clinical practice since 1987. Since then it has become apparent that this class of<br />

agents, while slowing progression of disease in HIV-infected patients, is limited in both activity and the<br />

duration of the clinical responses produced. Therefore in the search for better anti-HIV agents, the focus<br />

of effort was expanded to include the search for clinically useful inhibitors of a second viral enzyme,<br />

namely the protease. In contrast to reverse transcriptase, for which activity is required prior to the<br />

integration of viral genetic information into the host cell chromosomes, the viral protease plays a key<br />

role late in the virus life cycle and inhibitors of this enzyme display equal anti-viral activity in chronic<br />

and acute infection models in vitro [5].<br />

The HIV protease (HIV PR) is encoded <strong>by</strong> the 5' portion of the retroviral pol gene, which encodes all<br />

replicative enzymes. Viral structural proteins (p24,<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_1.html [2/29/2004 2:14:53 AM]


Document<br />

p17, p9, and p7) and replicative enzymes (protease, reverse transcriptase/ RnaseH, and integrase) are<br />

translated as either polyprotein P55-GAG, or a larger frameshift product P160-GAG-POL. In the<br />

process of virus assembly these polyproteins are proteolytically cleaved <strong>by</strong> the protease and this<br />

processing step, both in its timing and accuracy, is essential for the formation of infectious particles of<br />

HIV [6]. It was also shown early on that the inactivation of HIV PR, either <strong>by</strong> chemical inhibition or<br />

certain mutations, leads to the production of immature, noninfectious viral particles [7,8].<br />

Page 2<br />

Structurally HIV PR is a 99-amino-acid protein translated initially as a central part of the P160-GAG-<br />

POL polyprotein precursor. The autocatalytic processing from the 160 kDa precursor is poorly<br />

understood, but most likely occurs during the process of budding of pre-formed viral particles from the<br />

host cell [9]. After release from the precursor polyprotein, HIV PR forms a homodimer and acts in trans<br />

to correctly process GAG and GAG-POL polyproteins—a process required for formation of the viral<br />

capsid and nucleoprotein core.<br />

Retroviral proteases such as HIV PR are the latest additions to the wellstudied family of aspartic<br />

proteases. This family of enzymes, which includes, among others, proteases such as pepsin, renin, and<br />

cathepsins D and E, has been intensely studied in the past, and the knowledge gained from studies of<br />

these enzymes allowed early inferences as to the structure and function of the dimeric HIV PR.<br />

Moreover, the intensive effort over the past two decades to make inhibitors of human renin provided<br />

impetus for the early design of inhibitors of HIV PR. In fact, some of the renin inhibitors have turned<br />

out to be effective inhibitors of retroviral aspartic proteases as well and have served as the starting point<br />

for drug design. As a result of this many early inhibitors of HIV PR were peptidyl in nature and the best<br />

known example of such compounds is Ro31-8959, better known as saquinavir, a hydroxyethylaminecontaining<br />

mimetic of a hexapeptide substrate [10]. This potent inhibitor of HIV PR was discovered<br />

using a substrate-<strong>based</strong> rational approach to drug design and displays extremely high in vitro activity<br />

against clinical isolates and laboratory strains of HIV. Saquinavir has been recently approved <strong>by</strong> the<br />

FDA for the treatment of AIDS in combination with nucleoside inhibitors of reverse transcriptase, and<br />

the discovery of this compound was the first breakthrough and the starting point for many other<br />

innovative designs.<br />

Determination of the crystal structures of HIV PR gave new impetus to the design of novel inhibitors.<br />

One measure of the intensity with which new inhibitors were designed or discovered is the total number<br />

of crystal structures of inhibitory complexes, currently exceeding 250, that have been determined over<br />

the past 5 years. Very detailed crystallographic analysis combined with extensive biochemical<br />

characterization and site-specific mutagenesis studies made HIV PR perhaps the best characterized<br />

enzyme to date.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_2.html [2/29/2004 2:14:57 AM]


Document<br />

Page 3<br />

Based on the avalanche of papers describing the structure-<strong>based</strong> design of various HIV PR inhibitors, it<br />

would be reasonable to assume that, with the exception of saquinavir, all other HIV PR inhibitors that<br />

entered the stage of preclinical or clinical development were discovered using the elements of a structure<strong>based</strong><br />

approach. From the long list of more than 30 inhibitors considered as clinical candidates [11],<br />

currently there are three compounds (saquinavir, ritonavir, and indinavir) already approved <strong>by</strong> the FDA<br />

as anti-HIV drugs. Many factors that are requisite for in vivo activity in AIDS patients can only be<br />

predicted a priori in a very general sense. For instance, erratic oral bioavailability in humans, first-pass<br />

metabolism, binding to plasma proteins or tissue distribution may disqualify a perfect in vitro inhibitor<br />

of HIV replication and such properties can be very poorly predicted <strong>by</strong> any process of drug design. A<br />

potential answer to these problems is the parallel design of several chemically distinct compounds that<br />

may have similar in vitro activity but significantly different in vivo properties. The application of protein<br />

structure-<strong>based</strong> design offers such possibilities and in this text the discovery and optimization of<br />

different series of potent inhibitors of HIV PR will be discussed. In order to familiarize the reader with<br />

the architecture of HIV PR and the properties of its active site, the first paragraphs are devoted to the<br />

detailed description of the x-ray structures of the enzyme followed <strong>by</strong> several examples of inhibitors in a<br />

bound conformation.<br />

A. Three-Dimensional <strong>Structure</strong> of HIV PR<br />

Retroviral proteases such as HIV PR were tentatively assigned to the aspartic protease family on the<br />

basis of putative active-site sequence homology [12]. Mammalian aspartic proteases are bilobal, singlechain<br />

enzymes in which each lobe (or domain) contributes an aspartic acid residue to the active site [13].<br />

The active site itself is formed at the interface on the N- and C-terminal domains and exhibits<br />

approximate two-fold symmetry. Since the retroviral proteases are only about one-third the size of the<br />

two-domain eukaryotic enzymes, they were hypothesized to function as dimers in which each monomer<br />

contributes a single aspartic acid to the active site [14]. Obligate homodimeric proteases, in addition to<br />

providing a regulatory mechanism to control activation of the enzyme, represent the most efficient use<br />

of genetic information which, in retroviruses, is naturally parsimonious.<br />

The crystal structures of HIV PR confirmed the predicted dimeric character of the enzyme [15,16]<br />

(Figure 1). In all published crystallographic investigations of the unliganded form of the enzyme, the<br />

monomers are related to each other <strong>by</strong> crystallographic two-fold symmetry and are necessarily identical.<br />

The general topology of the HIV PR monomer is similar to that of a single-domain<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_3.html [2/29/2004 2:15:00 AM]


Document<br />

Figure 1<br />

Stereo view of the α-carbon backbone of HIV PR dimer. (a) The apoenzyme<br />

with flaps in the “open” conformation. (b) Inhibited form of HIV PR with flaps in a<br />

“closed” conformation. For clarity, the inhibitor is removed from the active site.<br />

pepsin-like aspartic protease and consists of antiparallel β-strands and a short, two-turn α-helix<br />

connected <strong>by</strong> loops of varying length. The dimer interface is formed <strong>by</strong> an antiparallel β-sheet<br />

comprising two strands from each monomer. The hydrophobic residues from those β-strands and two<br />

symmetry-related α-helices form the core of the dimer. The dimer is further stabilized <strong>by</strong> a net of<br />

hydrogen bonds involving the residues around the catalytic aspartic acids. The active site is formed <strong>by</strong><br />

the dimer interface and is composed of equivalent contributions of residues from each monomer. The<br />

substrate-binding cleft is bound on one side <strong>by</strong> the active site aspartic acid (Asp25/25') and on the other<br />

side <strong>by</strong> a pair of two-fold related, antiparallel β-hairpin structures, commonly referred to as “flaps.”<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_4.html [2/29/2004 2:15:23 AM]<br />

Page 4


Document<br />

The conserved active-site residues (Asp25, Thr26, and Gly27 from both monomers) form a symmetrical<br />

and highly hydrogen-bonded arrangement virtually identical to that described for pepsin [17]. The two<br />

aspartates are nearly coplanar with the “inner” carboxylate oxygens hydrogen bonded to the amide<br />

hydrogens of Gly27/27'. This designation (e.g. Gly 27/27') will be used throughout this text to indicate<br />

equivalent residues of the dimer. The two threonines are inaccessible to solvent and are hydrogenbonded<br />

to the main-chain amide groups of the other monomer, forming a rigid network called a<br />

“fireman's grip” [17]. As in the case of the structures of eukaryotic pepsins, there is electron density for<br />

a water molecule bound between the two carboxylates of the active-site aspartates.<br />

Page 5<br />

In the structure of the apo-form of HIV PR, the flaps from both monomers are related <strong>by</strong><br />

crystallographic two-fold symmetry and can be considered as being in an open conformation. In the<br />

structures of related proteases from Rous Sarcoma Virus and HIV-2, the flaps are either<br />

crystallographically disordered or in a partly closed conformation [18]. This suggests that, in solution, in<br />

the absence of ligands, the flaps are rather flexible and that the stable conformation of the flaps observed<br />

in the crystal structure of the apo-enzyme of HIV PR could be considered to result from kinetic trapping<br />

during the crystallization process.<br />

In the apo-form of HIV PR, the active site residues are located at the bottom of a rather shallow groove.<br />

Upon binding an inhibitor, the protease undergoes significant structural changes, particularly apparent in<br />

the flap region. As a result, a tunnel-like site is formed, which runs diagonally across the dimer<br />

interface. The tunnel has a volume of approximately 1140 Å 3 and is 23 Å long. Because of the dimeric<br />

nature of HIV PR, the active site has approximate two-fold symmetry with the dyad axis intersecting the<br />

plane of the catalytic aspartates. Along the active site tunnel, starting from the central aspartates, there<br />

are distinct subsites S1, S2, S3, and S4, and corresponding symmetry related subsites S1', S2' S3', and<br />

S4' (Figure 2). It should be noted that in this chapter, the convention of Schechter and Burger [19] will<br />

be used to describe enzyme specificity subsites (S1, S1', etc.) and the corresponding side chains of<br />

inhibitors (P1, P1', etc.). The boundaries of the subsites are formed <strong>by</strong> residues from both monomers of<br />

HIV PR. All subsites, with the exception of S4/S4', which are exposed to solvent, are bounded <strong>by</strong> mostly<br />

aliphatic side chains and have hydrophobic character. The borders of the S1/S1' subsites are formed <strong>by</strong><br />

the side chains of Ile23/23', Ile50/50', Ile84/84', Pro81/81', the γ carbon of Thr80/80', carboxylates of the<br />

active site Asp25/25', and the carbonyl oxygens of Gly27/27'. The S2/S2' subsites are bounded <strong>by</strong><br />

Val32/32', Ile50/50', Ile47/47', Leu76/76', Ala28/28', and the carboxylates of Asp30/30'. The S3/S3',<br />

subsites are partly exposed to solvent and are bordered <strong>by</strong> the side chains of Leu23/23', Val82/82',<br />

Pro81/81', and the guanidinium groups of Arg8/8', which form a salt bridge with the carboxylates of<br />

Asp29/29'. Most of<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_5.html [2/29/2004 2:15:26 AM]


Document<br />

Figure 2<br />

Schematic representation of the specificity subsites of the HIV PR active site with bound<br />

peptidic inhibitor JG-365. Amino acids forming the boundaries of the particular<br />

subsites are shown.<br />

the hydrogen bond donor and acceptor functional groups of the active site are located in an approximate<br />

plane that lies along the long axis of the tunnel and is somewhat perpendicular to the plane of the<br />

subsites. The hydrogen-bonding functionalities include the carboxylates of the catalytic aspartates, the<br />

carbonyl oxygens of Gly27 and Gly48, the amide nitrogens of Asp29' and Gly48, the carboxylate of<br />

Asp29', and the dimer symmetry-related groups on the other side of the active site. Additional groups<br />

capable of forming hydrogen bonds with ligands are located in the outer part of the S2/S2' subsite and<br />

include the amide nitrogens and the carboxylates of Asp30/30'. There are five conserved water<br />

molecules in the active site of HIV PR. Four of the waters are symmetrically distributed in the S3/S3'<br />

subsites and one, hereafter called Wat301, is located near the two-fold axis of the dimer and, in the<br />

presence of most inhibitors, is approximately tetrahedrally coordinated <strong>by</strong> the hydrogen bonds formed<br />

between carbonyl oxygens of the ligand(s) and the amide nitrogens of Ile50/50' of the flaps. In the<br />

ligand-bound form of HIV PR, Wat301 is completely inaccessible to solvent, and it has been speculated<br />

that its functional substitution could be energetically favorable [18] or at least may lead to discovery of<br />

novel nonpeptidic inhibitors [20]. Thus, there are 18 hydrogen bond donors or acceptors in<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_6.html [2/29/2004 2:15:34 AM]<br />

Page 6


Document<br />

Page 7<br />

the active site of HIV PR-16 that could form hydrogen bonds directly and two in which the interaction is<br />

mediated <strong>by</strong> the conserved Wat301. The total solvent accessible surface area of the eight subsites of the<br />

HIV PR active site is approximately 1150 Å 2. Because of the large number of groups with hydrogen-bondforming<br />

potential, 450 Å 2 of the surface has a polar character, and the nonpolar area of the subsites is<br />

slightly larger, approximately 700 Å 2.<br />

B. Structural Flexibility of HIV PR<br />

In the process of viral assembly, HIV PR specifically cleaves nine cleavage sites on GAG and GAG-POL<br />

polypeptides [21]. Examination of the amino acid composition of the recognized substrate sites (Table 1)<br />

indicates their hydrophobic character and significant sequence variability. The loose specificity of HIV PR<br />

most likely reflects its functions in a world of reduced complexity within the confines of the budding<br />

virion. The length of the viral protein precursors (approximately 1500 amino acids) reduces the number of<br />

potential sequences the protease must discriminate from in selecting its nine cleavage sites. Therefore,<br />

HIV PR and other retroviral proteases are not enzymes that have evolved to carry out a single reaction at a<br />

rapid rate, but rather enzymes with minimum specificity required to cleave the viral precursors in a<br />

specific and orderly manner.<br />

The loose specificity requirements demonstrated <strong>by</strong> effective binding and catalytic processing of all nine<br />

sequences, albeit at different rates [22], was the<br />

Table 1 The Sequences of the Proteolytic Processing Sites of HIV-1<br />

HIV-1 PR<br />

Cleavage sites<br />

Scissile bond<br />

P17/P24 V S Q N Y P I V Q N<br />

P24/P2 K A R V L A E A M S<br />

P2/P7 S A T I M M Q R G N<br />

P7/P1 E R Q A N F L G K I<br />

P1/P6 R P G N F L Q S R P<br />

TF/PR V S F S F P Q I T L<br />

PR/RT C T L N F P I S P I<br />

RT/RN G A E T F Y V D G A<br />

RN/IN I R K V L F L D G I<br />

Schechter-Berger<br />

notation<br />

P5 P4 P3 P2 P1 P1' P2' P3' P4' P5'<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_7.html (1 of 2) [2/29/2004 2:15:46 AM]


Document<br />

TF—transframe, PR—protease, RT—reverse transcriptase, RH—RNAse H, IN—integrase. The location of the<br />

processing sites in HIV-1 were determined <strong>by</strong> protein sequencing of HIV-1 virion proteins.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_7.html (2 of 2) [2/29/2004 2:15:46 AM]


Document<br />

Page 8<br />

first indication that the recognition subsites of the HIV PR can display flexibility upon binding of<br />

substrates or inhibitors. Early crystal structures of the HIV PR apo-enzyme and complexes with peptidic<br />

inhibitors showed several conformations of the active site forming flaps, which include the residues<br />

Met46/46' to Ile54/54' [15,16]. Increased availability of coordinates of HIV PR complexed with various<br />

inhibitors and crystallized in different crystallographic space groups allowed for more rigorous<br />

examination of domain movements and structural changes in the active site.<br />

The alignment of several crystal structures of HIV PR in a common frame of reference, which most<br />

commonly includes the region around the symmetryrelated active site triad Asp25/25'-Thr26/26'-<br />

Gly27/27', will highlight those regions of the backbone where significant displacement occurs upon<br />

accommodating the individual inhibitors. Examination of the aligned structures, which included<br />

examples of all classes of inhibitors, indicated only small variation of the backbone and limited<br />

movements in the two binding loops, comprising residues Leu76-Ile84 from both monomers. The loops<br />

form the outer walls of subsites S1/S1' and S3/S3' with inward-facing hydrophobic side chains of<br />

isoleucines and valines. The flexibility of these loops, which in some cases can move outward <strong>by</strong> as<br />

much as 2.5 Å, has a significant impact on the volume of the specificity subsites, which in turn can<br />

accommodate corresponding P1/P1' and P3/P3' moieties of various sizes. Interestingly, the predominant<br />

resistance-causing mutations are located on the same loops and involve changes in residues Val82/82'<br />

and Ile84/84' (see below). It should be noted, that while the alignment of several crystal structures<br />

provides information about the flexible regions, the extent of flexibility of the residues around the HIV<br />

PR active site can be limited <strong>by</strong> crystal packing forces and may represent a crystallographic artifact. In<br />

all characterized crystal forms of HIV PR [23] the loops 76–84 and 46–56 participate in crystal lattice<br />

formation and the particular conformation of these loops can be driven <strong>by</strong> crystallization conditions or<br />

interactions with other molecules related <strong>by</strong> the crystallographic symmetry.<br />

C. Inhibitors of HIV PR<br />

In general, inhibitors of HIV PR can be divided into three distinct groups. The first group includes<br />

peptidic inhibitors that utilize various transition-state dipeptide analogs such as statine, hydroxyethylene,<br />

and hydroxyethylamine incorporated into peptidic frameworks of differing lengths. Several crystal<br />

structures of this type of inhibitor complexed with HIV PR were solved and the structural information<br />

provided a wealth of information as to the minimum size of inhibitors, geometry of hydrogen bonds<br />

formed within the active site, and the structural flexibility of the subsites (for reviews see Reference 18<br />

and 23). The second and perhaps largest group of HIV PR inhibitors includes peptidomimetic<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_8.html [2/29/2004 2:15:50 AM]


Document<br />

compounds that utilize similar transition-state analogs and retain at least one peptide bond with a side<br />

chain corresponding to a naturally occurring amino acid. Several compounds from this group have<br />

excellent pharmacokinetic and antiviral properties and, in fact, all three HIV PR inhibitors approved for<br />

clinical use (saquinavir, ritonavir, and indinavir) belong to this class of compounds. The last and the<br />

smallest group of HIV PR inhibitors has a distinct nonpeptidic character. Compounds from this class<br />

were discovered either <strong>by</strong> screening libraries of existing compounds or <strong>by</strong> structure-<strong>based</strong> de novo<br />

design. Illustrative examples of inhibitors belonging to all three classes and a brief description of the<br />

discovery of selected compounds are presented below.<br />

D. Peptidic Inhibitors of HIV PR<br />

Page 9<br />

The concept of peptidic inhibitors of HIV PR can be exemplified <strong>by</strong> the crystal structure of the statinecontaining<br />

peptidic compound AG1002 (Figure 3) [23]. In AG1002, the statine moiety replaces the<br />

scissile dipeptide while the flanking amino acids were derived from the natural substrate cleaved <strong>by</strong> HIV<br />

PR. The inhibitor binds to the active site in an extended conformation with the central hydroxyl group of<br />

the statine moiety forming hydrogen bonds with the active-site aspartic acids 25/25'. The peptidic<br />

backbone and the side chains of the<br />

Figure 3<br />

Stereo view of the peptidic inhibitor AG1002 bound to the active site of HIV PR.<br />

The distribution of the specificity subsites S and S' is similar to that shown in<br />

Figure 2. The boundaries of the HIV PR active site are indicated <strong>by</strong> the dotted<br />

surface.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_9.html (1 of 2) [2/29/2004 2:16:10 AM]


Document<br />

Page 10<br />

inhibitor form 16 hydrogen bonds and occupy subsites from S4 to S1, S2', and S3'. The carbonyl<br />

oxygens of P2 and P1' accept two hydrogen bonds from the flap water Wat301, which in effect is nearly<br />

tetrahedrally coordinated. Due to the structural nature of statine, which lacks the P1' side chain, the S1'<br />

pocket remains unoccupied. The S1 subsite is only partially filled <strong>by</strong> the P1 side chain of leucine. The<br />

P2 and P2' side chains of asparagine and glutamine form hydrogen bonds with Asp30' and 30, while the<br />

aliphatic carbons of both side chains make several hydrophobic contacts in the S2 and S2' pockets<br />

respectively. Despite the large number of hydrogen bonds formed within the HIV PR active site,<br />

AG1002 has rather low inhibitory potency with a binding constant of 0.55 μM The low binding constant<br />

most likely reflects the absence of the P1' group, the free energy required for desolvation of the<br />

hydrophilic side chains, and the charged N- and C-termini as well as entropic effects caused <strong>by</strong> the<br />

flexible nature of the heptapeptide.<br />

Other interesting examples of peptidic inhibitors are compounds utilizing other transition-state analogs,<br />

e.g. reduced amide-containing hexapeptide MVT-101 [24], hydroxyethylene-containing octapeptide U-<br />

85548e [25], and hydroxyethylamine-containing heptapeptide JG-365 [26]. All these compounds bind to<br />

the active site of HIV PR in a similar extended conformation and the small differences in the geometry<br />

of hydrogen bonds formed with HIV PR can be attributed to the different character and length of the<br />

transition-state analogs. The chemical structures and inhibition constants of these inhibitors are<br />

summarized in Table 2. Note that the inhibition constants cited throughout this chapter and in Tables 2,<br />

3, and 6 were determined in different laboratories—often using significantly different assay<br />

conditions—and therefore might not be meaningfully comparable.<br />

Due to their substantial size and peptidic nature, inhibitors from this class were not suitable for clinical<br />

application. Nevertheless, the structural information derived from many crystal structures of peptidic<br />

inhibitors bound to the HIV PR active site was critical for subsequent modeling and design of the next<br />

generation of peptidomimetic and nonpeptidic inhibitors of HIV PR.<br />

E. Peptidomimetic Inhibitors of HIV PR<br />

<strong>Design</strong> and <strong>Structure</strong> of Ro-31-8959 (Saquinavir)<br />

The strategy of designing saquinavir was <strong>based</strong> on the transition-state mimetic concept, an approach that<br />

has been used successfully in the design of potent inhibitors of renin and other aspartic proteases [10].<br />

From the variety of nonscissile transition-state analogs of a dipeptide, the hydroxyethylamine mimetic<br />

was selected because it most readily accommodates the amino acid moiety characteristic of the Phe-Pro<br />

and Tyr-Pro cleavage sequence of the<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_10.html [2/29/2004 2:16:13 AM]


Document<br />

Page 11<br />

retroviral substrates. In the first step of design, the dipeptide analog consisting of<br />

Phe[CH(OH)CH 2N]Pro was used to determine the minimum sequence required for potent inhibition.<br />

From this study a compound was selected that included benzyloxycarbonyl at the N-terminal side of the<br />

inhibitor followed <strong>by</strong> the P2 asparagine, the hydroxyethylamine isostere with side chains of<br />

phenylalanine and proline in the P1 and P1' positions respectively and the NH-t-butyl group at the Cterminal<br />

part. In the following design, the side chain of proline was consequently modified to a<br />

piperidine and finally to a decahydroisoquino-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_11.html (1 of 2) [2/29/2004 2:16:45 AM]


Document<br />

Page 12<br />

line moiety, and the N-terminal benzyloxycarbonyl group was replaced <strong>by</strong> the quinoline-2-carbonyl. The<br />

resulting compound, Ro-31-8959, was one of the first peptidomimetic inhibitors with very high antiviral<br />

potency and became a benchmark for further design of HIV PR inhibitors [10].<br />

The high-resolution crystal structure of saquinavir bound to the active site of HIV PR was solved in<br />

many laboratories [23,27]. The incorporation of decahydroisoquinoline moiety, which can be considered<br />

as a conformationally restrained mimic of cyclohexylalanine, has some important consequences. First,<br />

the length of the C-terminal part of the inhibitor has been restricted to the P2' residue which, in<br />

saquinavir, consists of a NH-t-butyl group. Second, it restrained the conformational freedom of the<br />

otherwise peptidic backbone, minimizing the entropic penalty to the free energy of binding. In the<br />

crystal structure of saquinavir with HIV PR (Figure 4), the decahydroisoquinoline in the preferred chairchair<br />

conformation, makes extended hydrophobic contacts in the S1' subsite. The bond between the<br />

methylene carbon and the nitrogen of decahydroisoquinoline is in the low-energy equatorial<br />

conformation and the nitrogen, even if protonated, is not in a position to form a hydrogen bond with the<br />

active-site residues. The central hydroxyl group is in the R(syn) conformation and is within the<br />

hydrogen-bond-forming distance with both carboxylates<br />

12640-0012a.gif<br />

Figure 4<br />

Stereo view of the peptidomimetic inhibitor Ro 31-8959 (saquinavir) bound<br />

to the active site of HIV PR. The distribution of the specificity subsites S and<br />

S' is identical to that shown in Figure 2. Note the stacking interaction<br />

between the quinoline moiety and the P1 side chain of phenylalanine.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_12.html [2/29/2004 2:16:50 AM]


Document<br />

Page 13<br />

of the active-site aspartates. Similar to Ag1002 and other peptidic inhibitors, the carbonyl oxygens of the<br />

P2 and P1' amides are within hydrogen-bonding distance of the flap water molecule; however, the<br />

geometry of the second hydrogen bond is distorted due to the additional spacing between both carbonyl<br />

groups. The nitrogen of the t-butylamide is displaced from the normal P2' position <strong>by</strong> approximately 1.8<br />

Å and, as a result, cannot form a direct hydrogen bond with the carbonyl oxygen of Gly27. Instead the tbutylamide<br />

nitrogen interacts via highly ordered water molecules with the amide nitrogen of Asp29 and<br />

the carbonyl oxygen of Gly27. The aliphatic t-butyl moiety occupies the S2' subsite and the position of<br />

the backbone in this region prohibits any further extension into the S3' pocket. The P1 and P2 side<br />

chains of phenylalanine and asparagine, respectively, occupy the corresponding subsites and have a<br />

similar conformation to the equivalent groups observed in peptidic inhibitors. In the crystal structure, the<br />

N-terminal quinoline-2-carboxylate is moved to the side and, as a result, the carbonyl oxygen forms<br />

hydrogen bonds with the ordered water molecule and with the amide nitrogen of Asp29'. The quinoline<br />

ring is in a low-energy conformation with respect to the preceding carbonyl oxygen, which places the<br />

aromatic nitrogen in unfavorable close contact (3.3 Å) to the carbonyl oxygen of the flap Gly48.<br />

Because of the absence of any further contacts with the HIV PR active site residues, the contribution of<br />

the quinoline moiety to the free energy of binding remains unclear. Perhaps in solution, a stacking<br />

interaction of the P1 phenyl ring and the aromatic quinoline restricts the conformational freedom of Ro-<br />

31-8959, in effect diminishing the free-energy loss due to the entropic and desolvation effects.<br />

Saquinavir, despite its distinct peptidomimetic character is a very potent inhibitor of HIV PR with an<br />

inhibition constant of 0.9 nM and an antiviral IC50 in vitro of 0.020 μM [10]. Although it suffers from a<br />

low oral bioavailability (5–10% in humans), it became an important starting point for the design of<br />

second generation, less-or nonpeptidic inhibitors. Saquinavir became the first HIV PR inhibitor<br />

approved <strong>by</strong> the FDA for treatment of AIDS.<br />

<strong>Design</strong> and <strong>Structure</strong> of ABT-538 (Ritonavir)<br />

An interesting concept for designing specific HIV PR peptidomimetic inhibitors with internal two-fold<br />

symmetry was first formulated <strong>by</strong> John Erickson and his colleagues from Abbott Laboratories [28].<br />

They reasoned that if HIV PR incorporates symmetry into its active site structure, compounds that<br />

mimic this symmetry might be novel, more specific, and potent inhibitors and, furthermore, due to the<br />

bidirectionality of peptide bonds, might be sufficiently less peptidic in character and pharmacologically<br />

superior to the classical peptide-<strong>based</strong> compounds. The crystal structure of one of the first compounds<br />

from this series (A74704) verified the assumption of symmetrical binding conformation in the<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_13.html [4/5/2004 4:44:15 PM]


Document<br />

active site of HIV PR. The inhibitor consists of the central diamino alcohol moiety with symmetrically<br />

distributed phenylalanine side chains and two flanking, Cbz-blocked, valine residues. Except for the<br />

asymmetric hydroxyl group, A74704 binds to the active site in a symmetric mode and the inconsistent<br />

distribution of the terminal Cbz groups is most likely caused <strong>by</strong> crystal lattice contacts and may not<br />

reflect the binding mode in solution [28].<br />

Page 14<br />

The design of symmetrical inhibitors was further extended to include a series of diamino, diol core units,<br />

in which the C2 axis bisects the bond connecting the two hydroxy-bearing carbon atoms [29]. Such<br />

inhibitors consistently showed greater potency than A74704, but the relative potencies of the diols<br />

differed for different diastereomers, and they did not exhibit a uniform dependence on the<br />

stereochemistry at the hydroxymethyl position. Surprisingly, high-resolution crystal structures of HIV<br />

PR with all possible diol diastereomers, (S,S, R,R and R,S) revealed that most of the inhibitors bind in a<br />

clearly asymmetric fashion placing only one of the diol hydroxyl groups on the C2 axis dissecting the<br />

active site of HIV PR and the catalytic carboxylates of Asp25/25'. The asymmetric placement of diols<br />

causes translation of inhibitors along the long axis of the active site and, as a result, the midpoint of the<br />

compounds is displaced <strong>by</strong> up to 0.9 Å from the two-fold axis of the HIV PR. Nevertheless, the dihedral<br />

angles of the symmetry-related bonds are in most cases within 10°, and the inhibitors maintain overall<br />

symmetry in the bound conformation [23,29].<br />

The ABT-538 design was a direct consequence of the pioneering work with peptidomimetic compounds<br />

with the internal C2 symmetry [30]. Since the high-resolution crystal structures of a family of diolcontaining<br />

compounds indicated that in most cases only one of the diol hydroxyls interacts with the<br />

catalytic aspartic acids 25/25', in subsequent designs the noninteracting hydroxyl group was replaced <strong>by</strong><br />

a hydrogen. This substitution reduced the free energy penalty required for desolvation of the hydroxyl<br />

group and increased the inhibitory potency without perturbing the binding mode of the compounds [30].<br />

In the further search for related inhibitors with improved oral bioavailability, the focus of effort<br />

concentrated on the effect that molecular size, aqueous solubility, and hydrogen-bonding capability has<br />

on pharmacokinetic behavior. This study resulted in a smaller compound, A-80987, in which the P2'<br />

side chain of valine was eliminated and the terminal 2-pyridinyl group was replaced <strong>by</strong> 3-pyridinyl<br />

moiety that makes van der Waals contacts in the S2' subsite and forms a hydrogen bond with the amide<br />

nitrogen of Asp30 [31]. The pharmacokinetic properties of A-80987 were significantly improved over<br />

larger, symmetrical compounds from this series and, at the same time, the high antiviral activity typical<br />

for these inhibitors was largely unaffected. In subsequent optimization, which focused on the metabolic<br />

stability of these inhibitors in vivo, the electron-rich and oxidation-prone pyridinyl groups were replaced<br />

<strong>by</strong> thiazoles.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_14.html [4/5/2004 4:44:17 PM]


Document<br />

Page 15<br />

Thiazoles are less electron-rich isosteres of pyridines and therefore it was speculated that compounds<br />

with such substitution may have improved metabolic stability [30]. The modeling of A-82200 in which<br />

the N-terminal pyridinyl group was substituted <strong>by</strong> a 4-thiazolyl moiety indicated that the 5-membered<br />

ring binds in the S3 subsite and can be further derivatized at the 2 position <strong>by</strong> an isopropyl group. The<br />

isopropyl functionality makes van der Waals contacts with Val82 and fills the hydrophobic part of the<br />

S3 subsite in nearly optimal fashion.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_15.html (1 of 2) [4/5/2004 4:44:40 PM]


Document<br />

Page 16<br />

The resulting compound, ABT-538 (Table 3), binds to the active site of HIV PR in an extended<br />

conformation. The central, asymmetric hydroxyl group is within hydrogen-bonding distance of the<br />

catalytic aspartates 25/25', and the P1/P1' phenylalanine side chains are symmetrically distributed in the<br />

corresponding subsites. The nitrogens of the symmetric amide bonds on both sides of the central<br />

aminoalcohol are barely within the hydrogen-bonding distance of the carbonyl oxygens of Gly27/27'<br />

(3.4 Å) and the carbonyl oxygens of these amide bonds participate in the tetrahydral coordination of the<br />

flap water molecule Wat301. On the N-terminal side of the compound, the P2 side chain of valine fills<br />

the S2 subsite and the terminal 2-isopropyl-4-thiazolyl makes hydrohobic contacts with the residues in<br />

the S3 pocket and has a stacking interaction with the P1 phenylalanine. On the C-terminal side, the 5thiazole<br />

is positioned to interact within the S2' subsite, and the nitrogen on the 5-membered ring is<br />

within hydrogen-bonding distance of the amide of Asp30'.<br />

Despite two peptide bonds present in ABT-538, this compound has substantial oral availability in<br />

humans and a very high antiviral activity in vivo [30]. Recently, ABT-538, better known as ritonavir,<br />

has been approved <strong>by</strong> the FDA for treatment of AIDS in combination with inhibitors of the reverse<br />

transcriptase.<br />

<strong>Design</strong> and <strong>Structure</strong> of L-735,524 (Indinavir)<br />

Indinavir is another example of very potent peptidomimetic compound discovered using the elements<br />

the crystal structure-<strong>based</strong> design [32] and SAR (structure activity relationship). The starting point for<br />

the design was a series of compounds containing the hydroxyethylene isostere of a scissile dipeptide<br />

[33]. An example of compounds from these series is L-685,434, which consists of a tert-butylcarbamate<br />

group forming the P2 moiety, symmetrically distributed phenylalanine side chains in the P1/P1', and the<br />

indanol group in the P2' portion of the inhibitor. Although very potent, the optimized molecules from<br />

this series lacked aqueous solubility and an acceptable pharmacokinetic profile [32]. The Merck group<br />

hypothesized that incorporation of a basic amine-containing functionality, such as the<br />

decahydroisoquinoline group of saquinavir, into the backbone of L-685,434 series might improve the<br />

solubility and bioavailability of this type of compound. Also the replacement of the P2/P1<br />

functionalities, the tert-butylamide and phenylalanine side chain <strong>by</strong> the decahydroisoquinoline tertbutylamide,<br />

would generate a novel class of hydroxylaminepentanamide isostere with potentially<br />

improved metabolic stability in vivo. An additional strong argument for using decahydroisoquinoline as<br />

an isostere of P1/P2 moieties was the restricted conformational freedom of the enclosed-into-a-ring<br />

basic amine, which should decrease the entropy change upon binding to HIV PR in a similar fashion to<br />

that observed in saquinavir. In<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_16.html [4/5/2004 4:44:41 PM]


Document<br />

Page 17<br />

the resulting chimeric inhibitor the central hydroxyl group forms hydrogen bonds with the catalytic<br />

aspartic acids 25/25' and the hydrophobic side chains of the P1/P1' decahydroisoquinoline and<br />

phenylalanine respectively are separated from the central hydroxyl-bearing carbon <strong>by</strong> the methylene<br />

linkers forming a pseudosymmetrical arrangement. In the subsequent optimization of inhibitors from<br />

this novel series, a smaller piperazine group was substituted for the decahydroisoquinoline group, which<br />

offered a possibility to expand from the N4 position to the partially lipophilic S3 subsite. One of the first<br />

compounds from the piperazine series possessed a benzyloxycarbonyl moiety attached to the piperazine<br />

ring and the additional hydrophobic interaction in the S3 subsite was reflected <strong>by</strong> substantial increase in<br />

both intrinsic potency and in the ability to inhibit viral spread in infected cells in vitro. Finally, the<br />

replacement of the benzyloxycarbonyl group <strong>by</strong> the 3-pyridylmethyl moiety (Table 3) provided both<br />

lipohilicity for binding to the HIV PR active site and a weakly basic nitrogen that increased aqueous<br />

solubility and oral bioavailability. The crystal structure of L-735,524 (indinavir) bound to the active site<br />

of HIV PR [34] indicates that the 3-pyridylmethyl group attached to the N4 position of the piperazine<br />

ring makes hydrophobic contacts with the residues in the S3 and S1 pockets and the tert-butyl moiety<br />

fills the S2 subsite in the fashion previously observed in the structure of saquinavir. The positions of the<br />

P2 and P1' carbonyls maintain the proper alignment to form hydrogen bonds with the flap water<br />

Wat301. The terminal indanol group of indinavir occupies the S2' subsite with the hydroxyl group<br />

within hydrogen-bonding distance of the amide nitrogen of Asp29.<br />

The high aqueous solubility and largely nonpeptidic character of indinavir may be responsible for the<br />

good oral bioavailability, respectable pharmacokinetic profile, and high antiviral activity observed with<br />

this compound. Similar to saquinavir and ritonavir, indinavir has been recently approved <strong>by</strong> the FDA for<br />

treatment of AIDS.<br />

F. Nonpeptidic Inhibitors of HIV PR<br />

The nonpeptidic inhibitors of HIV PR can be divided into two subclasses. Compounds that belong to the<br />

first group maintain the general binding mode of the peptidomimetic inhibitors including formation of<br />

the key hydrogen bonds with the active site residues. An example of such nonpeptidic inhibitors of HIV<br />

PR is AG1343 (nelfinavir). The second group of nonpeptidic HIV PR inhibitors includes compounds<br />

with a binding mode significantly different from that described for the peptidomimetic compounds.<br />

Most inhibitors in this latter class were initially discovered <strong>by</strong> screening natural-products libraries or <strong>by</strong><br />

structure<strong>based</strong> de novo design. The most interesting examples of the nonpeptidic inhibitors from this<br />

group are the independently discovered but structurally<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_17.html [4/5/2004 4:44:43 PM]


Document<br />

Page 18<br />

related 4-hydroxypyrans and 4-hydroxycoumarins, the cyclic urea-<strong>based</strong> DMP323 series, and AG1284.<br />

<strong>Design</strong> and <strong>Structure</strong> of AG1343 (Nelfinavir)<br />

Analysis of the crystal structure of saquinavir with HIV PR indicated that while the nonpeptidic<br />

components of the ligand, namely the decahydroisoquinoline and the t-butylamide moieties fill the S1'<br />

and S2' subsites nearly optimally, the N-terminal portion offered the possibility for remodeling, aimed at<br />

the elimination of the peptidic character. Also, the contribution of the quinoline to the binding affinity to<br />

HIV PR was difficult to rationalize. Since the removal of quinoline resulted in a nearly 1000-fold loss in<br />

binding constant, it was concluded that the stacking interactions of the P1 phenyl ring and the P3<br />

aromatic moiety of quinoline are necessary for the conformational stability of Ro 31-8959. In an attempt<br />

to redesign the N-terminal part of the ligand, the nonpeptidic portions of the P1' and P2' were maintained<br />

but for reasons of synthetic accessibility, the decahydroisoquinoline moiety was replaced <strong>by</strong> an orthosubstituted<br />

benzylamide [35]. Crystallographic analysis of both compounds showed that saquinovir and<br />

the modified LY289612 bind essentially identically to the active site of HIV PR and their inhibition<br />

constants and antiviral activity were very similar (Table 4 and Table 3).<br />

In the first attempt to functionally substitute the P2 side chain of asparagine, the isophthalic-acidcontaining<br />

compound was modeled and the low-energy conformation of the aromatic ring, required for<br />

binding in the S2 subsite, was stabilized <strong>by</strong> a tertiary carboxamide in the P3 region of the inhibitor [36].<br />

The analysis of the binding mode and interactions of the isophthalic ring in the S2 subsite indicated a<br />

lipophilic pocket deep on the border between the S2 and S1' subsites, which could be conveniently filled<br />

with a methyl group extending from the 2 position of the ring. The resulting compound II in Table 4 lost<br />

most of the peptidomimetic character of LY289612 but retained its inhibitory potency.<br />

In an independent line of design, the relationship between the P1 phenylalanine side chain and the P3<br />

quinoline was investigated. In the crystal structure of saquinavir bound to the active site of HIV PR<br />

(Figure 4), the aromatic ring of the P1 phenylalanine makes several van der Waals contacts with<br />

residues forming the S1 subsite. Computer modeling indicated that an extension of the phenylalanine<br />

side chain to phenethyl (homophenylalanine) would lead to prohibitive close contacts of the phenyl ring<br />

with the aliphatic side chains of HIV PR. On the other hand, replacement of the γ-carbon of the<br />

homophenylalanine <strong>by</strong> sulphur, which has a more acute C-S-C bond angle, would direct the aromatic<br />

ring into the neighbouring S3 subsite without changing the desired lipophilic nature of the P1 side chain.<br />

The increased area of hydrophobic interactions in the S1 and S3 subsites <strong>by</strong> compounds with the Sphenylcysteine<br />

and<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_18.html [4/5/2004 4:44:45 PM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_19.html (1 of 2) [4/5/2004 4:44:51 PM]<br />

Page 19


Document<br />

Page 20<br />

S-naphthylcysteine derived side chains in P1 resulted in a substantial increase in the inhibition constants<br />

[37]. The increase in the binding affinity to the low picomolar range in enzyme inhibition assay, allowed<br />

for subsequent truncation of the P3 quinoline moiety. The final compound from this miniseries<br />

(compound III in Table 4) consisted of the ortho-substituted benzamide in the P1' and P2', Snaphthylcysteine<br />

in P1 and asparagine in P2. Despite reduced molecular weight, the inhibition constant<br />

of this compound for HIV PR was comparable to LY289612.<br />

The observation that a larger, nonpeptidic moiety in the P1 could eliminate the need for the P3 side<br />

chain led to hybrid molecules that incorporated ring structures as the P2 component and maintained the<br />

P1 S-naphthylcysteine side chain of compound III. In this miniseries several bicyclic functionalities<br />

were modeled as the P2 substituents and one example, compound IV utilizing a tetrahydroquinoline<br />

group, is shown in Table 4 [38]. In subsequent modeling, it was noticed that the P2 bicyclic functionality<br />

might be replaced <strong>by</strong> 2,3-disubstituted phenyl rings. In particular, a methyl substitution in position 2<br />

would increase the area of hydrophobic interaction in a manner previously observed in the isophthal<br />

series. Addition of a hydrophilic functionality attached at position 3 could increase the solubility of the<br />

compound and contribute to the binding constant <strong>by</strong> forming a hydrogen bond with the carboxylate<br />

oxygen of Asp30. A compound with a 2-methyl-3-hydroxy substitution pattern was synthesized and<br />

showed an improved inhibition constant of 3 nM in the HIV PR enzyme assay (Table 4). The crystal<br />

structure of compound V with HIV PR was solved and indicated the predicted binding mode with the<br />

possibility of a stacking interaction between the P2 phenyl and the P1 thio-naphthyl groups and the<br />

expected hydrophobic and hydrogen-bonding interactions of the P2 moiety with the protein side chains<br />

in the corresponding specificity pocket [38].<br />

As with the optimized compounds from other series, compound V suffered from low aqueous solubility.<br />

The replacement of the P1' aryl group <strong>by</strong> the basic amine-containing decahydroisoquinoline<br />

dramatically increased the solubility and allowed for truncation of the P1 S-naphthylcysteine side chain<br />

to S-phenylcysteine without any loss of inhibitory activity. The resulting compound VI, AG1343 or<br />

nelfinavir, has an inhibition constant of 1.9 nM in the HIV PR enzyme assay and respectable antiviral<br />

activity with an IC 90 of 60 nM [39]. The nonpeptidic character, pH-dependent solubility profile, and the<br />

small molecular weight of nelfinavir may contribute to its good pharmacokinetic profile in humans<br />

[40,41]. Currently, this compound is being tested and is in the advanced phase of clinical trials.<br />

The crystal structure of nelfinavir bound to the active site of HIV PR is shown in Figure 5. The general<br />

binding mode of this compound, in particular the path of the backbone, is similar to the binding mode of<br />

peptidyl inhibitors. Nevertheless, the lack of any peptide bonds utilizing naturally occurring amino<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_20.html [4/5/2004 4:44:58 PM]


Document<br />

Figure 5<br />

Stereo view AG1343 (nelfinavir) bound to the active site of HIV PR.<br />

Page 21<br />

acids qualifies nelfinavir to be a member of the group of nonpeptidic inhibitors of HIV PR. The unique,<br />

and perhaps crucial hydrogen-bonding interaction of the P2 hydroxyl group with the carboxylate oxygen<br />

of Asp30, combined with the smaller area of hydrophobic contacts in the S1 and S3 specificity subsites<br />

are the principal differences from other clinically active HIV PR inhibitors and may contribute to a<br />

distinct resistance pattern and point to additional utility of nelfinavir in the treatment of AIDS.<br />

<strong>Design</strong> and <strong>Structure</strong> of DMP323<br />

A cyclic urea-containing HIV PR inhibitor, DMP323, was discovered using de novo structure-<strong>based</strong><br />

design principles. Similar to the concept of Erickson and his co-workers from Abbott Laboratories, the<br />

group from DuPont-Merck attempted to take advantage of the two-fold symmetry of HIV PR in<br />

designing compounds that maintained the interaction of the diol with the catalytic aspartic acids 25/25'<br />

and at the same time were able to functionally displace the ubiquitous flap water molecule Wat301.<br />

They hypothesized that incorporation of the binding features of this structural water molecule into an<br />

inhibitor would be beneficial because of the entropic gain due to its displacement and because the<br />

conversion of a flexible linear inhibitor into a rigid, cyclic structure with restricted conformation should<br />

provide an additional, positive entropic effect. In the initial design, a cyclohexanone with the ketone<br />

oxygen as the structural<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_21.html [4/5/2004 4:45:10 PM]


Document<br />

Page 22<br />

water mimic was used and in subsequent synthetic targets the cyclohexanone ring was enlarged to a 7membered<br />

ring to incorporate a diol functionality. This target was further modified to a cyclic urea,<br />

which can be symmetrically substituted from both nitrogens without creating unnecessary stereocenters.<br />

The crystal structures of about 10 cyclic-urea-<strong>based</strong> inhibitors with HIV PR were solved [42]. In all<br />

cases, the C2 symmetric inhibitors bind to the HIV PR active site with the diad symmetry axes of the<br />

protease and the compounds being nearly coincident. The 7-membered ring of the inhibitors is roughly<br />

perpendicular to the plane of the catalytic aspartates 25/25' and both hydroxyl groups of the diol are<br />

positioned to interact with their carboxylates. The carbonyl oxygen of the inhibitors accepts hydrogen<br />

bonds from backbone amides of symmetrically distributed residues Ile50/50' of the flap. In the structure<br />

of DMP323, symmetrically substituted moieties of hydroxymethylbenzyls and phenylalanines extend<br />

towards the S2/S2' and S1/S1' subsites respectively and are involved in van der Waals interactions with<br />

the hydrophobic residues of these pockets [42].<br />

The interaction of DMP323 with the residues of HIV PR are restricted to the central four specificity<br />

subsites of the active site. Despite this limited area of hydrophobic interaction and hydrogen bonding<br />

restricted to the central cyclic urea functionality, DMP323 is a very potent inhibitor of HIV PR with<br />

good antiviral activity in vitro (Table 5). The limited solubility of this compound was perhaps<br />

responsible for erratic oral availability in humans, and after short trials, DMP323 was withdrawn from<br />

the clinical investigation. Nevertheless, the discovery of this class of compounds represents a very<br />

interesting and, <strong>by</strong> now, classical example of de novo structure-<strong>based</strong> drug design.<br />

<strong>Design</strong> and <strong>Structure</strong> of AG1284<br />

Another compound discovered <strong>by</strong> the application of de novo structure-<strong>based</strong> design is AG1284 [43]. In<br />

the initial design of a lead compound, the nonpeptidic hydroxyethyl-t-butylbenzylamide portion of<br />

LY289612 occupying the S1' and S2' subsites was retained as a “starting module.” In attempting to fill<br />

the pockets related <strong>by</strong> the dimer two-fold symmetry it was discovered that, <strong>by</strong> extending a two-carbon<br />

fragment from the central hydroxyl carbon, the S1 subsite could be accessed <strong>by</strong> an aromatic ring. The<br />

ring was oriented orthogonal to the observed P1 phenyl group of the classical inhibitors and this allowed<br />

further extension off the ortho position towards the S2 subsite. In order to maintain the critical hydrogen<br />

bond to the flap water Wat301, in the initial compounds an acylated amino group was used, replaced in<br />

subsequent designs <strong>by</strong> a benzamide functionality. In this model, the geometry of the hydrogen bonds to<br />

the flap water was somewhat perturbed, and the nitrogens of the t-butyl amides on both sides of the<br />

compound were in a position to interact favorably with solvent,<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_22.html [4/5/2004 4:45:29 PM]


Document<br />

Page 23<br />

potentially lowering the desolvation penalty. The absence of hydrogen-bonding interactions with the<br />

carbonyl oxygens of Gly27/27' was viewed as a positive factor, since the accumulated structural and<br />

mechanistic information suggested that formation of these hydrogen bonds may not be energetically<br />

favorable [23]. The compound was synthesized as a racemic mixture of two enantiomers of the central<br />

hydroxymethyl group and had the inhibition constant of 24 μM. Despite the modest binding constant of<br />

compound II (Table 6) and very low water solubility, the co-crystal structure with HIV PR was solved at<br />

2.3 Å resolution, providing a critical starting point for further design. The inhibitor was found to bind<br />

largely as anticipated with the two aromatic rings occupying the S1 and S1' subsites and the two<br />

benzamide carbonyls forming hydrogen bonds to the flap water Wat301. Both benzamide nitrogens<br />

interact via a string of highly ordered water molecules with the amide nitrogens of Asp29/29'. The<br />

crystal structure of the complex indicated that the S enantiomer was the more active component of the<br />

racemic mixture and this was confirmed <strong>by</strong> stereoselective synthesis of subsequent compounds [44].<br />

In the subsequent designs, the ortho-substituted benzyl rings were consecutively replaced <strong>by</strong> larger<br />

naphthyl groups that occupied more of the S1–S3 and S1'–S3' subsites. The increased area of<br />

hydrophobic interactions with the residues in these subsites was reflected in a substantial improvement<br />

in the<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_23.html [4/5/2004 4:45:41 PM]


Document<br />

Page 24<br />

binding constant and also in reduced aqueous solubility. Also, due to the very tight fit of both naphthyl<br />

moieties in the S1 and S1' subsites, subsequent design targeting the S3/S3' subsites proved to be difficult<br />

and synthetically challenging [44]. In the search for a simpler solution, the di-tertiary amides were<br />

designed using the crystal structure of compound II (Table 6) as a starting model. Branching from the<br />

amide nitrogens provided an interesting possibility to access S2–S3/S2'–S3' subsites while<br />

simultaneously increasing the solubility and stability of the compounds. In the first design, the<br />

hydroxyethyl moieties were fused to the amide nitrogens and the hydroxyl groups were intended to form<br />

hydrogen bonds with the amide nitrogens of Asp29/29' (compound III in Table 6). The addition of both<br />

hydroxyethyl groups resulted in a rather significant increase in the binding constant, and the racemic<br />

mixture had the K i of 1.1μM. When the crystal structure of compound III complexed with HIV PR was<br />

solved at 2.2 Å resolution, it was observed that the inhibitor had undergone an inversion in binding<br />

mode relative to the secondary amide series. The phenyl groups of compound III occupied the S2/S2'<br />

subsites, switching positions with the t-butyl groups, which were in turn occupying the S1/S1' pockets<br />

(Figure 6). Due to this change in binding mode, the R enantiomer would be expected to be preferred<br />

relative to S. The final position of the hydroxyethyl moieties was less effected <strong>by</strong> the change, and both<br />

hydroxyls were within hydrogen-bonding distance from the amide nitrogens of Asp 29/29'. In the S2/S2'<br />

pockets, the phenyl groups occupied only a fraction of subsites, but the interaction was strengthened <strong>by</strong><br />

highly ordered water molecules involved in electrostatic interaction with the aromatic rings and <strong>by</strong><br />

forming hydrogen bonds to Asp30/30'. Interestingly the position of the hydrogen bonds with respect to<br />

the flap water was significantly disturbed in the new binding mode, and the conserved Wat301 was no<br />

longer tetrahydrally coordinated [43,45].<br />

This unanticipated change in binding mode presented a potential for new avenues of design different<br />

from those of the secondary amides. The ability to design into neighboring subsites depends to a large<br />

extent on the positions of bond vectors suitable for substitution in the bound conformation of a given<br />

inhibitor. These vectors in the crystallographically discovered new binding mode of compound III were<br />

positioned ideally to access unfilled space in the S3/S3' pockets. The discovery of this new conformation<br />

of compound III highlighted the power of crystallographic feedback in the process of inhibitor design<br />

and, without this structural information, further design in this series would have been severely impeded.<br />

Inspection of the crystal structure of compound III bound to the active site of HIV PR revealed<br />

lipophilic cavities extending off the S1/S1' subsites adjacent to the t-butyl groups of the benzamidine<br />

moiety. The cavities are bordered <strong>by</strong> flexible loops around Pro81/81' and previous crystallographic<br />

studies indicated that both loops can move back <strong>by</strong> up to 2.5 Å, extending the size and<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_24.html [4/5/2004 4:45:43 PM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_25.html (1 of 2) [4/5/2004 4:46:06 PM]<br />

Page 25


Document<br />

Figure 6<br />

Change of the binding mode of compound III observed during<br />

iterative design of AG1284. (a) Crystallographically determined<br />

binding mode of compound II. Pseudosymmetrically distributed aryl groups<br />

are bound in the S1 and S1' specificity subsites. (b) Crystallographically<br />

determined binding mode of compound III. Note the inversion of the binding<br />

mode. The ortho-substituted benzyl groups bind in a pseudosymmetric<br />

fashion in the S2 and S2' subsites.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_26.html (1 of 2) [4/5/2004 4:46:26 PM]<br />

Page 26


Document<br />

Page 27<br />

volume of the active site. With this in mind, the dimethylbenzyl group was attached to compound III and<br />

the additional phenyl ring was accommodated well in the lipophilic pocket of the S1'/S3' sides. As the<br />

S1 pocket was not fully occupied, a Monte Carlo-<strong>based</strong> De Novo Ligand-Generating program<br />

(MCDNLG) [46] was used to identify other amide substituents that might fill this subsite more<br />

effectively. From several moieties identified <strong>by</strong> the MCDNLG program, a larger cyclopentylethyl group<br />

showing very good shape complementarity to the S1/S3 subsite was selected for synthesis. In addition,<br />

due to the asymmetrical nature of this compound, additional space was identified at the bottom of the<br />

S2' pocket that was conveniently filled with either a methyl or a chlorine group on the 5 position of the<br />

benzamidine ring. The inhibition constant of the resulting compound (compound IV in the Table 6) was<br />

0.008 μM, which represents approximately a 2500-fold improvement over the first compound from this<br />

series.<br />

The crystal structure of compound IV or AG1284 complexed with HIV-1 PR was solved, revealing<br />

excellent complementarity between the ligand and protein. The ligand forms only 4 hydrogen bonds<br />

with either protein functional groups or ordered water molecules, in contrast to the nine hydrogen bonds<br />

formed <strong>by</strong> peptidomimetic LY289612, despite their similar binding affinities. The nonpeptidic character<br />

of AG1284 may have contributed to good oral bioavailability and pharmacokinetics in three animal<br />

species [43].<br />

Despite very good inhibitory potency on the enzyme level, AG1284 has rather modest antiviral activity<br />

in vitro (Table 6). The reason for this discrepancy is unclear but could be related to the low water<br />

solubility and higher affinity for membranes, which may effect cell partitioning. A similar lack of<br />

correlation between the potency of enzyme inhibition and antiviral activity has been previously observed<br />

with other HIV PR inhibitors [11].<br />

Hydroxypyrans and Hydroxycoumarins<br />

The lead compounds for the 4-hydroxypyran and 4-hydroxycoumarin series were discovered in<br />

biological screens as low potency inhibitors of HIV PR [47–49]. Successful structure solution of both<br />

lead compounds with HIV PR enabled rapid optimization of their enzyme inhibitory potencies and anti-<br />

HIV activities, and one of these compounds, U96988, has already entered Phase I clinical testing<br />

[49,50]. The binding mode of this type of inhibitor differs substantially from the classical<br />

peptidomimetic compounds and is somewhat similar to de novo-designed compounds from the cyclic<br />

urea series. In the case of 4-hydroxycoumarin, the two oxygen atoms of the lactone functionality are<br />

positioned within hydrogen-bonding distance of the two NH amides of Ile50/50' on the flap, replacing<br />

the ubiquitous water molecule Wat301. The 4-hydroxyl group (Table 5) is located within hydrogenbonding<br />

distance of the two catalytic<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_27.html [4/5/2004 4:46:27 PM]


Document<br />

Page 28<br />

aspartic acid residues Asp25/25' and this hydrogen-bonding network of the 4-hydroxycoumarin defines<br />

the essential pharmacophore of this new class of inhibitors. In the structure of U96988, this<br />

pharmacophore is pseudosymmetrically subsituted <strong>by</strong> an ethyl and a phenyl group at the C-3a and an<br />

ethyl and a benzyl group at the C-6a positions. These four substituents extend into the central core of<br />

S2/S2' subsites, where they make van der Waals contacts with the hydrophobic residues of the active site<br />

[49]. With a molecular weight of 362 U96988 is the smallest inhibitor of HIV PR in clinical testing. It<br />

suffers from rather low antiviral activity (ED90 of ~ 10 μM)but can be considered as the first in a series<br />

of this promising class of nonpeptidic HIV PR inhibitors.<br />

II. Structural Basis of Resistance of HIV PR Inhibitors<br />

The dimeric character and the two-fold symmetry of the active site, in which the monomers contribute<br />

equivalent residues to symmetrically distributed specificity subsites, led to early speculations that HIV<br />

PR may be less susceptible to resistance than, for example, reverse transcriptase. In the case of retroviral<br />

proteases, a single base mutation in the viral genome corresponds to two changes in the threedimensional<br />

structure and two structurally identical changes in the active site could result in an enzyme<br />

with a drastically modified specificity profile and impaired catalytic activity. Identification of HIV PR<br />

variants in cell-culture experiments clearly indicated, however, that this class of drugs is not immune to<br />

the challenge of viral resistance. It should be stressed that HIV, unlike other human viruses, is<br />

characterized <strong>by</strong> a dynamic viral turnover in the steady state [51,52]. The rapid replication rate coupled<br />

with the lengthy duration of infection will favor the emergence of resistant mutants to targeted antiviral<br />

agents [53].<br />

The accumulated data from cell-culture sequential-passage experiments with several structurally<br />

different inhibitors and from the resistant variants identified during clinical exposure to four HIV PRtargeting<br />

drugs indicate a very complex pattern of mutations in the structure of HIV PR. In contrast to<br />

mutations in the reverse transcriptase, which frequently cause multihundredfold resistance [54], single<br />

base changes in the HIV PR gene (i.e., two identical substitutions per protease dimer) lead in most cases<br />

to 5–10-fold decrease in the antiviral potency of a given drug [11]. It has been shown for the most<br />

clinically studied HIV PR inhibitors, such as indinavir and ritonavir, that the clinical manifestation of<br />

resistance (increase in the viral load and decrease in the CD4 count) requires the simultaneous<br />

appearance of several mutations [55,56]. For example the resistant HIV strain isolated from patients<br />

exposed for 40 weeks to indinavir carried mutations at residues 10/10'L > R, 46/46'M > I, 63/63'L > P,<br />

82/82'V > T, and 84/84'I > V [59,60]. However, the combination of these five<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_28.html [4/5/2004 4:46:29 PM]


Document<br />

Figure 7<br />

Cartoon representation of the HIV PR dimer. The sites of primary<br />

resistance-causing mutations in the active site are indicated. For clarity, the names of the<br />

residues are shown for one monomer only.<br />

Page 29<br />

mutations (ten assuming the dimeric nature of HIV PR) changed the susceptibility of the resistant strain<br />

to indinavir <strong>by</strong> only eight-fold if compared to the wild type HIV.<br />

The resistance-causing mutations are localized in a few “hot spots” in the structure of HIV PR and can<br />

be divided into two groups. The first group consists of the primary mutations located directly in the<br />

active site and includes changes at residues Val82/82', Ile84/84', somewhat less frequently at Gly48/48'<br />

and, in the case of nelfinavir, Asp30/30' (Figure 7). Residues 82/82' and 84/84' are located on the<br />

flexible loops that form the outer walls of the S3/S3' and S1/S1' subsites, respectively. In the resistant<br />

variants, valine 82/82' is most frequently substituted <strong>by</strong> the smaller side chain of alanine or the larger<br />

side chains of phenylalanine or isoleucine [57,58]. The change in position 82/82' is usually accompanied<br />

<strong>by</strong> a substitution of Ile84/84', most commonly to the smaller amino acids alanine or valine [57]. From<br />

the clinically tested compounds, ritonavir and indinavir, which were optimized to form strong<br />

hydrophobic interactions with the side chain of Val82/82' in the S3 subsite, suffer most significantly<br />

from any change at this position. On the other hand, the antiviral activities of saquinavir, and nelfinavir,<br />

which do not form any interaction in the S3/S3' subsite are not affected <strong>by</strong> mutations at Val82/82' and<br />

are only marginally cross resistant to changes involving Ile84/84' [57,58,62]. The resistance-causing<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_29.html [4/5/2004 4:46:39 PM]


Document<br />

Page 30<br />

mutation of Asp30/30' to asparagine seems to be specific for nelfinavir and was initially observed in cellculture<br />

sequential passage experiments [62]. Recently, the same phenotypic change was confirmed as<br />

the predominant mutation in the resistant variants appearing in AIDS patients exposed to low doses of<br />

this HIV PR inhibitor [64]. The molecular basis of resistance involving this mutation can be rationalized<br />

as follows: in the crystal structure of nelfinavir with the wild type HIV PR, the 3-hydroxyl group of the<br />

2,3-substituted phenyl group is within hydrogen-bonding distance of the carboxylate oxygen of Asp30 in<br />

the S2 subsite. Due to the expected coulombic character of this interaction, the hydrogen bond formed<br />

with the negatively charged carboxylate of Asp30 would be expected to be a relatively strong one. The<br />

change of the negatively charged carboxylate of Asp30/30' to the amide oxygen of the asparagine side<br />

chain should reduced the strength of this interaction. Apparently the loss in the enthalpic contribution to<br />

the free energy of binding is only partially balanced <strong>by</strong> the entropic gain caused <strong>by</strong> the difference in<br />

desolvation of a charged vs. neutral side chain of the receptor, leading to decreased binding affinity of<br />

nelfinavir and eventually to viral resistance.<br />

An additional resistance-causing mutation that qualifies as a primary mutation involves the change of<br />

Gly48/48' to valine. This particular mutation seems to be specific for saquinavir and was observed both<br />

in cell-culture sequential passage experiments and in AIDS patients exposed to this inhibitor [61,65].<br />

Located on the lower strands of the active-site forming flaps, Gly48/48' can be considered a part of the<br />

S4/S4' subsites. The replacement of the glycine hydrogen <strong>by</strong> the rigid side chain of valine has most<br />

likely a dual effect: first it has a direct impact on the interaction of the quinoline moiety of saquinavir<br />

with the active site of HIV PR, and second it may change the mobility of the flaps, which in turn will<br />

effect the binding kinetics of the natural substrates or inhibitors. Although none of the other clinically<br />

tested inhibitors form any interaction with this part of the flap, the HIV variants with mutation of<br />

Gly48/48' seem to be cross-resistant to all compounds, which is reflected <strong>by</strong> a 3–5-fold reduction of<br />

their antiviral activity [55,62].<br />

While the effect of primary mutations on reduced binding affinities of inhibitors can be at least partially<br />

explained in view of the accumulated structural data, the function of secondary, or compensatory<br />

mutations in the resistant HIV PR is difficult to rationalize as yet. The predominant compensatory<br />

mutations observed in the resistant variants involve residues Leu63/63', Ala71/71', Met46/46',<br />

Asn88/88', Leu10/10', and Leu90/90' (Figure 8) [60,63]. Changes of these residues alone do not confer<br />

viral resistance, but their appearance increases the viability of the virus carrying the primary mutations<br />

in the active site of protease. All these residues are located far away from the active site of HIV PR do<br />

not participate in any apparent way in the inhibitor binding and it seems unlikely that they form a longrange<br />

interaction with the natural<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_30.html [4/5/2004 4:46:41 PM]


Document<br />

Figure 8<br />

Cartoon representation of the HIV PR dimer. The sites of compensatory<br />

mutations are indicated.<br />

substrate. Also, the reported compensatory mutations are conservative in nature and have no effect on<br />

the overall distribution of atomic charges on the surface of HIV PR.<br />

Page 31<br />

Sequence polymorphism at the Leu63/63' position, located on the surface at the base of HIV PR, has<br />

been observed in clinical isolates of the virus not exposed to any HIV PR inhibitors. Variations of<br />

Ala71/71', where the side chains are buried very close to Leu63/63', are less commonly found in clinical<br />

isolates. After a prolonged challenge <strong>by</strong> HIV PR inhibitors, Leu63/63' changes to proline and Ala71/71'<br />

to valine.<br />

The side chains of Met46/46' are fully exposed to solvent and these residues are located on the βhairpins<br />

that form the active side flaps. It has been speculated that the compensatory change of<br />

Met46/46' to isoleucine or phenylalanine may affect the dynamics of the flap movement, which in turn<br />

could influence the rates of catalytic activity of HIV PR impeded <strong>by</strong> the primary mutations in the active<br />

site [58].<br />

Any changes to Asn88/88' and Leu90/90', buried in the body of HIV PR, most likely affect the structural<br />

stability of the enzyme. The side chains of Asn88/88' form buried hydrogen bonds and replacement of<br />

this residue <strong>by</strong> aspartic acid or serine not only eliminates some of these bonds but also introduces<br />

unfavorable interactions in the core of HIV PR. Similarly, Leu90/90' is buried in a tight hydrophobic<br />

space close to the “fireman's grip” motif, which<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_31.html (1 of 2) [4/5/2004 4:46:51 PM]


Document<br />

involves the catalytic asparates 25/25'. The structural effect of a mutation of Leu90/90' to the larger<br />

methionine is rather difficult to predict since it can either rigidify or destabilize the HIV PR dimer or it<br />

may have an effect on the catalytic efficiency of the “resistant” enzyme.<br />

Page 32<br />

The complicated pattern of HIV resistance to protease inhibitors, in particular the appearance of<br />

compensatory mutations that alone do not confer any resistance, suggests that the key to understanding<br />

the basis of decreased susceptibility of the virus to a given drug is the kinetics of specific processing of<br />

the GAG and GAG-POL polyproteins. The reduction in sensitivity of a mutant HIV PR towards any<br />

inhibitor can be conveniently reflected <strong>by</strong> the ratio of K i mutant/K i wild type. However, this reduced<br />

inhibitor sensitivity is only one component that distinguishes mutant-form from wild-type proteases. For<br />

virus encoding of a mutant HIV PR to be viable, the mutant protease must be capable of a minimal<br />

(although not yet quantified) level of enzymatic activity towards all substrates required for maturation of<br />

the virions. This proteolytic efficiency is reflected in the specificity constants (K cat/K m) as determined for<br />

mutant and wild type HIV PRs. In order to rationalise these potentially conflicting relationships between<br />

enzymes, substrates, and inhibitors, Gulnik and his colleagues [66] introduced the term “Vitality<br />

Factor,” in which<br />

In order for the “Vitality Factor” to be predictive for the level of resistance expected for a particular drug<br />

or combination of drugs for a given resistant strain of HIV, the determination of the specificity constants<br />

(K cat/K m for mutants) must be repeated for all nine known substrates processed <strong>by</strong> HIV PR. The<br />

inhibition constants of a given compound should not depend on the substrate, but the K cat/K m ratios do<br />

and therefore vitality values will differ for different substrates. It will be expected that the mean for all<br />

nine “vitality” values will be predictive for the change in antiviral activity for a particular compound.<br />

Although those data will be derived from in vitro experiments and are clearly not without some<br />

limitations, they may help in understanding the molecular basis of resistance and may contribute some<br />

value to possible multidrug strategies for the clinical management of AIDS.<br />

III. Perspective<br />

HIV PR inhibitors with acceptable oral availability and pharmacokinetic properties offer great promise<br />

for the treatment of HIV infection and AIDS. Efficacy studies of indinavir, ritonavir, or nelfinavir using<br />

plasma viral RNA as a marker have demonstrated up to three log reductions in RNA copy numbers that<br />

are<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_32.html [4/5/2004 4:46:54 PM]


Document<br />

Page 33<br />

sustained in many patients [67–69]. In contrast, nucleoside antiretroviral therapy that targets reverse<br />

transcriptase rarely results in more than one log reduction of viral RNA, indicating fairly poor inhibition<br />

of viral replication <strong>by</strong> this class of compounds. One of the reasons for the apparent greater in vivo<br />

antiviral activity of HIV PR inhibitors could be the mechanistic difference of the two enzymes and their<br />

respective activities in the viral life cycle. However, growing evidence of retroviral resistance to<br />

protease inhibitors remains a concern. The availability of several chemically distinct HIV PR inhibitors,<br />

including the second generation of compounds currently under preclinical development, offers a<br />

possibility of combining two or more drugs that share little cross-resistance. Also, it seems reasonable to<br />

evaluate these compounds in combination with various nucleoside and nonnucleoside reverse<br />

transcriptase inhibitors. Early clinical data from such combination therapy indicates reduction of<br />

retroviral RNA in plasma to levels lower than the currently available limit of detection [70]. This is the<br />

first indication that the application of well-chosen combination therapy can place AIDS patients in<br />

prolonged virologic and clinical remission.<br />

Undoubtedly, protein crystallography and other elements of structure-<strong>based</strong> drug design were widely<br />

applied in the discovery of HIV PR inhibitors. It will be prudent to assume that, in the absence of<br />

structural feedback, rapid discovery of several chemically different and potent inhibitors of HIV PR<br />

would have been severely impeded if not even impossible. However, structure-<strong>based</strong> drug design still<br />

remains a new and developing technology. Further success of this drug discovery technique largely<br />

depends on the development of methods of computational chemistry. Several computational approaches<br />

such as ALADDIN [71], DOCK [72], and MCDNLG [46] have been applied with a limited degree of<br />

success in a search for novel inhibitors of HIV PR and these methods will be developed further. The<br />

most difficult and challenging computational task required for full implementation of structure-<strong>based</strong><br />

drug design involves assigning a priority to designed compounds before their synthesis, <strong>by</strong> computation<br />

of the absolute free energy of binding or <strong>by</strong> prediction of the relative difference in the binding constants<br />

of chemically related compounds. While the former approach is technically very difficult, due to the size<br />

of configurational space that must be sampled and the limited accuracy of the force field that describes<br />

atomic interactions in the molecular system [73], the latter approach has had some successes [74,75].<br />

Nevertheless, owing to the various assumptions and approximations that underlie these techniques, such<br />

methods are useful only as order-of-magnitude estimates [74]. Further improvements of these methods<br />

heavily depend on the availability of structural and thermodynamic data for several closely related<br />

compounds that could be used to calculate parameters required for the implementation of such<br />

thermodynamic-integration cycles. The large number of high-resolution crystal structures of HIV PR<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_33.html [4/5/2004 4:46:56 PM]


Document<br />

Page 34<br />

complexed with various inhibitors offers a unique opportunity for the development of such<br />

computational methods if the structural data can be coupled with thermodynamic measurements of<br />

inhibitor-protein binding. These include direct measurements <strong>by</strong> microcalorimetry of the association<br />

constant, K, and in addition the enthalpy, entropy, heat capacity, and stoichiometry of binding. The<br />

combination of such thermodynamic and structural data will lead to a more precise understanding of the<br />

factors that influence binding and, ultimately, will lead to new general design principles that can be<br />

applied to drug discovery in the area of AIDS as well as other challenging diseases.<br />

Acknowledgments<br />

I wish to thank all my co-workers from Agouron who contributed to these studies, in particular J.<br />

Davies, S. Reich, M. Melnick, V. Kalish, A. Patick, L. Musick, and B-W. Wu. Steven Kaldor from Ely<br />

Lilly is acknowledged for his contribution in designing AG1343 (nelfinavir). I would like to thank<br />

Richard Ogden for critical reading of the manuscript and D. Olson for expert assistance in preparing the<br />

manuscript.<br />

References<br />

1. Mitsuya H, Yarchoan R Broder S. Molecular Targets for AIDS therapy. Science 1990;<br />

249:1533–1543.<br />

2. DeClerq E. Toward improved anti-HIV chemotherapy: Therapeutic intervention with HIV infections.<br />

J. Med. Chem. 1995; 38:2491–2517.<br />

3. Tomaselli AG, Howe JW, Sawyer TK, Wlodawer A, Henrikson RL. HIV-1 protease as a target for<br />

drug design. Chimica Oggi 1991; 9:6–14.<br />

4. Ding J, Das K, Yadav PNS, Hsiou Y, Zhang W, Hughes SH, Arnold E. Structural studies of HIV-1<br />

reverse transcriptase and implications for drug design. In: <strong>Structure</strong>-Based <strong>Drug</strong> <strong>Design</strong>. New York:<br />

Marcel Dekker 1996, in press.<br />

5. Perno CF, Bergamini A, Pesce CD. Inhibition of the protease of human immunodeficiency virus<br />

blocks replication and infectivity of the virus in chronically infected macrophages. J. Infect. Dis. 1993;<br />

168:1148–1156.<br />

6. Kohl NE, Emini EA, Schleif WA, Davis LJ, Heimbach JC, Dixon RAF, Scolnick EM, Sigal, IS.<br />

Active human immunodeficiency virus protease is required forviral infectivity. Proc. Natl. Acad. Sci.<br />

USA, 1988; 85:4186–4690.<br />

7. McQuade TK, Tomasselli AG, Liu L, Karacostas V, Moss B, Sawyer TK, Heinrikson RL, Tarpley<br />

WG. A synthetic HIV-1 protease inhibitor with antiviral activity arrests HIV-like particle maturation.<br />

Science 1990; 247:454–456.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_34.html (1 of 2) [4/5/2004 4:46:59 PM]


Document<br />

8. Kaplan AH, Zack JA, Knigge M, Paul DA, Kempf DJ, Norbeck DW, Swanstrom R. Partial inhibition<br />

of the human immunodeficiency virus type 1 protease results in aberrant virus assembly and the<br />

formation of noninfectious particles. J. Virol. 1993; 67:4050–4055.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_34.html (2 of 2) [4/5/2004 4:46:59 PM]


Document<br />

9. Ratner L. HIV life cycle and genetic approaches. Perspectives in drug discovery and design. 1993;<br />

1:3–22.<br />

10. Roberts NA, Martin JA, Kinchington D, Broadhurst AV, Craig JC, Duncan IB, Galpin SA, Handa<br />

BJ, Kay J, Krohn A, Lambert RW, Merrett JH, Mills JS, Parkes KEB, Redshaw S, Ritchie AJ, Taylor<br />

DL, Thomas GJ, Machin PJ. Rational design of peptide-<strong>based</strong> HIV proteinase inhibitors. 1990;<br />

248:358–361.<br />

11. Winslow DL, Otto MJ. HIV protease inhibitors. Current Science 1995; 9:183–192.<br />

12. Toh H, Ono M, Saigo K, Miyata T. Retroviral protease-like sequence in the yeast transposon TY1.<br />

Nature 1985; 315:691–693.<br />

13. Davies DR. The structure and function of the aspartic proteinases. Annu. Rev. Biophys. Biophys.<br />

Chem. 1990; 19:189–215.<br />

14. Pearl LN, Taylor WR. A structural model for the retroviral protease. Nature 1987; 329:351–354.<br />

Page 35<br />

15. Navia MA, Fitzgerald PMD, McKeever BM, Leu C-T, Heimbach JC, Herber WK, Sigal IS, Darke<br />

PL, Springer JP. Three-dimensional structure of aspartyl protease from human immunodeficiency virus<br />

HIV-1. Nature 1989; 333:615–620.<br />

16. Wlodawer A, Miller M, Jaskolski M, Sathyanarayana BK, Baldwin E, Webber IT, Selk LM,<br />

Clawson L, Schneider J, Kent SBH. Conserved folding in retroviral proteases: Crystal structure of a<br />

synthetic HIV-1 protease. Science 1989; 245:616–621.<br />

17. James MNG, Sielecki A. <strong>Structure</strong> and refinement of penicillopepsin at 1.8 Å resolution. J. Mol.<br />

Biol. 1983; 163:299–361.<br />

18. Wlodawer A, Erickson JW. <strong>Structure</strong>-<strong>based</strong> inhibitors of HIV-1 protease. Annu. Rev. Biochem.<br />

1993; 62:543–585.<br />

19. Schechter I, Burger A. On the size of the active site in proteases. Biochem. Biophys. Res. Commun.<br />

1967; 27:157–162.<br />

20. Lam PYS, Jadhav PK, Eyermann CJ, Hodge CN, Ru Y, Bacheler LT, Meek JL, Otto MJ, Rayner<br />

MM, Wong N, Chang CH, Webber PC, Jackson DA, Sharpe TR, Erickson-Viitanen S. Rational design<br />

of potent bioavailable, nonpeptide cyclic ureas as HIV protease inhibitors. 1994; 263:380–384.<br />

21. Pettit SC, Michael SF, Swanstrom R. The specificity of the HIV-1 protease. Perspectives in <strong>Drug</strong><br />

Discovery and <strong>Design</strong>. 1993; 1:69–83.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_35.html (1 of 2) [4/5/2004 4:47:02 PM]


Document<br />

22. Krausslich HG, Ingraham RH, Skoog MT, Wimmer E, Pallai PV, Carter CA. Activity of purified<br />

biosynthetic proteinase of human immunodeficiency virus on natural substrates and synthetic peptides.<br />

Proc. Natl. Acad. Sci. USA. 1989; 86:807–811.<br />

23. Appelt K. Crystal structures of HIV-1 protease—inhibitor complexes. Perspectives in <strong>Drug</strong><br />

Discovery and <strong>Design</strong>. 1993; 1:23–48.<br />

24. Miller M, Schneider J, Sathanarayana BK, Toth MV, Marshall GR, Clawson L, Selk L, Kent SBH,<br />

Wlodawer A. <strong>Structure</strong> of complex of synthetic HIV-1 protease with a substrate-<strong>based</strong> inhibitor at 2.3 A<br />

resolution. Science. 1989; 246:1149–1151.<br />

25. Jaskolski M, Tomasselli AG, Sawyer TK, Staples RL, Heinrikson RL, Schneider J, Kent SBH,<br />

Wlodawer A. <strong>Structure</strong> at 2.5 Å resolution of chemically synthesized Human Immunodeficiency Virus<br />

type 1 protease complexed with a hydroxyethylene-<strong>based</strong> inhibitor. v Biochemistry. 1991;<br />

30:1600–1609.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_35.html (2 of 2) [4/5/2004 4:47:02 PM]


Document<br />

Page 36<br />

26. Swain AL, Miller M, Green J, Rich DH, Schneider J, Kent, SBH, Wlodawer A. X-ray<br />

crystallographic structure of a complex between a synthetic protease of Human Immunodeficiency Virus<br />

1 and a substrate-<strong>based</strong> hydroxyethylamine inhibitor. Proc. Natl. Acad. Sci. USA. 1990; 87:8805–8809.<br />

27. Krohn A, Redshaw S, Ritchie JC, Graves BJ, Hatada MH. Novel binding mode of highly potent HIV<br />

proteinase inhibitors incorporating the (R)-hydroxyethylamine isostere. J. Med. Chem. 1991;<br />

34:3340–3342.<br />

28. Erickson J, Neidhart DJ, VanDrie J, Kempf DJ, Wang XC, Norbeck DW, Plattner JJ, Rittenhouse<br />

JW, Turon M, Wideburg N, Kohlbrenner WE, Simmer R, Helfrich R, Paul DA, Knigge, M. <strong>Design</strong>,<br />

activity, and 2.8 Å crystal structure of a C2 symmetric inhibitor complexed to HIV-1 protease. Science.<br />

1990; 249:527–533.<br />

29. Hosur MV, Bhat TN, Kempf DJ, Baldwin ET, Liu B, Gulnik S, Wideburg NE, Norbeck DW, Appelt<br />

K, Erickson JW. Influence of stereochemistry on activity and binding modes for C2 symmetry-<strong>based</strong><br />

diol inhibitors of HIV-1 protease. J. Amer. Chem. Soc. 1994; 116:847–855.<br />

30. Kempf DJ, Marsh KC, Denissen JF, McDonald E, Vasavanonda S, Flentge CA, Green BE, Fino L,<br />

Park CH, Kong XP, Wideburg NE, Saldivar A, Ruiz L, Kati WM, Sham HL, Robins T, Stewart KD,<br />

Hsu A, Plattner JJ, Leonard JM, Norbeck DW. ABT-538 is a potent inhibitor of human<br />

immunodeficiency virus protease and has high oral availability in humans. Proc. Natl. Acad. Sci. USA.<br />

1995; 92:2484–2488.<br />

31. Kempf DJ, Marsh KC, Fino LC, Bryant P, Craig-Kennard A, Sham HL, Zhao C, Vasavanonda S,<br />

Kohlbrenner WE, Wideburg NE, Saldivar A, Green BE, Herrin T, Norbeck D.W. <strong>Design</strong> of orally<br />

available, symmetry-<strong>based</strong> inhibitors of HIV protease. Bioorg. Med. Chem. Lett. 1994; 2:847–858.<br />

32. Holloway MK, Wai JM, Halgren TA, Fitzgerald PMD, Vacca JP, Dorsey BD, Levin RB, Thompson<br />

WJ, Chen LJ, deSolms SJ, Gaffin N, Ghosh AK, Giuliani EA, Graham SL, Guare JP, Hungate RW, Lyle<br />

TA, Sanders WM, Tucker TJ, Wiggins M, Wiscount CM, Woltersdorf OW, Young SD, Darke PL,<br />

Zugay JA. A priori prediction of activity fro HIV-1 protease inhibitors employing energy minimization<br />

in the active site. J. Med. Chem. 1995; 38:305–317.<br />

33. Vacca JP, Guare JP, deSolms SJ, Sanders WM, Giuliani EA, Young SD, Darke PL, Zugay J, Sigal<br />

IS, Schleif W, Quintero J, Emini E, Anderson P, Huff JR. L-687,908, a potent hydroxyethylenecontaining<br />

HIV protease inhibitor. J. Med. Chem. 1991; 34:1225–1228.<br />

34. Dorsey BD, Levin RB, McDaniel SL, Vacca JP, Guare JP, Darke PL, Zugay JA, Emini EA, Schleif<br />

WA, Quintero LC, Lin JH, Chen I-W, Holloway MK, Fitzgerald PMD, Axel MG, Ostovic D, Anderson<br />

PS, Huff JR. L-735,524: the design of a potent and orally available HIV protease inhibitor. J. Med.<br />

Chem. 1994; 37:3443–3451.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_36.html (1 of 2) [4/5/2004 4:47:03 PM]


Document<br />

35. Kaldor SW, Hammond M, Dressman BA, Frtiz JE, Crowell TA, Hermann RA. New peptide<br />

isosteres useful for the inhibition of HIV—1 protease. Bioorg. Med. Chem. Lett. 1994; 4:1385–1390.<br />

36. Kaldor SW, Dressman BA, Hammond M, Appelt K, Burgess JA, Lubbehausen PP, Muesing MA,<br />

Hatch SD, Wiskerchen MA, Baxter AJ. Isophthalic acid derivatives: amino acid surrogates for the<br />

inhibition of HIV-1 protease. Bioor. Med. Chem. Lett. 1995; 5:721–726.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_36.html (2 of 2) [4/5/2004 4:47:03 PM]


Document<br />

Page 37<br />

37. Kaldor SW, Appelt K, Fritz JE, Hammond M, Crowell TA, Baxter AJ, Hatch SD, Wiskerchen MA,<br />

Muesing MA. A systemic study of P1–P3 spanning sidechains for the inhibition of HIV-1 protease.<br />

Bioorg. Med. Chem. Lett. 1995; 5:715–720.<br />

38. Kalish JV, Tatlock JH, Davies JF, Kaldor SW, Dressman BA, Reich S, Pino M, Nyugen D, Appelt<br />

K, Musick L, Wu B-W. <strong>Structure</strong>-<strong>based</strong> drug design of nonpeptidic P2 substituents for HIV-1 protease<br />

inhibitors. Bioorg. Med. Chem. Lett. 1995; 5:727–732.<br />

39. Kaldor SW, Kalish VJ, Davies JF, Appelt K, Tatlock JH, Dressman BA, Campanale KM, Burgess<br />

JA, Lubbehusen PL, Muesing MA, Hatch SD, Shetty BV, Patick AK, Kosa MB, Khalil DA. AG1343: a<br />

potent orally bioavailable inhibitor of HIV-1 protease. J. Med. Chem. 1996; submitted for publication.<br />

40. Shetty B, Kosa M, Khalil DA, Webber S. Preclinical Pharmacokinetics and Distribution to Tissue of<br />

AG1343, an inhibitor of human deficiency virus type 1 protease. Antimicr. Ag. and Chemoth. 1996;<br />

40:110–114.<br />

41. Quart BD, Chapman SK, Peterkin J, Webber S, Oliver S. Phase 1 safety, tolerance,<br />

pharmacokinetics and food effect studies of AG1343—a novel HIV protease inhibitor. Abst. LB3. In:<br />

Proceedings of the 2nd National Conference on Human Retroviruses and Related Infections. 1995:163.<br />

42. Lam PYS, Jadhaw PK, Eyermann CJ, Hodge CN, Lee YR, Bacheler LT, Meek JL, Otto MJ, Rayner<br />

MM, Wong YN, Chang C-H, Weber PC, Jackson DA, Sharpe TR, Erickson-Viitanen S. Rational design<br />

of potent, bioavailable, nonpeptide cyclic ureas as HIV protease inhibitors. Science, 1994; 263:380–384.<br />

43. Reich SH, Melnick M, Davies JF, Appelt K, Lewis KK, Fuhry MA, Pino M, Trippe AJ, Nguyen D,<br />

Dawson H, Wu B-W, Musick L, Kosa M, Kahil D, Webber S, Gehlhaar DK, Andrada D, Shetty B.<br />

Protein structure-<strong>based</strong> design of potent, orally bioavailable, nonpeptide inhibitors of human<br />

immunodeficiency virus protease. Proc. Natl. Acad. Sci. 1995; 92:3298–3302.<br />

44. Reich SH, Melnick M, Pino M, Fuhry MA, Trippe AJ, Appelt K, Davies JF, Wu B-W, Musick L.<br />

<strong>Structure</strong>-<strong>based</strong> design and synthesis of substituted 2-butanols as nonpeptidic inhibitors of HIV protease:<br />

secondary amide series. J. Med. Chem. 1996; 39:2781–2794.<br />

45. Melnick M, Reich SH, Lewis KK, Mitchell L, Ngyen D, Trippe AJ, Dawson H, Davies JF, Appelt<br />

K, Wu B-W, Musick L, Gehlhaar DK, Webber S, Shetty B, Kosa M, Kahil D, Andrada D. Bis-tertiary<br />

amide inhibitors of the HIV-1 protease generated via protein structure-<strong>based</strong> iterative design. J. Med.<br />

Chem. 1996; 39:2795–2811.<br />

46. Gehlhaar DK, Moerder KE, Zichi D, Sherman CJ, Ogden RC, Freer ST. De novo design of enzyme<br />

inhibitors <strong>by</strong> Monte Carlo ligand generation. J. Med. Chem. 1995; 38:466–472.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_37.html (1 of 2) [4/5/2004 4:47:05 PM]


Document<br />

47. Tummino PJ, Ferguson D, Hupe L, Hupe D. Competitive inhibition of HIV-1 protease <strong>by</strong> 4-hydroxybenzopyran-2-ones<br />

and 4-hydroxyphenylpyran-2-ones. Biochem. Biophys. REs. Commun. 1994;<br />

200:1658–1664.<br />

48. Tummino PJ, Fergusson D. Hupe D. Competitive inhibition of HIV-1 protease <strong>by</strong> warfarin<br />

derivatives. Biochem. Biophys. Res. Commun. 1994; 201:290–294.<br />

49. Thaisrivongs S, Tomich PK, Watenpaugh KD, Chong KT, Howe WJ, Yang K-T, Strohbach JW,<br />

Turner SR, McGRath JP, Bohanon JCL, Mulichak AM, Spinelli PA, Hinshaw RR, Pagano PJ, Moon JB,<br />

Ruwart MJ, Wilkinson KF, Rush BD,<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_37.html (2 of 2) [4/5/2004 4:47:05 PM]


Document<br />

Page 38<br />

Zipp GL, Dalga RJ, Schwende FJ, Howard GM, Padbury GE, Toth LN, Zhao Z, Koeplinger KA,<br />

Kakuk TJ, Cole SL, Zaya RM, Piper RC, Jeffrey P. <strong>Structure</strong>-<strong>based</strong> design of HIV protease<br />

inhibitors: 4-hydroxycoumarins and 4-hydroxy-2-pyrones as nonpeptidic inhibitors. J. Med. Chem.<br />

1994; 37:3200–3204.<br />

50. Vara Prasad JVN, Para KS, Lunney EA, Ortwine DF, Dunbar Jr. JB, Fergusson D, Tummino PJ,<br />

Hupe D, Tait BD, Domagala JM, Humblet C, Bhat TN, Liu B, Guerin DMA, Baldwin ET, Erickson JW,<br />

Sawyer TK. Novel series of achiral low molecular weight and potent HIV-1 protease inhibitors. J. Am.<br />

Chem. Soc. 1994; 116:6989–6990.<br />

51. Ho DD, Neuman AU, Pereison AS, Chen W, Leonard JM, Markowitz M. Rapid turnover of plasma<br />

virions and CD4 lymphocytes in HIV-1 infection. Nature 1995; 373:123–126.<br />

52. Wie X, Ghosh SK, Taylor ME, Johnson VA, Emini EA, Deutsch P, Lifson JD, Bonhoeffer S, Nowak<br />

MA, Hahn BH, Saag MS, Shaw GM. Viral dynamics in HIV-1 infection. Nature 1995; 373:117–122.<br />

53. Coffin JM. HIV population dynamics in vivo: implications for genetic variation, pathogenesis and<br />

therapy. Science 1995; 267:483–489.<br />

54. De Clercq E. HIV resistance to reverse transcriptase inhibitors. Biochem. Pharmacol. 1994;<br />

47:155–169.<br />

55. Mo H, Markowitz M, Ho DD. Patterns of specific mutations in HIV-1 protease that confer resistance<br />

to protease inhibitors in clinical development. Third International Workshop on HIV drug resistance.<br />

Kauai, Hawaii, August 1994. Abstract 13.<br />

56. Markowitz M, Mo H, Kempf DJ, Norbeck D. Selection and analysis of human immunodeficiency<br />

virus type 1 variants with increased resistance ABT-538, a novel protease inhibitor. J. Virol 1995;<br />

69:701–706.<br />

57. Kaplan AH, Michael SF, Wehbie RS, Knige MF, Paul DA, Everitt L, Kempf DJ, Norbeck DW,<br />

Erickson JW, Swanstrom R. Selection of multiple human immunodeficiency virus type 1 variants that<br />

encode viral proteases with decreased sensitivity to an inhibitor of the viral protease. Proc. Natl. Acad.<br />

Sc. USA 1994; 91:5597–5601.<br />

58. Erickson JW, Baldwin ET, Bhat TN, Gulnik S, Leu B, Yr B. Structural basis of drug resistance to<br />

HIV-1 protease inhibitors. Third International Workshop on HIV <strong>Drug</strong> Resistance. Kuaui, Hawaii,<br />

August 1994, Abstract 2.<br />

59. Condra JH, Schleif WM, Blahy OH, Gabryelski LJ, Graham DJ, Quintero JC, Rhodes A, Robbins<br />

HL, Roth E, Shivaprakash M, Titus PL, Yang Y, Emini EA. Mutations in HIV protease conferring<br />

resistance to inhibitor L735,524. Abstract 187. In: Proceedings of the 2nd National Conference on<br />

Human Retroviruses and Related Infections. 1995:88.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_38.html (1 of 2) [4/5/2004 4:47:07 PM]


Document<br />

60. Condra JH, Schleif WA, Blahy OM, Gabryelski LJ, Graham DJ, Quintero JC, Rhodes A, Robbins<br />

HL, Roth E, Shivaprakash M. In vivo emergence of HIV-1 variants resistant to multiple protease<br />

inhibitors. Nature, 1995; 374:569–571.<br />

61. Jackobsen H, Yasargil K, Winslow JC, Craig JC, Krohn A, Duncan IB, Mous J. Characterization of<br />

human immunodeficiency virus type 1 mutants with decreased sensitivity to proteinase inhibitor RO<br />

31–8959. Virology, 1995; 206:527–534.<br />

62. Patick AK, Mo H, Markowitz M, Appelt K, Wu B-W, Musick L, Kalish V, Kaldor SW, Reich SH,<br />

Ho D, Webber S. Antiviral and resistance studies of AG1343, an<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_38.html (2 of 2) [4/5/2004 4:47:07 PM]


Document<br />

orally bioavailable inhibitor of human immunodeficiency virus protease. Antimicr. Ag. and<br />

Chemoth. 1996; 40:292–297.<br />

63. Korant B, Lu Z, Strack P, Rizzo C. HIV protease mutations leading to reduced inhibitor<br />

susceptibility. In: Intracellular Protein Catabolism. New York: Plenum Press, 1996:241–250.<br />

Page 39<br />

64. Patick AK, Duran M, Cao Y, Pei Z, Keller MR, Peterkin J, Chapman S, Anderson B, Markowitz M.<br />

Genotypic and phenotypic characterization of HIV-1 variants isolated from in vitro selection studies and<br />

from patients treated with the protease inhibitor, nelfinavir. Fifth International Workshop on HIV <strong>Drug</strong><br />

Resistance, Whistler, Canada, 1996:29.<br />

65. Jacobsen H, Brun-Vezinet F, Duncan I, Hanggi M, Ott M, Vella S, Weber J, Mous J. Genotypic<br />

characterization of HIV-1 from patients after prolonged treatment with protease inhibitor saquinavir. In:<br />

Abstracts of the 3rd International Workshop on HIV <strong>Drug</strong> Resistance. London: MediTech Media,<br />

1994:16.<br />

66. Gulnik SV, Suvorov LI, Liu B, Yu B, Anderson B, Mitsuya H, Erickson JW. Kinetic<br />

characterization and cross-resistance patters of HIV-1 protease mutants selected under in vitro drug<br />

pressure. Biochemistry 1995; 34:9282–9287.<br />

67. Stein DS, Fish DG, Chodakewitz J. A 24-week open-label phase I evaluation of the HIV protease<br />

inhibitor L 735,524. Abstract LB1 Second National Conference on Human Retroviruses and Related<br />

Infections, Washington D.C., 1995.<br />

68. Markowitz M, Jalil L, Hurley A. Evaluation of the antiviral activity of orally administered ABT-538,<br />

an inhibitor of HIV-1 protease. Abstract 185 Second National Conference on Human Retroviruses and<br />

Related Infections, Washington D.C., 1995.<br />

69. Gathe Jr. J, Burkhardt B, Hawley P, Conant M, Peterkin J, Chapman S. A randomized Phase II study<br />

of Virocept , a novel HIV protease inhibitor, used in combination with stavudine (D4T) vs. stavudine<br />

(D4T) alone. Abstract Mo.B.413 In: XI International Conference on AIDS, Vancouver, 1996:25.<br />

70. Hammer S. Advances in antiretroviral therapy and viral load monitoring. Abstract Mo.01 In:<br />

Abstracts of the XI International Conference on AIDS, Vancouver, 1996:2.<br />

71. VanDrie JH, Weininger D, Martin YC. Aladdin: an integrated tool for computer-assisted molecular<br />

design and pharmacophore recognition from geometric, steric, and substrate searching of threedimensional<br />

structures. Computer-Aided Mol. <strong>Design</strong>. 1989; 3:225–234.<br />

72. Kuntz ID, Blanley JM, Oatley SJ, Langridge R, Ferrin TE. A geometry approach to macromoleculeligand<br />

interactions. J. Mol. Biol. 1982; 161:269–278.<br />

73. Van Gunsteren WF Berendsen HJC. Computer simulation of molecular dynamics: methodology,<br />

application and perspectives in chemistry. Angew. Chem. Int. Ed. Eng. 1990; 29:992–996.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_39.html (1 of 2) [4/5/2004 4:47:09 PM]


Document<br />

74. Reddy RM, Varney MD, Kalish V, Viswanadhan VN, Appelt K. Calculation of relative differences<br />

in the binding free energies of HIV-1 protease: a thermodynamic cycle perturbation approach. J. Med.<br />

Chem. 1994; 37:1145–1152.<br />

75. Verkhivker G, Appelt K, Freer ST, Villafranca JE. Empirical free energy calculations of ligandprotein<br />

crystallographical complexes—knowledge-<strong>based</strong> ligand-protein interaction potentials applied to<br />

the prediction of human immunodeficiency virus protease binding affinity. Protein Engineering, 1995;<br />

8:677–691.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_39.html (2 of 2) [4/5/2004 4:47:09 PM]


http://legacy.netlibrary.com/reader/message.asp?message=811&BookID=12640&FileName=Page_40.html<br />

The requested page could not be found.<br />

Return to previous page<br />

http://legacy.netlibrary.com/reader/message.asp?message=811&BookID=12640&FileName=Page_40.html [4/5/2004 4:47:11 PM]


Document<br />

2<br />

Structural Studies of HIV-1 Reverse Transcriptase and Implications for<br />

<strong>Drug</strong> <strong>Design</strong><br />

Jianping Ding, Kalyan Das, Yu Hsiou,<br />

Wanyi Zhang, and Edward Arnold<br />

Center for Advanced Biotechnology and Medicine, and Rutgers University,<br />

Piscataway, New Jersey<br />

Prem N. S. Yadav<br />

University of Medicine and Dentistry of New Jersey, Piscataway, New Jersey<br />

Stephen H. Hughes<br />

ABL-Basic Research Program, NCI-Frederick Cancer Research and<br />

Development Center, Frederick, Maryland<br />

I. Introduction<br />

Page 41<br />

Like all other retroviruses, human immunodeficiency virus type 1 (HIV-1) contains the multifunctional<br />

enzyme reverse transcriptase (RT). Retroviral RTs have a DNA polymerase activity that can use either<br />

an RNA or a DNA template and an RNase H activity. HIV-1 RT is essential for the conversion of singlestranded<br />

viral RNA into a linear double-stranded DNA that is subsequently integrated into the host cell<br />

chromosomes [1–4]. In this conversion process HIV-1 RT catalyzes the incorporation of approximately<br />

20,000 nucleotides. Chemotherapeutic agents have been identified that target virtually all stages of the<br />

HIV-1 replication cycle (see review [5]). Since both the polymerase and RNase H activities of HIV-1<br />

RT are essential, inhibiting either step blocks viral replication. Therefore, HIV-1 RT is an important<br />

target for the treatment of AIDS. Two major classes of antiviral agents that inhibit HIV-1 RT<br />

polymerization have been identified; these are nucleoside RT inhibitors (NRTIs) (Figure la) and<br />

nonnucleoside RT inhibitors (NNRTIs) (Figure 1b). Nucleoside analogs, such as 3'-azido-2',3'dideoxythymidine<br />

(AZT), 2',3'-dideoxyinosine (ddI), 2',3'-dideoxycytidine (ddC), 2',3'-dideoxy-3'thiacyti-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_41.html [4/5/2004 4:47:13 PM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_42.html (1 of 2) [4/5/2004 4:47:17 PM]<br />

Page 42


Document<br />

Figure 1<br />

(a) Chemical structures of representative nucleoside analog<br />

inhibitors of HIV-1 RT. AZT: 3'-azido-2',3'-dideoxythymidine; d4T:<br />

2',3'-didehydro-2',3'-dideoxythymidine; ddI: 2',3'-dideoxyinosine;<br />

ddC: 2',3'-dideoxycytidine; 3TC: 2',3'-dideoxy-3'-thiacytidine;<br />

PMEA: 9-(2-phosphonylmethoxylethyl)adenine.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_42.html (2 of 2) [4/5/2004 4:47:17 PM]


Document<br />

(b) Chemical structures of representative nonnucleoside inhibitors of HIV-1 RT.<br />

Nevirapine: 11 -cyclopropyl-5,11 -dihydro-4-methyl-6H-dipyrido (3,2-b:2'3'-e)(1,4)<br />

diazepin-6-one; α-APA: αanilinophenylacetamine; TIBO:<br />

tetrahydroimidazo-(4,5,1-jk)<br />

(1,4)-benzo-diazepin-2(1H)-one and thione;<br />

pyridinones; HEPT: 1-{(2-hydroxyethoxy)<br />

methyl}-6-(phenylthio)thymine;<br />

BHAP: bis(heteroaryl)piperazine; TSAO:<br />

{2',5'-bis-O-(tert-butyldimethylsilyl)}-3'<br />

-spiro-5''-(4''-amino-1",2"-oxathiole)-2",<br />

2"-dioxide; L-737,126: 5-chloro-<br />

3-(phenylsulfonyl)indole-2-carboxamide; TBA:<br />

1-(2',6'-difluoro-phenyl)-1H,3H-thiozolo -(3,4-a) -benzimidazole;<br />

quinoxaline S-2720: 6-chloro-3,3-dimethyl-4-(isopropenyloxycarbonyl)<br />

-3-4-dihydroquinoxalin-2(1H)-thione.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_43.html (1 of 2) [4/5/2004 4:47:21 PM]<br />

Page 43


Document<br />

dine (3TC), and 2',3'-didehydro-2',3'-dideoxythymidine (d4T), have been widely used in the treatment of<br />

HIV-1 infections [5–7]. However, the effectiveness of these drugs is limited <strong>by</strong> their cytotoxicity and the<br />

rapid emergence of drug-resistant viral strains [5,8–12]. Nonnucleoside inhibitors, e.g., the HEPT<br />

derivatives [13], TIBO derivatives [14], nevirapine [15], pyridinones [16], BHAP derivatives [17], TBA<br />

derivatives [18,19], TSAO derivatives [20], α-APA [21], and quinoxalines (HBY) [22,23], are potent<br />

inhibitors of HIV-1 RT (see reviews [5,11,12]). While these inhibitors differ considerably in chemical<br />

structure, all of them are quite specific for HIV-1 RT and inhibit neither HIV-2 RT nor variety of<br />

cellular polymerases. Challenging a virus with these drugs<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_43.html (2 of 2) [4/5/2004 4:47:21 PM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_44.html [4/5/2004 4:48:34 PM]<br />

Page 44


Document<br />

Figure<br />

Continued<br />

Page 45<br />

rapidly selects viral strains containing drug-resistance mutations [9,24,25]. HIV-1 viral variants are<br />

known whose RT is resistant to all of the currently available drugs/inhibitors (see reviews<br />

[5,9,11,12,26,27]). In some cases, drug-resistant variants can be selected in very short periods of time<br />

[9], a consequence of the high viral load and rapid turnover of viral populations in infected individuals<br />

[28–30]. A better understanding of how these viral variants confer resistance should provide insight into<br />

the limitations of their genetic flexibility. In the past few years, substantial progress has been made in<br />

understanding the three-dimensional structure of HIV-1 RT. This paper will discuss the recent<br />

biochemical, genetic, and clinical data of HIV-1 RT in the context of the crystal structure of HIV-1 RT<br />

and prospects for development of more effective inhibitors of HIV-1 replication.<br />

II. Three-Dimensional <strong>Structure</strong>s of HIV-1 RT<br />

Three-dimensional crystal structures of HIV-1 RT have been determined both for the unliganded form of<br />

the protein and for complexes with either template-primer substrate or nonnucleoside inhibitors (Figure<br />

2 and Table 1). <strong>Structure</strong>s of HIV-1 RT have been determined in complexes with a series of NNRTIs,<br />

including nevirapine [31–33], 1051U91 (a nevirapine analog) [33], α-APA R95845 [34], α-APA<br />

R90385 [33], HEPT [33], 8-Cl TIBO (R86183) [35], and 9-Cl TIBO (R82913) [36,37]. The structure of<br />

HIV-1 RT in a ternary complex with a 19-mer/18-mer double-stranded DNA (dsDNA) template-primer<br />

and an antibody Fab fragment has been described [38]. In addition, structures of unliganded HIV-1 RT<br />

have also been solved in multiple crystal forms [39–43]. The structure of a polypeptide corresponding to<br />

the fingers and palm subdomains of the HIV-1 RT polymerase domain has also been determined [44].<br />

HIV-1 RT is an asymmetric heterodimer consisting of the p66 (66 kDa) and p51 (51 kDa) subunits. The<br />

N-terminal 440 residues of the p66 subunit constitute the polymerase domain and the C-terminal 120<br />

residues of p66 form the RNase H domain; the p51 subunit has the same amino acid sequence as the<br />

polymerase domain of the p66 subunit [1,2]. The polymerase domain of the p66 subunit has been<br />

likened to a human right hand. On this basis the subdomains of both p66 and p51 have been designated<br />

as fingers, palm, thumb, and connection (Figure 2) [31,38]. In the p66 subunit, the fingers, palm, and<br />

thumb subdomains form a large cleft that can accommodate the DNA substrate. The polymerase active<br />

site, which contains three strictly conserved aspartic acid residues (Asp110, Asp185, and Asp186), is<br />

located in the DNA-binding cleft and is part of the p66 palm subdomain (Figure 3) [31,38]. In the p51<br />

subunit, however, the thumb is rotated away from the fingers and the connection subdomain is folded<br />

over onto the palm subdomain between the fingers and thumb subdomains. As a<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_45.html [4/5/2004 4:48:36 PM]


Document<br />

Table 1 Crystal <strong>Structure</strong>s of HIV-1 Reverse Transcriptase<br />

PDB entry HIV-1 RT <strong>Structure</strong><br />

Resolution<br />

Range (Å)<br />

R-Factor/Free R-Factor<br />

(%)<br />

Data Completeness<br />

(%)<br />

Page 46<br />

Temperature<br />

(°C) References<br />

1hmv unliganded RT 6.0-3.0 25.4/29.7 85 -165 41<br />

1rtj unliganded RT 25.0-2.35 21.9 89.5 -173 42<br />

1dlo unliganded RT 8.0-2.7 23.0/33.6 99.5 -165 43<br />

1hmi RT/DNA/Fab 15.0-3.0 26.0 88.1 -10 38<br />

3hvt RT/nevirapine 8.0-2.9 26.6 95.6 -165 31,32<br />

1vrt RT/nevirapine 25.0-2.2 18.6 87.1 16 33<br />

1rth RT/1051U91 25.0-2.2 21.4 81.4 16 33<br />

1hni RT/α-APA (R95845) 10.0-2.8 25.5/36 78.5 -15 34<br />

1vru RT/α-APA (R90385) 25.0-2.4 18.7 86.5 16 33<br />

1hnv RT/8-Cl TIBO (R86183) 10.0-3.0 24.9/35.6 81 -10 35<br />

1rev RT/9-Cl TIBO (R82913) 25.0-2.6 22.4 80.7 -173 36<br />

1tvr RT/9-Cl TIBO (R82913) 10.0-3.0 25.9 72 -165 37<br />

1rti RT/HEPT 25.0-3.0 23.6 86.3 14 33<br />

1hrh RT (RNase H domain) 20.0-2.4 20.0 93.1 4 94<br />

1rdh RT (RNase H domain) 20.0-2.8 21.5 95.6 4 95<br />

1har RT (fingers and palm subdomains) 7.0-2.2 20.8/27.0 96 4 44<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_46.html [4/5/2004 4:48:38 PM]


Document<br />

Figure 2<br />

Ribbon diagrams of the three types of HIV-1 RT crystal structures. (a) <strong>Structure</strong> of the HIV-1 RT/DNA/Fab ternary complex<br />

[38]. The bound nucleic acid is shown as double-stranded helices with template strand in black and primer strand in gray. The Fab<br />

fragment is not shown. (b) <strong>Structure</strong> of the HIV-1 RT complexed with the NNRTI TIBO R86183 [34]. For the sake of clarity, the bound<br />

TIBO inhibitor is shown as an atomic model. (c) <strong>Structure</strong> of unliganded HIV-1 RT [41,43]. (d) A schematic diagram showing the<br />

resemblance of the HIV-1 RT p66 subunit to a human right hand. The RNase H domain, which has no counterpart for a human hand, is<br />

shown as an oval below the thumb. When a template-primer binds to HIV-1 RT, the fingers, palm, and thumb subdomains of p66 form a<br />

large cleft to bind the DNA. The polymerase active site (shown as a small circle) lies at the bottom of the DNA-binding cleft. The<br />

NNRTI binds in the highly hydrophobic NNIBP (shown as a large circle), which is located in the vicinity of the polymerase active site.<br />

The p66 thumb subdomain in the NNRTI-bound HIV-1 RT structures is in an upright position extended beyond that observed in the<br />

structure of RT with bound DNA. In the absence of any bound nucleic acid or NNRTI, however, the p66 thumb folds down into the<br />

DNA-binding cleft (shown as dashed drawing).<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_47.html [4/5/2004 4:48:52 PM]<br />

Page 47


Document<br />

Figure<br />

Continued<br />

consequence, the p51 submit has no cleft for binding nucleic acid substrates and hence no polymerase<br />

activity.<br />

Page 48<br />

There is considerable evidence showing that HIV-1 RT is quite flexible and that this flexibility is<br />

essential for DNA polymerization. Comparisons among DNA-bound, inhibitor-bound, and unliganded<br />

HIV-1 RT structures provide a demonstration of the enzyme's flexibility. When a DNA template-primer<br />

binds to HIV-1 RT, structural elements of the fingers, palm, and thumb subdomains of the p66 subunit<br />

form a clamp-like structure that holds the nucleic acid (Figure 2) [38]. The template-primer substrate<br />

interacts with amino acid residues of the fingers, palm, and thumb subdomains, especially in the regions<br />

denoted as “primer grip” and “template grip,” believed to position the template-primer precisely relative<br />

to the polymerase active site [38]. The primary contacts between the template-primer and the protein are<br />

along the sugar-phosphate backbone of the DNA and thus are not sequence-specific [38]. In the absence<br />

of nucleic acid template-primer or NNRTI, the thumb subdomain of p66 is folded down into the DNAbinding<br />

cleft and lies near the fingers subdomain (Figure 2) [40,41,43]. As a consequence, the DNAbinding<br />

cleft is closed. However, even in the absence of a bound nucleic acid, binding of an NNRTI<br />

induces both short-range and long-range structure distortions, including a hinge-like movement near the<br />

base of the p66 thumb that constrains the p66 thumb in a conformation that is extended beyond the<br />

upright<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_48.html [4/5/2004 4:48:59 PM]


Document<br />

Figure 3<br />

Overall structure of HIV-1 RT p66/p51 heterodimer [38] showing the locations<br />

of the major target sites for anti-HIV-1 RT inhibitors. The NRTIs target the dNTP-binding<br />

site/the polymerase active site (shown as a small striped circle) that lies at the<br />

floor of the DNA-binding cleft. The NNRTIs bind to the NNIBP (shown as a large<br />

dotted circle), which is near, but distinct from, the polymerase active site. The RNase H<br />

domain is located at the C-terminal of the p66 subunit. The RNase H catalytic site<br />

(shown as a medium circle) is an attractive target site for anti-HIV-1 drugs. The HIV-1<br />

RT p66/p51 heterodimer interface is shown as a dashed line. Since HIV-1 RT functions<br />

as a heterodimer, any inhibitors that could interfere with the dimerization process might<br />

also be potential drugs for treating HIV-1 infection.<br />

position of the thumb observed in the HIV-1 RT/DNA/Fab structure [31,33–37,43].<br />

In the unliganded structure of HIV-1 RT reported <strong>by</strong> Esnouf et al. [42], the p66 thumb subdomain is in<br />

an upright conformation different from that observed in the other unliganded HIV-1 RT structures but<br />

similar to that found in the DNA-bound HIV-1 RT and NNRTI-bound HIV-1 RT structures. Esnouf et<br />

al. [42] contend that the upright conformation of the p66 thumb subdomain in their unliganded RT<br />

structure is appropriate for unliganded HIV-1 RT and that the binding of an NNRTI does not affect the<br />

conformation of the p66 thumb. However, we believe that the conformation of the p66 thumb in the<br />

Esnouf et al. structure may well be a result of the method used to prepare the<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_49.html (1 of 2) [4/5/2004 4:49:12 PM]<br />

Page 49


Document<br />

Page 50<br />

crystals, which were produced <strong>by</strong> soaking out a weakly bound NNRTI (HEPT) from pregrown crystals.<br />

When the crystals were grown in the presence of HEPT, the p66 thumb was presumably in an upright<br />

position similar to that seen in all of the known NNRTI-bound HIV-1 RT structures. As the weakly<br />

bound HEPT diffused out, the molecular packing arrangement may have constrained the position of the<br />

p66 thumb subdomain. As a consequence, this unliganded HIV-1 RT structure may represent an<br />

intermediate between the other unliganded structures and the structure of HIV-1 RT with a bound<br />

NNRTI. It is evident that the p66 thumb subdomain has considerable flexibility and can adopt<br />

substantially different conformations during the binding of template-primer or inhibitors and,<br />

presumably during DNA polymerization as well.<br />

III. Polymerase Active Site of HIV-1 RT and the NRTIs<br />

Polymerization of DNA <strong>by</strong> HIV-1 RT involves a sequential stepwise binding of the template-primer and<br />

deoxynucleoside triphosphate (dNTP) substrates at the polymerase active site [45,46]. The incoming<br />

dNTP is covalently linked via the α-phosphorus to the 3'-oxygen of the primer strand, accompanied <strong>by</strong><br />

the release of pyrophosphate. An essential requirement for the polymerization reaction is the presence of<br />

a 3'-OH group at the end of the primer strand. Nucleoside analogs contain a modified sugar moiety in<br />

which the 3'-OH group is replaced <strong>by</strong> another group (e.g., hydrogen, halogen, or azido) (Figure la). To<br />

exert their antiviral activity at the level of RT, the NRTIs must be phosphorylated successively to the 5'monophosphate,<br />

5'-diphosphate, and 5'-triphosphate forms <strong>by</strong> a series of kinases. Once the NRTI is<br />

converted to the triphosphate form and interacts with RT, it can inhibit polymerization in two possible<br />

ways. One possibility is that the NRTI binds preferentially to the dNTP-binding site and competitively<br />

inhibits the binding of natural dNTPs. Another possibility, which seems to be the predominant mode of<br />

inhibition, is that the NRTI is incorporated into the growing chain and acts as a terminator of chain<br />

elongation. Once an NRTI is incorporated, no additional nucleotides can be added to the DNA chain<br />

since the primer terminal 3'-OH group (the site of phosphodiester bond formation) is absent.<br />

Phosphorylation is a crucial step in the intracellular metabolism of the NRTI and often is a limiting<br />

factor for the antiviral activity of the NRTI [47]. In an attempt to <strong>by</strong>pass the first phosphorylation step,<br />

several acyclic nucleoside phosphonates have been developed in which the sugar moiety of normal<br />

NRTIs is replaced with an acyclic phosphonate group, such as 9-(2-phosphonylmethoxyethyl)adenine<br />

(PMEA), (R)-9-(2-phosphonylmethoxypropyl)adenine (PMPA), and (S)-9-(3-fluoro-2phosphonylmethoxyethyl)<br />

adenine (FPMPA) (Figure la) (see reviews [5,11]). These acyclic nucleoside<br />

phosphonates are dideoxynucleoside<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_50.html [4/5/2004 4:49:13 PM]


Document<br />

monophosphate analogs which can easily be converted to the active triphosphate form <strong>by</strong> adding two<br />

additional phosphates [48,49], and can inhibit HIV replication [50,52].<br />

Page 51<br />

Analysis of the structure of the HIV-1 RT/DNA/Fab complex showed that the dNTP-binding site is<br />

composed of both protein and nucleic acid. In addition to the 5'-terminus of the template nucleotide and<br />

three carboxylate residues (Asp110, Asp185, and Asp186), the amino acid residues Asp113, Tyr115,<br />

Phe116, Gln151, Phe160, and possibly Met184 and Lys219 form part of the putative dNTP-binding site<br />

[12,53] (Figure 4 and Table 2). It is important to realize that the precise composition, position, and<br />

conformation of the template-primer can influence the recognition and incorporation of incoming<br />

nucleotides at the polymerase catalytic site. In the wild-type HIV-1 RT, the dNTP-binding<br />

Figure 4<br />

Stereoview of the polymerase active site of HIV-1 RT [38]. The amino acid<br />

residues that compose the putative dNTP-binding site, including the three catalytically<br />

essential aspartic acids, are shown with side chains. The double-stranded nucleic acid is<br />

shown with the atomic model in the HIV-1 RT/DNA/Fab complex. The dNTP-binding<br />

site consists of structural elements from both protein and nucleic acid. The precise<br />

composition, position, and conformation of the template-primer can affect the recognition of<br />

incoming dNTPs at the polymerase active site.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_51.html (1 of 2) [4/5/2004 4:49:23 PM]


Document<br />

Table 2 HIV-1 RT Amino Acid Residues Composing the Putative dNTP-Binding Sites and the Locations of NRTI-<br />

Resistance Mutations (see also [12])<br />

dNTP-Binding Site Nucleoside <strong>Drug</strong>-Resistance Mutation Site<br />

Residue Location Mutation Location Possible Effects References<br />

Asp110 β6 Met41Leu αA template binding 55<br />

Asp113 β6-αC loop Ile50Thr β2 unclear 154<br />

Try115 αC Lys65Arg β3–β4 template binding 154<br />

Phe116 αC Asp67Asn β3–β4 template binding 155<br />

Gln151 β8-αE Thr69Asp β3–β4 template binding 156<br />

Phe160 αE Lys70Arg β3–β4 template binding 155<br />

Asp185 β9–β10 Leu74Val β4 template binding 8<br />

Asp186 β9–β10 Val75Thr β4 template binding 163<br />

Lys219 β11 Glu89Gly β5 dsDNA binding 157<br />

Tyr115Phe αC dNTP binding 170<br />

Ile135Thr β7–β8 unclear 158,159<br />

Gln151Met β8-αE dNTP binding<br />

Met184Val, Ile β9–β10 dNTP-binding/fidelity 154,160,161<br />

Thr215Tyr, Phe,<br />

Cys<br />

β11 indirect effect/dNTP<br />

binding<br />

155,162<br />

Lys219Gln β11 dNTP binding 155<br />

Page 52<br />

site can accommodate both the normal dNTP substrates and dideoxynucleoside analogs. The majority of<br />

mutations that confer resistance to NRTIs are not located at the dNTP-binding site; however, they<br />

appear to influence the geometry of the dNTP-binding site indirectly in a way that permits RT to<br />

discriminate between a normal dNTP and a modified nucleoside triphosphate (see discussion in the next<br />

section).<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_52.html (1 of 2) [4/5/2004 4:49:26 PM]


Document<br />

Analysis of various HIV-1 RT structures has revealed an unusual β-turn geometry for the YMDD motif<br />

at the polymerase catalytic site of p66 [33,34,41–43]. The energetically unfavorable main chain<br />

conformation of Met184 (torsion angles φ~60° and ϕ~-120°) is stabilized <strong>by</strong> a hydrogen bond of its<br />

carbonyl oxygen to the side chain of either Gln182 [33,41,43] or Gln161 [42]. It has been suggested that<br />

this novel β-turn geometry might be required to position the aspartic acids in precisely the correct way<br />

for catalysis [41].<br />

IV. Mechanism of NRTI-Resistance Mutations<br />

Development of resistance to NRTIs has been a major problem with clinical use of these drugs. Careful<br />

analysis of mutations that confer resistance to different<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_52.html (2 of 2) [4/5/2004 4:49:26 PM]


Document<br />

Page 53<br />

NRTIs in light of available structural data might provide information that could be used in the<br />

development of improved NRTIs that are more effective against the commonly observed NRTI-resistant<br />

HIV-1 variants. A viable drug-resistant RT mutant should be able to recognize and incorporate normal<br />

nucleoside triphosphate, yet reject a nucleoside analog. The only difference between normal nucleotide<br />

substrates and the NRTIs is the modification of the sugar moiety. This alteration may affect sugar<br />

puckering and the conformation of the glycosyl bond. Recognition of these differences could render the<br />

triphosphate form of the nucleoside analog a good substrate for wild-type RT but a poor substrate for a<br />

drug-resistant variant of RT.<br />

Structural analysis of HIV-1 RT has shown that most of the NRTI-resistance mutations are not located<br />

close to the putative dNTP-binding site and are unlikely to have a direct impact on the binding of dNTP<br />

analogs (Figure 5 and<br />

Figure 5<br />

A close-up view showing the relative locations of the commonly identified<br />

drug-resistance mutations for NRTIs (in dark-gray) and for NNRTIs (in light-gray) with<br />

respect to the bound DNA. Most of the NRTI-resistance mutations are not located at the<br />

putative dNTP-binding site, but are at positions to have potential interactions with the<br />

nucleic acid template-primer. Conversely, all the NNRTI-resistance mutations are<br />

clustered around the NNIBP and have direct contacts with NNRTIs or have direct effect on<br />

inhibitor binding.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_53.html (1 of 2) [4/5/2004 4:49:46 PM]


Document<br />

Page 54<br />

Table 2) [12,54]. For example, none of the five mutations, in HIV-1 RT Met41Leu, Asp67Asn,<br />

Lys70Arg, Thr215Tyr, and Lys219Gln [55] (Table 2) consistently associated with resistance to AZT,<br />

are at locations close to the dNTP-binding site. However, most (but not all, c.f. [56]) biochemical studies<br />

have failed to show that recombinant HIV-1 RT enzymes containing these mutations are more resistant<br />

to inhibition <strong>by</strong> AZT triphosphate than the wild-type HIV-1 RT [27,57,58]. Other mutations that confer<br />

resistance to NRTIs have been identified at positions 50, 65, 69, 74, 75, 89, 115, 135, 151, and 184 of<br />

HIV-1 RT (Figure 5 and Table 2). Most of these mutations do not lie close to the dNTP-binding site<br />

(Met184Val/Ile are the exception), but instead are located at positions where they could interact with the<br />

nucleic acid template-primer [12,59,60]. Biochemical data have shown that only when the 5'-template<br />

extension length is greater than three nucleotides does the wild-type RT begin to incorporate<br />

dideoxynucleotides effectively [54]. If the template extension is less than three nucleotides in length,<br />

wild-type HIV-1 RT is resistant to dideoxynucleotides. On the other hand, HIV-1 RT variants containing<br />

the mutations Leu74Val or Glu89Gly did not readily incorporate dideoxynucleotides either with short or<br />

long template extensions [54]. Based on both structural and biochemical data, it was proposed that<br />

mutations that cause HIV-1 RT to have reduced sensitivity to NRTIs exert their effects via interactions<br />

with the nucleic acid template-primer, which consequently alter the geometry of the polymerase active<br />

site [54]. It has been suggested that mutations that confer resistance to foscarnet might use a similar<br />

mechanism [61]. One possible exception to this mechanism might be the mutations of Met184Val and<br />

Met184Ile (see review [27]). Part of the highly conserved YMDD motif, Met184 is adjacent to residues<br />

Asp185 and Asp186, which are two of the three catalytically essential aspartic acid residues at the<br />

polymerase active site. In addition, Met184 appears to interact with the ribose moiety of the 3'-terminal<br />

nucleotide of the primer strand [12,38,53] (Figure 4). Therefore, mutations at this position could affect<br />

interactions with the incoming dNTP directly and/or alter the positioning of the nucleic acid. These<br />

mechanisms are not mutually exclusive and which mechanism is responsible for resistance has not yet<br />

been resolved [62]. There are two recent reports suggesting that the Met184Val mutant HIV-1 RT has<br />

approximately three-fold higher fidelity than the wild-type enzyme [63,64]. Based on these data, it was<br />

suggested that this increase in fidelity might reduce the overall rate of generation of viral variants in<br />

patients treated with 3TC or other dideoxynucleosides [64]. However, owing to both theoretical and<br />

technical problems with these analyses, these conclusions are controversial. Determination of crystal<br />

structures of both wild-type and mutant HIV-1 RT complexed with individual NRTIs in the presence of<br />

a variety of template-primers and/or dNTP substrates should provide a better understanding of the<br />

mechanisms of dNTP selection and drug resistance.ed at Arial for catalysis [41].<br />

IV. Mechanism of NRTI-Resistance Mutations<br />

Development of resistance to NRTIs has been<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_54.html [4/5/2004 4:49:47 PM]


Document<br />

V. <strong>Drug</strong> <strong>Design</strong> Targeting at the Polymerase Active Site<br />

Page 55<br />

All of the existing NRTIs contain a modified sugar moiety that lacks the 3'-OH group that is essential<br />

for incorporation of the next nucleotide. Modification can also be made on other functional groups such<br />

as the base and the triphosphate moieties. It may be worthwhile to try to alter the base moiety of the<br />

NRTIs to produce compounds that will be more specific to HIV-1 RT (i.e., less cytotoxic to normal<br />

cellular polymerases) and more effective against both wild-type or drug-resistant viral variants.<br />

<strong>Structure</strong>-activity analysis indicates that the pyrimidine moiety of the NRTIs can be modified at the C5<br />

position. An AZT derivative that has a 3'-azido group on the sugar moiety and a methyl group at the C5<br />

position of the pyrimidine moiety showed potent antiviral activity [65]. Substitution of the methyl group<br />

with a chlorine atom at position C5 of AZT results in a compound that has strong anti-HIV-1 activity<br />

[66]. Other possible substitutions at the C5 position include other halogen atoms or an ethyl group.<br />

Another possible drug-design strategy would be to devise compounds that can interface with the binding<br />

of the metal ions (Mg 2+ or Mn 2+) at the polymerase active site. Metal ions appear to be important in<br />

DNA polymerase catalysis. Based on the structural and biochemical data, a two-metal dependent<br />

mechanism of polymerization has been postulated [53,67,68] that is similar to that proposed for other<br />

DNA polymerases [69–71]. In this model, the metal ions mediate interactions between the three<br />

catalytically essential aspartic acid residues (Asp100, Asp185, and Asp186) and the α-, β-, and γphosphates<br />

of the incoming dNTP and promote the nucleophilic attack on the α-phosphate <strong>by</strong> the<br />

oxygen atom of the 3'-OH group of the primer strand. In the structure of the fingers and palm<br />

subdomains of the RT of Moloney murine leukemia virus (MuLV), a single Mn 2+ ion was found bound<br />

to the two aspartic acid residues at the polymerase active site [72]. In the structure of the unliganded<br />

HIV-1 RT, an electron density peak was located at the polymerase active site with a good coordination<br />

geometry to the Oδ1 atoms of both Asp185 and Asp186 [43]. This electron-density peak is in a position<br />

similar to that of the Mn 2+ ion observed in the MuLV RT structure. It is possible that this position<br />

corresponds to a Mg 2+ ion-binding site [43]. It might be useful to design inhibitors that would influence<br />

the metal-ion coordination using either computer-<strong>based</strong> calculations (such as DOCK [73–75]) or <strong>based</strong><br />

directly on an analysis of HIV-1 RT structure. Crystal structures of HIV-1 RT complexed with<br />

Mg 2+/Mn 2+ ion(s) at the polymerase catalytic site in the presence of template-primer and/or dNTP<br />

substrates would be helpful in defining the target sites of inhibitors of this type. Further studies on the<br />

structure activity relationship of HIV-1 RT complexes with these inhibitors, if active, might ultimately<br />

lead to a new type of HIV-1 RT drug that would not compete with the dNTP binding but would affect<br />

the DNA polymerization mechanism.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_55.html [4/5/2004 4:49:49 PM]


Document<br />

Page 56<br />

It is also attractive to consider developing agents that bind to the HIV-1 RT polymerase active site but<br />

are not nucleoside analogs. Since the amino acid residues at the dNTP-binding site are highly conserved,<br />

viral variants resistant to such inhibitors may be significantly impaired in their polymerase activity.<br />

However, there is a good chance that drug resistance could result from mutations in HIV-1 RT that<br />

influence the precise positioning of template-primer [54]. Initial attempts to use this approach starting<br />

from the crystal structure of the HIV-1 RT/DNA/Fab complex [38] (Ding, et al., in preparation) have<br />

uncovered some interesting lead compounds (Kuntz, Kenyon, Arnold, Hughes, et al., unpublished).<br />

VI. NNRTIs and the NNIBP<br />

Nonnucleoside RT inhibitors (NNRTIs) constitute the other major class of HIV-1 RT inhibitors. Many<br />

structurally distinct families of NNRTIs have been identified, including HEPT [13], TIBO [14],<br />

nevirapine [15], BHAP [17], TBA [18,19], TSAO [76], α-APA [21], pyridinones [16] and quinoxalines<br />

(HBY) [22,23] (Figure 1b). However, development of drug resistance is a major problem when NNRTIs<br />

are used to treat AIDS patients. An ideal drug should be able to block replication of all viable strains of<br />

HIV-1, but should not inhibit normal cellular enzymes. In this regard, the known NNRTIs may be too<br />

specific. While these inhibitors do not inhibit cellular polymerases, they are also inactive against HIV-2<br />

RT (which can be viewed as an extreme variant of HIV-1 RT). In addition, drug-resistant variants of<br />

HIV-1 RT emerge rapidly in the presence of most inhibitors. In contrast, the NRTIs inhibit a broad<br />

spectrum of polymerases including the host cellular polymerases. Though it appears to be more difficult<br />

for the virus to evade NRTIs than NNRTIs (in general, it takes longer for the virus to develop resistance<br />

to NRTIs than NNRTIs), NRTI toxicity is a serious problem.<br />

Structural and biochemical studies have shown that all NNRTIs bind in a highly hydrophobic pocket in<br />

the p66 subunit, located approximately 10 Å away from the polymerase active site (Figures 2 and 3)<br />

[31,33–37]. Nevertheless, in all known structures of HIV-1 RT/NNRTI complexes, the bound NNRTIs<br />

have not been found to have any direct interactions with residues that compose the putative dNTPbinding<br />

site. The nonnucleoside inhibitor binding pocket (NNIBP) contains primarily amino acid<br />

residues from the β5–β6 loop (Pro95, Leu100, Lys101, and Lys103), β6 (Val106 and Val108), the<br />

β9–β10 hairpin (Val179, Tyr181, Tyr188, and Gly190), and the β12–β13 hairpin (Phe227, Trp229,<br />

Leu234, His235, and Pro236) of the p66 palm subdomain, and β15 (Tyr318) of the p66 thumb<br />

subdomain, as well as the β7–β8 connecting loop (Glu138) of the p51 fingers subdomain (Figure 6 and<br />

Table 3).<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_56.html [4/5/2004 4:49:51 PM]


Document<br />

Figure 6<br />

Superposition of the NNRTIs in the HIV-1 RT/TIBO complex [35], HIV-1 RT/α-APA complex [34], and HIV-1<br />

RT/nevirapine complex [32]. The side chains are shown for those amino acid residues that have close contacts with bound inhibitors<br />

and the three catalytically essential aspartic acids in the HIV-1 RT/TIBO complex. Most of the amino acid residues that form the<br />

NNIBP are hydrophobic. Though the NNRTIs are chemically and structurally diverse, the bound NNRTIs all assume a strikingly<br />

common butterfly-like shape.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_57.html [4/5/2004 4:50:13 PM]<br />

Page 57


Document<br />

Table 3 HIV-1 RT Amino Acid Residues Defining the Nonnucleoside Inhibitor-Binding Pocket (NNIBP) and the<br />

Locations of NNRTI-Resistance Mutations (see also [12])<br />

NNIBP<br />

Residues<br />

Pro95 β5<br />

Location Mutation Possible Effects References<br />

Ala98 β5–β6 Gly less bulky 9,164<br />

Leu100 β5–β6 Ile β-branch 9,164-166<br />

Lys101 β5–β6 Glu charge change 164<br />

Lys103 β5–β6 Asn, Gln charge loss, less bulky 9,24,164<br />

Val106 β6 Ala less bulky 9,58,165<br />

Val108 β6 Ile bulkier 9,164<br />

Glu138 β7–β8(p51) Lys charge change 166<br />

Val179 β9 Glu, Asp charge gain, bulkier 164<br />

Tyr181 β9 Cys, Ile aromaticity loss, less bulky 24,58,164<br />

Tyr188 β10 His, Cys, Leu aromaticity loss, less bulky 9,167<br />

Gly190 β10 Glu charge gain, bulkier 9,22,168<br />

Phe227 β12<br />

Leu228 β12 Phe aromaticity gain, bulkier 133<br />

Trp229 β12<br />

Glu233 β13 Val charge loss, less bulky 169<br />

Leu234 β13<br />

Pro236 β13–β14 Leu increase flexibility, bulkier 133<br />

Lys238 β14 Thr charge loss, less bulky 133<br />

Tyr318 β15<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_58.html (1 of 2) [4/5/2004 4:50:15 PM]<br />

Page 58


Document<br />

Most of the amino acid residues that form the binding pocket are hydrophobic and five of them are<br />

aromatic residues. The hydrophobic interactions of the side chains of these residues, especially Tyr181,<br />

Tyr188, and Trp229, with the hydrophobic moieties of the NNRTIs appear to be important for inhibitor<br />

binding [32–34,59]. Since most of the NNRTIs also contain polar group(s), they have the potential to<br />

form hydrogen bonds with surrounding amino acid residues either directly or via water bridges [33–36].<br />

In the structures of both liganded HIV-1 RT and the HIV-1 RT/DNA/Fab complex, there are two small<br />

surface depressions at the base of the NNIBP that are the putative entrances to the pocket [34,43]. One<br />

surface depression is located at the p66/p51 heterodimer interface and is composed of amino acid<br />

residues Leu100, Lys101, Lys103, Val179, Tyr181, and Tyr188 of p66, and Glu138 of p51 [34]. This<br />

putative entrance is narrow compared to the size of the NNRTIs. Another surface depression has been<br />

found at the location near the base of the p66 thumb subdomain between two adjacent structural<br />

elements: the β5–β6 connecting loop (Lys101 and Lys103) and the β13–β14 hairpin (Pro236 and<br />

Leu238) [43]. Since this site is also exposed to solvent, an NNRTI could approach the NNIBP from it.<br />

How-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_58.html (2 of 2) [4/5/2004 4:50:15 PM]


Document<br />

ever, once an NNRTI is bound to RT, only the first putative entrance remains accessible; the second<br />

disappears due to the conformational change and repositioning of the β12-β13-β14 sheet [43].<br />

Page 59<br />

It is evident, from comparison of the various HIV-1 RT structures that the NNIBP has a highly flexible<br />

structure that apparently allows the enzyme to accommodate various types of NNRTIs with different<br />

shapes and sizes. Despite apparent differences in the structures of the bound inhibitors, comparison of<br />

structures of several HIV-1 RT/NNRTI complexes revealed remarkable similarity in the geometry of<br />

both the bound inhibitors and the NNIBP [33,35]. All these chemically diverse NNRTIs assume a<br />

strikingly similar butterfly-like shape (Figure 6). The binding of NNRTIs in the NNIBP can be likened<br />

to a butterfly sitting on the β6-β10-β9 sheet and facing toward the putative entrance to the pocket. The<br />

angle between the two wings of the “butterfly” is approximately 112–115° in the TIBO, α-APA, and<br />

nevirapine complexes [35]. This angle might be critical in inhibitor binding and could be a crucial<br />

parameter in the design of new NNRTIs. There are many other NNRTIs that are significantly larger or<br />

smaller in size than α-APA, TIBO, or nevirapine. It is very likely that the NNIBP can adopt other<br />

conformations. For example, BHAP appears to be too large to fit into the NNIBP in any of the reported<br />

HIV-1 RT/NNRTI complexes. The NNIBP in the HIV-1 RT/BHAP complex would need to be<br />

significantly larger than that observed in the structures of the known HIV-1 RT/NNRTI complexes. It is<br />

possible that the BHAP inhibitor may not conform to a butterfly-like shape. This underscores the<br />

importance of solving crystal structures for as many HIV-1 RT/NNRTI complexes as possible.<br />

Additional structural and biochemical data for other HIV-1 RT/NNRTI complexes should provide the<br />

insight needed to define the limits of the flexibility of HIV-1 RT in the NNIBP region.<br />

VII. Process of NNRTI Binding<br />

In crystal structures of unliganded HIV-1 RT [40,41,43] and of HIV-1 RT/DNA/Fab complex [38], the<br />

NNIBP does not exist (although a small cavity is found in the region of the NNIBP proximal to the<br />

polymerase active site in the unliganded HIV-1 RT structure described <strong>by</strong> Esnouf et al. [42]). In these<br />

structures, the side chains of Tyr181 and Tyr188 in p66 point away from the polymerase active site and<br />

toward the hydrophobic core. However, in the HIV-1 RT/NNRTI complex structures, the side chains of<br />

Tyr181 and Tyr188 point toward the polymerase active site, and the side chain of Tyr181 is in a position<br />

that prevents Trp229 from occupying the position it has in the unliganded or DNA bound HIV-1 RT<br />

structures. Binding an NNRTI also moves the β12-β13-β14 sheet away from the hydrophobic core<br />

[34,35,37]. These conformational<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_59.html [4/5/2004 4:50:16 PM]


Document<br />

changes create the space in the pocket required to accommodate inhibitors. In other words, significant<br />

conformational changes occur during the process of inhibitor binding that lead to the formation of the<br />

NNIBP [33–35]. This observation also underscores the importance of determining structures of HIV-1<br />

RT with and without bound inhibitors.<br />

Page 60<br />

An obvious question is, what forces initiate this series of conformational changes during NNRTI<br />

binding? One possibility is the contacts between the inhibitor and the protein. Though the NNIBP is<br />

hydrophobic, there are three hydrophilic amino acid residues (Lys101 and Lys103 of p66, and Glu138 of<br />

p51) at the rim of the putative entrance(s) to the pocket. The flexible and polar side chains of these<br />

residues could assist in steering an inhibitor into the pocket and/or could block the bound inhibitor from<br />

escaping out of the pocket. Mutagenesis studies have shown that these three residues are important in<br />

the binding of NNRTIs. Though the importance could be explained in terms of the interactions between<br />

these residues and the bound inhibitor in the final complexes, interactions at the initial stages of inhibitor<br />

binding might also be crucial. The flexible and polar side chains of these residues might help in directing<br />

the inhibitor toward the entrance to the pocket via electrostatic interactions, in part <strong>by</strong> replacing the<br />

original hydrogen bonds between the drug and the solvent molecules. Any initial energy gains from such<br />

polar interactions could potentially be replaced <strong>by</strong> hydrogen bonds or other types of interactions<br />

between the inhibitor and alternative residues as the inhibitor moves deeper into the binding pocket. In<br />

addition, significant portions of the aromatic rings of both Tyr181 and Tyr188 are exposed at the bottom<br />

of the surface depression and offer the potential for early π-π interactions with the inhibitor. This type of<br />

π-π interaction might also play an important role in the initial approach of inhibitors to the binding<br />

pocket. This hypothesis may provide a kinetic explanation for the ineffectiveness of NNRTIs against<br />

viral strains of HIV-1 that carry nonaromatic amino acids at positions 181 and 188. As the solvated<br />

inhibitor approaches the enzyme and proceeds to enter the binding pocket, most of the water molecules<br />

of solvation are lost. The few water molecules that remain in the NNRTI-bound complex are typically<br />

located at the entrance to the pocket, forming water bridges between the inhibitor and one or two polar<br />

residues around the entrance [33,35,36]. Once the inhibitor is in place, the surface residues close down<br />

around the drug preventing it from escaping <strong>by</strong> effectively sealing the entrance to the pocket.<br />

VIII. Mechanisms of Inhibition <strong>by</strong> NNRTIS<br />

Based on structural, biochemical, and genetic data several hypotheses have been postulated about the<br />

mechanism(s) of inhibition of HIV-1 RT <strong>by</strong> NNRTIs. It is<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_60.html [4/5/2004 4:50:18 PM]


Document<br />

Page 61<br />

now clear that the binding of NNRTIs provokes substantial conformational changes in both secondary<br />

structural elements and in side chains of residues in the NNIBP. These conformational changes in the<br />

NNIBP could directly or indirectly affect the precise geometry and/or mobility of the near<strong>by</strong> polymerase<br />

catalytic site, especially the highly conserved YMDD motif and/or the divalent metal ions<br />

[31,33,34,42,68]. The binding of NNRTIs appears to lock the flexible hinge-like structure between the<br />

palm and thumb subdomains and restrict mobility of the thumb subdomain, placing constraints on the<br />

geometry of the DNA-binding cleft [12,31,34,43]. The “primer grip” (i.e., β12-β13-β14 sheet), which<br />

has close interactions with the 3'-terminus of the primer strand [38], forms a part of the NNIBP and is<br />

involved in the binding of NNRTIs. It has become apparent that binding of NNRTIs can substantially<br />

alter the conformation of the primer grip; this could affect the precise positioning of the primer strand<br />

relative to the polymerase active site [34,37]. Displacement of the primer grip <strong>by</strong> NNRTI binding could<br />

lead to repositioning of the primer terminus. This could explain the observation that dNTP binding is<br />

largely unaffected <strong>by</strong> NNRTI binding while the rate of the chemical step of DNA polymerization is<br />

reduced [77]. Long-range distortions of the HIV-1 RT structure <strong>by</strong> NNRTI binding can potentially<br />

account for NNRTI inhibition of polymerization [39,41,43] and alteration of RNase H cleavage<br />

specificity [43,78]. These possible mechanisms are not mutually exclusive and the binding of inhibitors<br />

might have multiple influences on HIV-1 RT polymerization. The exact mechanism(s) of inhibition is<br />

still under investigation.<br />

IX. NNRTI-Resistance Mutations<br />

Analyses of the crystal structures of HIV-1 RT complexed with various NNRTIs have indicated that<br />

amino acid residues whose mutations confer high levels of resistance to NNRTIs [9,11,12,26,27] are<br />

located close to the bound inhibitors (Figure 5 and Table 2). Subunit-specific mutagenesis studies have<br />

confirmed that mutations that confer resistance to the NNRTIs act directly through the change in the<br />

NNIBP itself [60,79]. In these studies, recombinant HIV-1 RTs that contained amino acid substitutions<br />

only in the p66 subunit were resistant to NNRTIs, while those containing the same amino acid<br />

substitutions only in the p51 subunit remained susceptible to the drugs. There is one exception: the<br />

amino acid substitution of Glu138 to Lys, which confers resistance to inhibitors only when it is present<br />

in the p51 subunit. Amino acid residue 138 is located in the β7–β8 connecting loop of the fingers<br />

subdomain. In the p51 subunit this residue forms a part of the NNIBP, while its counterpart in the p66<br />

subunit is far away from the pocket [12,60].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_61.html [4/5/2004 4:50:22 PM]


Document<br />

Page 62<br />

The mechanism(s) of resistance may depend on the specific amino acid change. It is likely that most<br />

NNRTI-resistance mutations exert their effects <strong>by</strong> altering interactions between protein side chains and<br />

the inhibitors [12,34,35]. <strong>Drug</strong>-resistance mutations that result in a decrease or increase in the size of<br />

side chains might lead to loss of favorable contacts or steric conflicts with bound inhibitors. Mutations<br />

that alter the local electrostatic potential, i.e., gain, loss, or inversion of charge, may change the affinity<br />

of the NNIBP for inhibitor binding. These altered interactions could interfere with the binding of<br />

NNRTIs to the hydrophobic pocket or conceivably could even relax the geometric distortion that the<br />

binding of an inhibitor causes in the vicinity of the polymerase active site.<br />

X. <strong>Design</strong> of Improved NNRTIs<br />

Different NNRTIs, even from the same class of compounds, show remarkable variations in their ability<br />

to inhibit HIV-1 replication and can give rise to different spectra of resistance mutations [9,11,26,27].<br />

For example, biochemical studies showed that the 8-chloro TIBO derivative R86183 is quite potent in<br />

inhibiting an HIV-1 strain containing the Tyr181Cys mutation, which is one of the frequently occurring<br />

HIV-1 RT mutations that gives rise to a high level of resistance to almost all NNRTIs, including other<br />

TIBO derivatives [80]. There are several other reports of NNRTIs that are also relatively effective in<br />

inhibiting the HIV-1 RT Tyr181Cys variant [81–84]. These results suggest that although all the<br />

inhibitors appear to bind in the NNIBP, there are differences in their specific interactions with HIV-1<br />

RT. Structural analyses of HIV-1 RT/NNRTI complexes and computer modeling studies confirmed that<br />

the exact conformations of the amino acid residues forming the NNIBP appear to vary in different<br />

complexes and that there are specific interactions between individual inhibitor and surrounding residues<br />

[33,35,36,85]. However, these differences have not been sufficiently large to allow a successful<br />

combination therapy to be developed using two or more of the currently available NNRTIs (discussed in<br />

more detailed in a later section) [9,11,26,27,86]. Systematic analysis of wild-type and drug-resistant<br />

mutant HIV-1 RT structures in complexes with various NNRTIs should provide additional insights<br />

about constraints that could be used to optimize the design of NNRTIs. This knowledge could guide<br />

development of more effective inhibitors for AIDs treatment.<br />

As discussed earlier, the bound NNRTIs in HIV-1 RT complexes determined so far conform to a<br />

common butterfly-like shape (Figure 6). A close inspection of interactions between inhibitors and<br />

protein reveals that though<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_62.html [4/5/2004 4:50:38 PM]


Document<br />

Page 63<br />

most of the amino acid residues forming the pocket adjust their side chains to make close contacts with<br />

the inhibitor, the inhibitor is not sufficient to fill all of the space in the pocket. There is space for<br />

additional nonpolar, polar, or charged groups. Modification of the inhibitor would result in adjustment<br />

of the orientation of the side chains and could improve interactions between the inhibitor and<br />

surrounding residues such as Leu100, Lys101, Lys103, Val106, or Leu234. Inhibitors designed to have<br />

more extensive interactions with essential elements in the pocket should minimize the chances of<br />

selecting resistant HIV-1 RT variants. From this point of view, NNRTIs that interact with the relatively<br />

conserved residues of the pocket, such as Trp229, Leu234, and Tyr318, may reduce the risk of<br />

encountering resistance mutations that do not have significant costs for the enzyme. In addition,<br />

compounds could be designed to contain functional groups (for example charged or polar groups) able<br />

to fill more of the available space of the NNIBP and also capable of specific hydrophilic interactions<br />

with the polar or charged side chains and/or with polypeptide backbone atoms of the NNIBP (for<br />

example the main chain amide nitrogens and carbonyl oxygens). The hydrophilic interactions between<br />

inhibitors and protein backbone atoms should be advantageous because mutations to any amino acid<br />

other than proline would not affect such contacts. In the structures of HIV-1 RT/NNRTI complexes, the<br />

bound inhibitors are located very close to the polymerase active site composed of the three catalytically<br />

essential aspartic acids Asp110, Asp185, and Asp186. It might be useful to design compounds that have<br />

a long and branched aliphatic group or a substituted aromatic group that could not only produce<br />

hydrophobic interactions with Tyr181, Tyr188, and Trp229, but could also be able to interact with the<br />

three aspartic residues or interfere with the metal ion(s) binding at the polymerase active site.<br />

XI. RNase/H-Active Site as a Potential <strong>Drug</strong> Target Site<br />

HIV-1 RT contains RNase H, which is responsible for degradation of viral RNA and removal of RNA<br />

primers for minus- and plus-strand DNA synthesis (see reviews [87–89]). The absolute requirement for<br />

virus-associated RNase H function [90–93] offers an additional target for antiretroviral drugs. The<br />

RNase H domain of HIV-1 RT is located at the C-terminus of the p66 subunit (Figures 2 and 3). In<br />

contrast to the polymerase domain of HIV-1 RT, the structure of the RNase H domain is quite similar in<br />

all known HIV-1 RT structures and conforms quite well with the structure of the isolated HIV-1 RNase<br />

H domain [94–95]. The relative stability of the structure of the RNase H domain suggests that the RNase<br />

H active site could be a relatively well-defined target for drug<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_63.html [4/5/2004 4:50:40 PM]


Document<br />

Page 64<br />

design. Mutagenesis studies have demonstrated the interdependence of DNA polymerase and RNase H<br />

activities. Mutations that disrupt one of the two enzymatic activities of HIV-1 RT often also impair the<br />

second activity [96–99]. Indeed, according to the crystal structure of HIV-1 RT, the polymerase active<br />

site and the RNase H active site are separated <strong>by</strong> approximately 17–18 nucleotides [38] and the RNase<br />

H domain has many contacts with the polymerase domain, especially with the connection subdomain of<br />

p66 and the thumb and connection subdomains of p51 [31,38,100]. Interactions between the polymerase<br />

domain and nucleic acid can modulate RNase H activity. Because the predominant contacts of HIV-1<br />

RT with template-primer occur in the vicinity of the polymerase active site, precise placement of the<br />

template strand relative to the RNase H active site may be regulated <strong>by</strong> the sequence and composition of<br />

the template-primer. Mutagenesis experiments showed that mutations located at or near the “template<br />

grip” in the polymerase domain of HIV-1 RT can have a greater effect on RNase H than on polymerase<br />

activity [99,101,102]. It was also reported that binding of the NNRTI nevirapine alters the cleavage<br />

specificity of RNase H [78]. Structural distortions in the position and conformation of template-primer<br />

induced <strong>by</strong> NNRTI-binding may account for alteration of the cleavage specificity of RNase H [43].<br />

Divalent metal ions such as Mg 2+ or Mn 2+ are essential for the RNase H activity [103–106]. The<br />

structure of the isolated RNase H domain crystallized in the presence of MnCl 2 revealed two tightly<br />

bound Mn 2+ ions in close proximity to four catalytically essential acidic residues, Asp443, Glu478,<br />

Asp498, and Asp549, that form the active site [94]. Biochemical data have shown that mutations of<br />

these conserved residues could either disrupt RNase H activity or lead to a highly unstable enzyme<br />

[107–109]. Based on the crystal structures, a two-metal ion-dependent catalytic mechanism for RNase H<br />

activity has been postulated [101], which is similar to that proposed for phosphoryl transfer reactions<br />

catalyzed <strong>by</strong> polymerases and their associated nucleases [67,69–71]. In contrast, in the structure E. coli<br />

RNase H reported <strong>by</strong> Katayanagi et al. [111] only one Mg 2+ ion was observed, and that led to the<br />

proposal of a single metal-ion catalyzed hydrolysis [112]. Interestingly, in the structure of unliganded<br />

HIV-1 RT reported <strong>by</strong> Rodgers et al. [41] and Hsiou et al. [43] only one Mg 2+ ion was found at the<br />

RNase H active site. The mechanism of RNase H cleavage and the exact role of metal ion(s) in the<br />

hydrolysis and formation of phosphodiester bonds are still under investigation (see review [89]).<br />

Very few inhibitors specifically target HIV-1 RNase H activity. Illimaquinone, a natural marine product,<br />

was shown to preferentially inhibit the HIV-1 RNase H activity [113,114]. However, this compound<br />

appears to react with a sulfhydryl group in the polymerase domain and not with RNase H itself. It may<br />

be possible to use the available information on structural and biochemi-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_64.html [4/5/2004 4:50:41 PM]


Document<br />

Page 65<br />

cal properties of the polymerase and RNase H of HIV-1 RT to design compounds that would bind at the<br />

RNase H active site or interfere with the metal ion(s) binding and inhibit the RNase H activity of HIV-1<br />

RT. However, for optimal utility, these compounds should selectively inhibit the RNase H activity of<br />

HIV-1 and not the RNase H activity of the host cells.<br />

XII. Other Possible Target Sites for HIV-1 RT Inhibitors<br />

Both polymerase and RNase H activities of HIV-1 RT require that the enzyme be in a dimeric form<br />

[115–118] (Figure 3). The exact role(s) of the p51 subunit in the enzymatic activities of HIV-1 RT are<br />

not yet known. The three-dimensional structure of HIV-1 RT shows that the interface between p66 and<br />

p51 primarily involves interactions between the p66 palm and the p51 fingers subdomains, between the<br />

p66 connection and the p51 connection and fingers subdomains, and between the RNase H and the p51<br />

thumb and connection subdomains [34,100,119] (Figure 3). A compound that would interfere with<br />

dimerization would be a potential candidate for an anti-AIDS drug.<br />

As discussed earlier, the flexibility of HIV-1 RT permits the enzyme to adopt different conformations.<br />

In the absence of bound DNA, the thumb and the fingers subdomains come together and close a major<br />

portion of the DNA-binding cleft [40,41,43]. Synthetic oligonucleotides that could interact with the<br />

specific or conserved regions of the DNA-binding cleft could potentially block binding of templateprimer<br />

substrates. An RNA pseudoknot has been reported to bind and specifically inhibit HIV-1 RT<br />

[120]. Chemical modification and substitution of specific groups in RNA ligands can change the<br />

structure of the pseudoknot, which could result in considerably more effective pseudoknot inhibitors<br />

with high binding specificity [121]. Studies employing the phosphorodithioate analogs of the primer<br />

sequence recognized <strong>by</strong> HIV-1 RT showed that these compounds can act as inhibitors and that inhibition<br />

is a function of both the sequence and length of these novel single-stranded nucleic acid oligomers<br />

[122,123].<br />

A series of natural products, i.e., trihydroxyquinolone compounds isolated from Red Sea marine<br />

organisms, were reported to inhibit the DNA polymerase activity of HIV-1 RT [124,125]. This type of<br />

inhibitor appears to have a mechanism of inhibition that is different from either the NRTI inhibition<br />

mechanism or the NNRTI inhibition mechanism. The inhibition is reversible and noncompetitive with<br />

respect to both dNTP and template-primer [125]. This result indicates that there are other potential<br />

binding sites for inhibitors of HIV-1 RT.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_65.html [4/5/2004 4:50:43 PM]


Document<br />

XIII. Useful Tools in <strong>Structure</strong>-Based <strong>Drug</strong> <strong>Design</strong><br />

Several computer modeling algorithms have been developed for structure-<strong>based</strong> drug design. Among<br />

them, DOCK [73–75] and 3D SEARCH [126] have been successfully applied in the design of HIV-1<br />

protease inhibitors. These programs search a target protein for invaginations, grooves, and recognition<br />

surfaces that could bind a potential receptor molecule. Compounds complementary to the putative<br />

receptor binding site in both shape and chemical properties can be identified through searching<br />

databases of small molecules, such as the Cambridge Crystallographic Database, the Fine Chemicals<br />

Directory, or other commercially available databases.<br />

Page 66<br />

An important issue in analyzing HIV-1 RT is the flexibility of the enzyme. Comparison of structures of<br />

unliganded HIV-1 RT and NNRTI-bound HIV-1 RT complexes has shown that the NNIBP is not<br />

present in the unliganded form [34,41,43]. This underscores the importance of searching both the<br />

unliganded HIV-1 RT and the HIV-1 RT complexes with inhibitors and substrates in order to identify<br />

any potential inhibitor-binding sites.<br />

Many other approaches have been and are being developed for computeraided design of inhibitors. For<br />

example, pharmacophore analysis can identify the spatial arrangement of groups or atoms common to all<br />

active inhibitor molecules and then incorporate these elements into a single molecule [127,128].<br />

Detailed analysis of the volumes occupied <strong>by</strong> different inhibitors bound to the same binding site could<br />

also provide new suggestions for inhibitor design. For example, the volume union of all known NNRTIs<br />

such as nevirapine, TIBO, α-APA, HEPT, and 1051U91 can be calculated. This type of analysis could<br />

be used to screen for new NNRTIs. Since the coordinates for a number of HIV-1 RT/NNRTI complex<br />

structures are now available in the Protein Data Bank, these approaches can be applied to the design of<br />

new or improved NNRTIs. Given the relatively high flexibility of the NNIBP region and the diversity of<br />

NNRTI structures, the NNIBP of HIV-1 RT could be a methodologically challenging yet extremely<br />

important target for structure-<strong>based</strong> drug design.<br />

XIV. Enzymatic Efficiency of <strong>Drug</strong>-Resistant HIV-1 RT Variants<br />

Analyses of viral population dynamics indicated that, although drug resistance cannot be seen as a<br />

positive outcome of chemotherapy, clinical progress can be made through the development of drugresistant<br />

viral variants (see review [30]).alyzed hy Arial H itself. It may be possible to use the available<br />

information on structural and biochemi-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_66.html [4/5/2004 4:50:44 PM]


Document<br />

Page 67<br />

Biochemical data show that HIV-1 replicates extremely rapidly in infected individuals and that the viral<br />

load is low in the early stages of the disease because the host immune system is initially successful in<br />

limiting viral replication [28,29]. When patients are treated with either RT and/or protease inhibitors,<br />

wild-type HIV-1 is rapidly replaced with drug-resistant variants. In fact, even in patients who have not<br />

received treatment with any anti-RT drugs, HIV-1 variants that contain residues corresponding to both<br />

NRTI- and NNRTI-resistance mutations in RT can be found as minor components of the viral<br />

population [129]. Similarly, viral variants that contain residues in protease sequence corresponding to<br />

protease inhibitor drug-resistance mutations have also been observed in patients prior to drug therapy<br />

(see review [130]). Enzymatic components found in a wild-type virus, such as RT or protease, are<br />

optimized for efficient viral replication [30]. In the absence of selective pressure (drug), the wild-type<br />

virus has a fitness advantage over drug-resistant viral variants. However, in the presence of drugs, drugresistant<br />

variants have a fitness advantage over the wild-type because the drug impairs efficiency of the<br />

target enzyme in the wild-type virus [30].<br />

Binding of an NRTI or an NNRTI to wild-type HIV-1 RT interferes with the polymerization reaction.<br />

However, the presence of resistant variants in the population allows the virus to escape, and the variants<br />

to rapidly replace the wild-type virus. Nevertheless, this escape has a price. When the optimized wildtype<br />

virus is replaced <strong>by</strong> the less fit drug-resistant variants, the relative fitness of the virus decreases. In<br />

other words, the enzymatic efficiency of a drug-resistant HIV-1 RT variant is impaired relative to the<br />

wild-type enzyme (see review [131]). If the enzymatic efficiency of a drug-resistant viral variant is<br />

sufficiently impaired, the replication of the variant virus would be significantly decreased. Thus, an<br />

antiviral drug will be useful not because it would completely stop the growth of HIV-1 but because it<br />

selects viral variants whose replication is significantly impaired. Positive clinical benefit results from the<br />

fact that the viral load is decreased owing to reduced replication of the variant virus. As predicted <strong>by</strong> this<br />

model, some HIV-1 RT and protease inhibitors seem to select for relatively less fit drug-resistant<br />

variants. For example, treatment of HIV-1 infection with HBY 097, a quinoxaline inhibitor, induces<br />

development of an HIV-1 RT variant containing the Gly190Glu mutation that appears to have<br />

substantially decreased polymerase activity and replicates relatively slowly [22,84]. Replacement of the<br />

hydrogen atom of Gly190 with an acidic side chain of Glu190 in the hydrophobic NNIBP apparently<br />

interferes with the stability of the enzyme as well as the ability of the NNIBP to bind a hydrophobic<br />

inhibitor. The relative inefficiency of HIV-1 variant containing the Gly190Glu mutation in RT can be<br />

viewed as a positive outcome of the selection pressure provided <strong>by</strong> this particular inhibitor. However,<br />

most of the HIV-1 variants selected <strong>by</strong><br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_67.html [4/5/2004 4:50:46 PM]


Document<br />

currently available antiviral agents are not significantly less fit than the wild-type virus and the clinical<br />

benefits are not obvious. In this regard, new drugs should be designed that would be intended to select<br />

HIV-1 RT variants that are significantly less fit and do not replicate efficiently.<br />

XV. Combination Therapy Using Multiple ANTI-HIV-1 <strong>Drug</strong>s<br />

Page 68<br />

Monotherapy using either NRTIs or NNRTIs has led to the emergence of drug-resistant viral strains of<br />

HIV-1. Though many drug-resistance mutations confer cross-resistance to other inhibitors belonging to<br />

the same class, there are indications that some mutations conferring resistance to certain inhibitors are<br />

incompatible (see reviews [5,11,131]). A multidrug clinical trial with HIV-1 infected patients has shown<br />

that AZT resistance can be reversed <strong>by</strong> mutations that confer resistance to ddI [8]. The Leu74Val<br />

mutation appears to suppress the effects of the Thr215Tyr mutation that confers resistance to AZT<br />

[8,27]. The Met184Val mutation, which causes resistance to 3TC or other oxathiolane-cytosine analogs,<br />

also appears to reverse the effects of the AZT-resistance mutations [27]. Recent clinical studies have<br />

shown that a combination of AZT and 3TC led to a considerable decrease in viral load and a substantial<br />

increase of CD4 cells when compared with monotherapy using AZT alone, even after emergence of the<br />

Met184Val mutation [132]. Another example is the Pro236Leu mutation that confers resistance to<br />

BHAP. The sensitivity of this HIV-1 RT variant to TIBO, nevirapine, and pyridinone is increased ten<br />

fold [133]. Although the NRTIs and NNRTIs target two distinct binding sites of HIV-1 RT and lead to<br />

different sets of resistance mutations, some of the NRTI- and NNRTI-resistance mutations also appear<br />

to be incompatible. For example, the NNRTI-resistance mutations Leu100Ile and Tyr181Cys have been<br />

shown to suppress the effects of some AZT-resistance mutations [11,134]. This has led to the suggestion<br />

that a combination of anti-HIV-1 drugs would be more effective in inhibiting HIV-1 replication than<br />

using individual drugs alone. In fact, both clinical and in vitro studies have shown that combination<br />

therapy has considerable advantages over monotherapy. At least in some cases, the effectiveness of the<br />

therapy increases with an increase in the number of drugs in the combination [5,135]. Combination<br />

therapy may, in addition to increasing antiviral activity, also slow emergence of drug-resistant variants<br />

and may have the added benefit that reducing the dosage of individual drugs can reduce toxicity. It is<br />

generally believed that synergistic drug interactions arise from the fact that certain combinations of drugresistance<br />

mutations are particularly detrimental for the enzyme (and, <strong>by</strong> extension, the virus). This has<br />

focused attention on<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_68.html [4/5/2004 4:50:47 PM]


Document<br />

Page 69<br />

determining the mechanism(s) underlying drug resistance and, from this understanding, to devise ways<br />

for identifying combinations of drugs which might provoke drug-resistance mutations incompatible with<br />

viral survival.<br />

Several protocols have been designed for combination therapy using a variety of anti-HIV-1 drugs (see<br />

reviews [5,11]). Combinations of different drugs that interact with the same binding site of the same<br />

viral protein but lead to mutually antagonistic or suppressive resistance mutations have been studied<br />

extensively, especially for the combined uses of different but structurally related NRTIs (see for<br />

example [136–138]) or NNRTIs [139–141]. Combinations of drugs or inhibitors that target different<br />

sites of the same viral protein, primarily the combination of NRTIs and NNRTIs of HIV-1 RT, show<br />

enhanced inhibition of HIV-1 RT polymerase activity and suppression of the emergence of drugresistance<br />

mutations (for example [142–146]). Experiments have also been conducted with combinations<br />

of drugs that target different viral proteins, e.g., inhibitors of virus adsorption, virus-cell fusion, and/or<br />

uncoating proteins have been tested in combination with protease inhibitors and/or RT inhibitors.<br />

Combinations of AZT with the glycosylation inhibitor castanospermine [147], or with the Tat inhibitor<br />

Ro 24-7429 [148], or with the protease inhibitor Ro 31-8959 [149] have been shown to potently inhibit<br />

HIV-1 viral replication in vitro.<br />

Combination therapy can increase the effectiveness of inhibition and significantly impair efficiency of<br />

viral replication. However, both NRTI- and NNRTI-resistance mutations can affect the positioning of<br />

the nucleic acid and/or the overall structure of HIV-1 RT [23]. These two sets of resistance mutations<br />

can communicate with each other and can result in cross resistance. Moreover, new drug-resistance<br />

mutations that confer cross-resistance to both NRTIs and NNRTIs can be selected, which reduce the<br />

effectiveness of some drug combinations (see reviews [5,11,26]). Biochemical studies showed that both<br />

HIV-1 RT mutants [150] and viral variants [151] could be obtained that are resistant to the combination<br />

of AZT, ddI, and nevirapine. In clinical trials, treatment with AZT and ddI or AZT and ddC led to a<br />

different spectrum of NRTI-resistance mutations [152,153]. The most notable of these new mutations is<br />

Gln151Met, which is located at a position close to the dNTP-binding site. Structural analysis of the HIV-<br />

1 RT/DNA/Fab complex suggests that the side chain of Gln151 in the wild-type enzyme may interact<br />

with the first unpaired template nucleotide. The side chain of this residue may play a role in selecting the<br />

correct base for the incoming nucleotide [72]. Since the RT mutant containing only the Gln151Met<br />

mutation can confer high-level resistance to a number of NRTIs, including AZT, ddI, and ddC, it is not<br />

clear why this mutation did not emerge in monotherapy of these NRTIs. However, Gln151 is relatively<br />

well conserved and mutations at this position may have an unfavorable impact on HIV-1 RT.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_69.html [4/5/2004 4:50:49 PM]


Document<br />

XVI. Perspective<br />

Page 70<br />

Substantial progress has been made in understanding the structure and function of HIV-1 RT and in the<br />

development of anti-HIV-1 inhibitors. However, the genetic flexibility of HIV-1 will continue to make<br />

development of a truly effective antiviral therapy for AIDS an exceptionally difficult task. We are<br />

beginning to understand how to circumvent drug resistance. The accumulated evidence has shown that<br />

the ability of the virus to develop drug resistance is limited and that the drug-resistant viral variants are<br />

less efficient than the wild-type virus. If the selection pressure provided <strong>by</strong> antiviral drugs makes the<br />

virus pay a sufficiently high price, then the viral load can be decreased and there will be a measurable<br />

clinical benefit. Based on a better understanding of the structure-function relationships of HIV-1 RT, we<br />

are now coming to grips with the mechanisms of polymerization, drug inhibition, and drug resistance.<br />

This information should make it possible to develop new or improved HIV-1 RT inhibitors that have<br />

different properties and provoke different patterns of drug-resistance mutations. Though it is likely that<br />

there will be no single drug which would be effective against all HIV-1 variants, we have reasons to<br />

believe that new or improved drugs or, more likely, new drug combinations, will be designed that are<br />

broadly effective against all of the HIV-1 variants that can grow efficiently. Detailed analysis of the<br />

conformational changes among the various HIV-1 RT structures may reveal additional sites (in addition<br />

to the currently known NRTI- and NNRTI-binding sites) for binding new inhibitors able to interfere<br />

with the polymerization and/or the flexibility of the enzyme required for its activity. The considerable<br />

physical and genetic flexibility of HIV-1 RT suggests that more effective anti-RT drugs should be<br />

designed to target the conserved portions of HIV-1 RT that the virus cannot easily afford to change.<br />

Such conserved elements can be identified <strong>by</strong> comparing the sequences of RTs from different<br />

retroviruses; the functions and relative importance of these conserved elements can be determined <strong>by</strong><br />

mutagenesis and biochemical and structural analyses. It is our hope that application of structure-<strong>based</strong><br />

drug design strategies may aid in the development of novel HIV-1 RT inhibitors for a more effective<br />

treatment of HIV-1 infection.<br />

Acknowledgments<br />

We thank the other members of the Arnold and Hughes laboratories and our collaborators for their<br />

helpful discussions and assistance, including Koen Andries, Gail Ferstandig Arnold, Paul Boyer, Arthur<br />

Clark, Jr., Paul Janssen, Jörg-Peter Kleim, Luc Koymans, Tack Kuntz, Karen Lentz, Chris Michejda,<br />

Henri Moereels, Manfred Roesner, Marilyn Kroeger Smith, Rick Smith, Jr., and<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_70.html [4/5/2004 4:50:51 PM]


Document<br />

Chris Tantillo. The work in Edward Arnold's laboratory has been supported <strong>by</strong> Janssen Research<br />

Foundation and NIH grants (AI 27690 and AI 36144). Research in Stephen Hughes' laboratory is<br />

sponsored in part <strong>by</strong> the National Cancer Institute, DHHS, under contract with ABL, and <strong>by</strong> NIGMS.<br />

References<br />

1. Goff SP. Retroviral reverse transcriptase: synthesis, structure, and function. J Acquired Immune<br />

Deficiency Syndromes 1990; 3:817–831.<br />

Page 71<br />

2. Jacobo-Molina A, Arnold E. HIV reverse transcriptase structure-function relationships. Biochemistry<br />

1991; 30:6351–6361.<br />

3. Whitcomb JM, Hughes SH. Retroviral reverse transcription and integration: progress and problems.<br />

Ann Rev Cell Biol 1992; 8:275–306.<br />

4. Le Grice SFJ. Human immunodeficiency virus reverse transcriptase. In: Skalka AM, Goff SP, eds.<br />

Reverse Transcriptase. Plainview, New York: Cold Spring Harbor Laboratory Press, 1993:163–191.<br />

5. De Clercq E. Toward improved anti-HIV chemotherapy: therapeutic strategies for intervention with<br />

HIV infections. J Med Chem 1995; 38:2491–2517.<br />

6. De Clercq E. HIV inhibitors targeted at the reverse transcriptase. AIDS Res Human Retroviruses<br />

1992; 8:119–134.<br />

7. Larder BA. Inhibitors of HIV reverse transcriptase as antiviral agents and drug resistance. In: Skalka<br />

AM, Goff SP, eds. Reverse Transcriptase. Plainview, New York: Cold Spring Harbor Laboratory Press,<br />

1993:205–222.<br />

8. St. Clair MH, Martin JL, Tudor-Williams G, Bach MC, Vavro CL, King DM, et al. Resistance to ddI<br />

and sensitivity to AZT induced <strong>by</strong> a mutation in HIV-1 reverse transcriptase. Science 1991;<br />

253:1557–1559.<br />

9. Richman DD. Resistance of clinical isolates of human immunodeficiency virus to antiretroviral<br />

agents. Antimicrob Agents Chemother 1993; 37:1207–1213.<br />

10. Schinazi RF. Competitive inhibitors of human immunodeficiency virus reverse transcriptase.<br />

Perspectives in <strong>Drug</strong> Discovery and <strong>Design</strong> 1993; 1:151–180.<br />

11. De Clercq E. HIV resistance to reverse transcriptase inhibitors. Biochem Pharmacol 1994;<br />

47:155–169.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_71.html (1 of 2) [4/5/2004 4:50:53 PM]


Document<br />

12. Tantillo C, Ding J, Jacobo-Molina A, Nanni RG, Boyer PL, Hughes SH, et al. Locations of anti-<br />

AIDS drug binding sites and resistance mutations in the three-dimensional structure of HIV-1 reverse<br />

transcriptase: implications for mechanisms of drug inhibition and resistance. J Mol Biol 1994;<br />

243:369–387.<br />

13. Miyasaka T, Tanaka H, Baba M, Hayakawa H, Walker RT, Balzarini J, et al. A novel lead for<br />

specific anti-HIV-1 agents: 1-[(2-hydroxyethoxy)methyl]-6-(phenylthio)thymine. J Med Chem 1989;<br />

32:2507–2509.<br />

14. Pauwels R, Andries K, Desmyter J, Schols D, Kukla MJ, Breslin HJ, et al. Potent and selective<br />

inhibition of HIV-1 replication in vitro <strong>by</strong> a novel series of TIBO derivatives. Nature 1990;<br />

343:470–474.<br />

15. Merluzzi VJ, Hargrave KD, Labadia M, Grozinger K, Skoog M, Wu JC, et al. Inhibition of HIV-1<br />

replication <strong>by</strong> a nonnucleoside reverse transcriptase inhibitor. Science 1990; 250:1411–1413.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_71.html (2 of 2) [4/5/2004 4:50:53 PM]


Document<br />

Page 72<br />

16. Goldman ME, Nunberg JH, O'Brien JA, Quintero JC, Schleif WA, Freund KF, et al. Pyridinone<br />

derivatives: specific human immunodeficiency virus type 1 reverse transcriptase inhibitors with antiviral<br />

activity. Proc Natl Acad Sci USA 1991; 88:6863–6867.<br />

17. Romero DL, Busso M, Tan C-K, Reusser F, Palmer JR, Poppe SM, et al. Nonnucleoside reverse<br />

transcriptase inhibitors that potently and specifically block human immunodeficiency virus type 1<br />

replication. Proc Natl Acad Sci USA 1991; 88:8806–8810.<br />

18. Chimirri A, Grasso S, Monforte A-M, Monforte P, Zappala M. Anti-HIV agents. I. Synthesis and in<br />

vitro anti-HIV evaluation of novel 1H,3H-thiazolo[3,4-a]benzimidazoles. Farmaco 1991; 46:817–823.<br />

19. Chimirri A, Grasso S, Monforte A-M, Monforte P, Zappala M. Anti-HIV agents. II. Synthesis and in<br />

vitro anti-HIV activity of novel 1H3H-thiazolo[3,4-a]benzimidazoles. Farmaco 1991; 46:925–933.<br />

20. Balzarini J, Perez-Perez M-J, San-Felix A, Schols D, Perno C–F, Vandamme A–M, et al. 2',5'-bis-O-<br />

(tert-butyldimethylsilyl)-3'-spiro-5''-(4''-amino-1",2"-oxathiole-2",2"-dioxide)pyrimidine (TSAO)<br />

nucleoside analogues: highly selective inhibitors of human immunodeficiency virus type 1 that are<br />

targeted at the viral reverse transcriptase. Proc Natl Acad Sci USA 1992; 89:4392–4396.<br />

21. Pauwels R, Andries K, De<strong>by</strong>ser Z, Van Dable P, Schols D, Stoffels P, et al. Potent and highly<br />

selective human immunodeficiency virus type 1 (HIV-1) inhibition <strong>by</strong> a series of αanilinophenylacetamide<br />

derivatives targeted at HIV-1 reverse transcriptase. Proc Natl Acad Sci USA<br />

1993; 90:1711–1715.<br />

22. Kleim J–P, Bender R, Billhardt U-M, Meichsner C, Riess G, Roesner M, et al. Activity of a novel<br />

quinoxaline derivative against human immunodeficiency virus type 1 reverse transcriptase and viral<br />

replication. Antimicrob Agents Chemother 1993; 37:1659–1664.<br />

23. Kleim J-P, Roesner M, Winkler I, Paessens A, Kirsch R, Hsiou Y, et al. Selective pressure of a<br />

quinoxaline class nonnucleoside inhibitor of human immunodeficiency virus type 1 (HIV-1) reverse<br />

transcriptase (RT) on HIV-1 replication results in the emergence of nucleoside RT inhibitor specific (RT<br />

L74->V/I, V75->L/I) HIV-1 mutants. Proc Natl Acad Sci USA 1996; 93:34–38.<br />

24. Nunberg JH, Schleif WA, Boots EJ, O'Brien JA, Quintero JC, Hoffman JM, et al. Viral resistance to<br />

human immunodeficiency virus type 1-specific pyridinone reverse transcriptase inhibitors. J Virol 1991;<br />

65:4887–4892.<br />

25. Mellors JW, Dutschman GE, Im G-J, Tramontano E, Winkler SR, Cheng Y-C. Rapid emergence of<br />

HIV-1 resistant to non-nucleoside inhibitors of reverse transcriptase. Antiviral Research 1992; 17 S1:48.<br />

26. Larder BA. Interactions between drug resistance mutations in human immunodeficiency virus type 1<br />

reverse transcriptase. J Gen Virol 1994; 75:951–957.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_72.html (1 of 2) [4/5/2004 4:50:55 PM]


Document<br />

27. Kimberlin DW, Coen, DM, Biron KK, Cohen JI, Lamb RA, McKinlay M, et al. Molecular<br />

mechanisms of antiviral resistance. Antiviral Res 1995; 26:369–401.<br />

28. Wei X, Ghosh SK, Taylor ME, Johnson VA, Emini EA, Deutsch P, et al. Viral dynamics in human<br />

immunodeficiency virus type 1 infection. Nature 1995; 373:117–122.<br />

29. Ho DD, Neumann AU, Perelson AS, Chen W, Leonard JM, Markowitz M. Rapid turnover of plasma<br />

virions and CD4 lymphocytes in HIV-1 infection. Nature 1995; 373:123–126.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_72.html (2 of 2) [4/5/2004 4:50:55 PM]


Document<br />

30. Coffin JM. HIV population dynamics in vivo: implications for genetic variation, pathogenesis, and<br />

therapy. Science 1995; 267:483–489.<br />

31. Kohlstaedt LA, Wang J, Friedman JM, Rice PA, Steitz TA. Crystal structure at 3.5 Å resolution of<br />

HIV-1 reverse transcriptase complexed with an inhibitor. Science 1992; 256:1783–1790.<br />

Page 73<br />

32. Smerdon SJ, Jager J, Wang J, Kohlstaedt LA, Chirino AJ, Friedman JM, et al. <strong>Structure</strong> of the<br />

binding site for nonnucleoside inhibitors of the reverse transcriptase of human immunodeficiency virus<br />

type 1. Proc Natl Acad Sci USA 1994; 91:3911–3915.<br />

33. Ren J, Esnouf R, Garman E, Somers D, Ross C, Kir<strong>by</strong> I, et al. High resolution structures of HIV-1<br />

RT from four RT-inhibitor complexes. Nature Struct Biol 1995; 2:293–302.<br />

34. Ding J, Das K, Tantillo C, Zhang W, Clark AD Jr., Jessen S, et al. <strong>Structure</strong> of HIV-1 reverse<br />

transcriptase in a complex with the nonnucleoside inhibitor α-APA R 95845 at 2.8 Å resolution.<br />

<strong>Structure</strong> 1995; 3:365–379.<br />

35. Ding J, Das K, Moereels H, Koymans L, Andries K, Janssen PAJ, et al. <strong>Structure</strong> of HIV-1<br />

RT/TIBO R 86183 reveals similarity in the binding of diverse nonnucleoside inhibitors. Nature Struct<br />

Biol 1995; 2:407–415.<br />

36. Ren J, Esnouf R, Hopkins A, Ross C, Jones Y, Stammers D, et al. The structure of HIV-1 reverse<br />

transcriptase complexed with 9-chloro-TIBO: lessons for inhibitor design. <strong>Structure</strong> 1995; 3:915–926.<br />

37. Das K, Ding J, Hsiou Y, Clark AD Jr., Moereels H, Koymans L, et al. Crystal structures of 8-Cl and<br />

9-Cl TIBO complexed with wild-type HIV-1 RT and 8-Cl TIBO complexed with the Tyr181Cys HIV-1<br />

RT drug-resistant mutant. J. Mol Biol 1996; 264:1085–1100.<br />

38. Jacobo-Molina A, Ding J, Nanni RG, Clark AD Jr., Lu X, Tantillo C, et al. Crystal structure of<br />

human immunodeficiency virus type 1 reverse transcriptase complexed with double-stranded DNA at<br />

3.0 Å resolution shows bent DNA. Proc Natl Acad Sci USA 1993; 90:6320–6324.<br />

39. Jager J, Smerdon S, Wang J, Boisvert DC, Steitz TA. Comparison of three different crystal forms<br />

shows HIV-1 reverse transcriptase displays an internal swivel motion. <strong>Structure</strong> 1994; 2:869–876.<br />

40. Raag R, Clark AD Jr., Ding J, Jacobo-Molina A, Lu X, Nanni RG, et al. 3.0 Å crystal <strong>Structure</strong> of<br />

HIV-1 reverse transcriptase without dsDNA reveals largescale motion of p66 thumb subdomain. Am<br />

Crystallogr Assoc Mtg Abstr, Ser 2 1994; 18:44.<br />

41. Rodgers DW, Gamblin SJ, Harris BA, Ray S, Culp JS, Hellmig B, et al. The structure of unliganded<br />

reverse transcriptase from the human immunodeficiency virus type 1. Pro Natl Acad Sci USA 1995;<br />

92:1222–1226.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_73.html (1 of 2) [4/5/2004 4:53:09 PM]


Document<br />

42. Esnouf R, Ren J, Ross R, Jones Y, Stammers D, Stuart D. Mechanism of inhibition of HIV-1 reverse<br />

transcriptase <strong>by</strong> non-nucleoside inhibitors. Nature Struct Biol 1995; 2:303–308.<br />

43. Hsiou Y, Ding J, Das K, Clark AD Jr., Hughes SH, Arnold E. <strong>Structure</strong> of unliganded HIV-1 reverse<br />

transcriptase at 2.7 Å resolution: implications of conformational changes for polymerization and<br />

inhibition mechanisms. <strong>Structure</strong> 1996; 4:853–860.<br />

44. Unge T, Knight S, Bhikhabhai R, Lovgren S, Dauter Z, Wilson K, et al. 2.2 Å resolution structure of<br />

the amino-terminal half of HIV-1 reverse transcriptase (fingers and palm subdomains). <strong>Structure</strong> 1994;<br />

2:953–961.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_73.html (2 of 2) [4/5/2004 4:53:09 PM]


Document<br />

45. Majumadar C, Abbotts J, Broder S, Wilson SH. Studies on the mechanism of human<br />

immunodeficiency virus reverse transcriptase: Steady-state kinetics, processivity and polynucleotide<br />

inhibiton. J Biol Chem 1988; 263:15657–15665.<br />

46. Kati WM, Johnson KA, Jerva LF, Anderson KS. Mechanism and fidelity of HIV reverse<br />

transcriptase. J Biol Chem 1992; 267:25988–25997.<br />

47. Balzarini J, Herdewijn P, De Clercq E. Differential patterns of intracellular metabolism of 2',3'didehydro-2',3'-dideoxthymidine<br />

and 3'-azido-2',3'- dideoxythymidine: two potent anti-human<br />

immunodeficiency virus compounds. J Biol Chem 1989; 264:6127–6133.<br />

48. Balzarini J, De Clercq E. 5' -Phosphoribosyl 1-pyrophosphate synthetase converts the acyclic<br />

nucleoside phosphonates 9-(3-hydroxy-2-phosphonyl-methoxypropyl) adenine and 9-(2phosphonylmethoxyethyl)adenine<br />

directly to their antivirally active diphosphate derivatives. J Biol<br />

Chem 1991; 266:8686–8689.<br />

49. Merta AI, Votruba J, Jindrich J, Holy A, Cihlar T, Rosenberg I, et al. Phosphorylation of 9-(2phosphonylmethoxyethyl)adenine<br />

and (S)-9-(3-fluoro-2-phosphonylmethoxypropyl)adenine <strong>by</strong> AMP<br />

(dAMP) kinase from L1210 cells. Biochem Pharmacol 1992; 44:2067–2077.<br />

50. De Clercq E. Broad spectrum anti-DNA virus and anti-retrovirus activity of<br />

phosphonylmethoxyalkyl-purines and -pyrimidines. Biochem Pharmacol 1991; 42:963–972.<br />

51. Balzarini J, Hao Z, Herdewijn P, Johns DG, De Clercq E. Intracellular metabolism and mechanism<br />

of anti-retrovirus action of 9-(2-phosphonyl-methoxyethyl) adenine, a potent anti-human<br />

immunodeficiency virus compound. Proc Natl Acad Sci USA 1991; 88:1499–1503.<br />

Page 74<br />

52. Balzarini J, Holy A, Jindrich J, Dvorakova H, Hao Z, Snoeck R, et al. 9-[(2RS)-3- fluoro-2phosphonylmethoxyethyl]<br />

derivatives of purines: a class of highly selective antiretroviral agents in vitro<br />

and in vivo. Proc Natl Acad Sci USA 1991; 88:4961–4965.<br />

53. Patel PH, Jacobo-Molina A, Ding J, Tantillo C, Clark AD Jr., Raag R, et al. Insights into DNA<br />

polymerization mechanisms from structure and function analysis of HIV-1 reverse transcriptase.<br />

Biochemistry 1995; 34:5351–5363.<br />

54. Boyer PL, Tantillo C, Jacobo-Molina A, Nanni RG, Ding J, Arnold E, et al. The sensitivity of wildtype<br />

human immunodeficiency virus type 1 reverse transcriptase to dideoxynucleotides depends on<br />

template length; the sensitivity of drug-resistant mutants does not. Proc Natl Acad Sci USA 1994;<br />

91:4882–4886.<br />

55. Kellam P, Boucher CA, Larder BA. Fifth mutation in human immunodeficiency virus type 1 reverse<br />

transcriptase contributes to the development of high-level resistance to zidovudine. Proc Natl Acad Sci<br />

USA 1992; 89:1934–1938.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_74.html (1 of 2) [4/5/2004 4:53:11 PM]


Document<br />

56. Martin JL, Wilson JE, Haynes RL, Furman PA. Mechanism of human immunodeficiency virus type<br />

1 resistance to dideoxyinosine. Proc Natl Acad Sci USA 1993; 90:6135–6139.<br />

57. Lacey SF, Reardon JE, Furfine ES, Kunkel TA, Bebenek K, Eckert KA, et al. Biochemical studies<br />

on the reverse transcriptase and RNase H activities from human immunodeficiency virus strains resistant<br />

to 3'-azido-3'-deoxythymidine. J Biol Chem 1992; 267:15789–15794.<br />

58. Larder BA. 3'-Azido-3'-deoxythymidine resistance suppressed <strong>by</strong> a mutation conferring human<br />

immunodeficiency virus type 1 resistance to nonnucleoside<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_74.html (2 of 2) [4/5/2004 4:53:11 PM]


Document<br />

reverse transcriptase inhibitors. Antimicrob Agents Chemother 1992; 36:2664–2669.<br />

Page 75<br />

59. Nanni RG, Ding J, Jacobo-Molina A, Hughes SH, Arnold E. Review of HIV-1 reverse transcriptase<br />

three-dimensional structure: implications for drug design. Perspectives in <strong>Drug</strong> Discovery and <strong>Design</strong><br />

1993; 1:129–150.<br />

60. Boyer PL, Ding J, Arnold E, Hughes SH. <strong>Drug</strong> resistance of human immunodeficiency virus type 1<br />

reverse transcriptase: subunit specificity of mutations that confer resistance to nonnucleoside inhibitors.<br />

Antimicrob Agents Chemother 1994; 38:1909–1914.<br />

61. Mellors JW, Bazmi HZ, Schinazi RF, Roy BM, Hsiou Y, Arnold E, et al. Novel mutations in reverse<br />

transcriptase of human immunodeficiency virus type 1 reduce susceptibility to foscarnet in laboratory<br />

and clinical isolates. Antimicrob Agents Chemother 1995; 39:1087–1092.<br />

62. Boyer PL, Hughes SH. Analysis of mutations of position 184 in the reverse transcriptase of human<br />

immunodeficiency virus type 1. Antimicrob Agents Chemother 1995; 39:1624–1628.<br />

63. Pandey VN, Kaushik N, Rege N, Sarafianos SG, Yadav NS, Modak MJ. Role of M184 of human<br />

immunodeficiency virus type-1 reverse transcriptase in the polymerase function and fidelity of DNA<br />

synthesis. Biochem 1996; 35:2168–2179.<br />

64. Wainberg MA, Drosopoulos WC, Salomon H, Hsu M, Borkow G, Parniak MA, et al. Enhanced<br />

fidelity of 3TC-selected mutant HIV-1 reverse transcriptase. Science 1996; 271:1282–1285.<br />

65. Mitsuya H, Yarchoan R, Broder S. Molecular targets for AIDS therapy. Science 1990;<br />

249:1533–1544.<br />

66. Balzarini J, Van Aerschot A, Herdewijn P, De Clercq E. 5-Chloro-substituted derivatives of 2',3'didehydro-2',3'-dideoxyuridine,<br />

3'-fluoro-2',3'-dideoxyuridine and 3'-azido-2',3'-dideoxyuridine as anti-<br />

HIV agents. Biochem Pharmacol 1989; 38:869–874.<br />

67. Steitz TA, Steitz JA. A general two-metal-ion mechanism for catalytic RNA. Proc Natl Acad Sci<br />

USA 1993; 90:6498–6502.<br />

68. Spence RA, Kati WM, Anderson KS, Johnson KA. Mechanism of inhibition of HIV-1 reverse<br />

transcriptase <strong>by</strong> nonnucleoside inhibitors. Science 1995; 267:988–993.<br />

69. Beese LS, Steitz TA. Structural basis for the 3'-5' exonuclease activity of Escherichia coli DNA<br />

polymerase I: a two metal ion mechanism. EMBO J 1991; 10:25–33.<br />

70. Steitz TA. DNA- and RNA-dependent DNA polymerases. Curr Opin Struct Biol 1993; 3:31–38.<br />

71. Joyce CM, Steitz TA. Function and structure relationships in DNA polymerases. Ann Rev Biochem<br />

1994; 63:777–822.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_75.html (1 of 2) [4/5/2004 4:53:12 PM]


Document<br />

72. Georgiadis MM, Jessen SM, Ogata CM, Telesnitsky A, Goff SP, Hendrickson WA. Mechanistic<br />

implications from the structure of a catalytic fragment of Moloney murine leukemia virus reverse<br />

transcriptase. <strong>Structure</strong> 1995; 3:879–892.<br />

73. Kuntz ID, Blaney JM, Oatley SJ, Langridge R, Ferrin TE. A geometric approach to macromolecularligand<br />

interactions. J Mol Biol 1982; 161:269–288.<br />

74. Kuntz ID. <strong>Structure</strong>-<strong>based</strong> strategies for drug design and discovery. Science 1992; 257:1078–1082.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_75.html (2 of 2) [4/5/2004 4:53:12 PM]


Document<br />

75. Ring CS, Sun E, McKerrow JH, Lee GK, Rosenthal PJ, Kuntz ID, et al. <strong>Structure</strong>-<strong>based</strong> inhibitor<br />

design <strong>by</strong> using protein models for the development of antiparasitic agents. Proc Natl Acad Sci USA<br />

1993; 90:3583–3587.<br />

Page 76<br />

76. Balzarini J, Perez-Perez M-J, San-Felix A, Velazquez S, Camarasa M-J, De Clercq E. 2',5'-bis-O-<br />

(tert-butyldimethylsilyl)-3'-spiro-5"-(4"-amino-1",2"-oxathiole-2",2"-dioxide) (TSAO) derivatives of<br />

purine and pyrimidine nucleosides as potent and selective inhibitors of human immunodeficiency virus<br />

type 1. Antimicrob Agents Chemother 1992; 36:1073–1080.<br />

77. Rittinger K, Divita G, Goody R. Human immunodeficiency virus reverse transcriptase substrateinduced<br />

conformational changes and the mechanism of inhibition <strong>by</strong> nonnucleoside inhibitors. Proc Natl<br />

Acad Sci USA 1995; 92:8046–8049.<br />

78. Palaniappan C, Fay PJ, Bambara RA. Nevirapine alters the cleavage specificity of ribonuclease H of<br />

human immunodeficiency virus 1 reverse transcriptase. J Biol Chem 1995; 270:4861–4869.<br />

79. Jonckheere H, Taymans J-M, Balzarini J, Velazquez S, Camarasa M-J, Desmyter J, et al. Resistance<br />

of HIV-1 reverse transcriptase against [2',5'-bis-O-(tert-butyldimethylsilyl)-3'-spiro-5"-(4"-amino-1",2"oxathiole-2",2"-dioxide)]<br />

(TSAO) derivatives is determined <strong>by</strong> the mutation Glu 138->Lys on the p51<br />

subunit. J Biol Chem 1994; 269:25255–25258.<br />

80. Pauwels R, Andries K, De<strong>by</strong>ser Z, Kukla MJ, Schols D, Breslin HJ, et al. New TIBO derivatives are<br />

potent inhibitors of HIV-1 replication and are synergistic with 2',3'-dideoxynucleoside analogues.<br />

Antimicrob Agents Chemother 1994; 38:2863–2870.<br />

81. Kashman Y, Gustafson KR, Fuller RW, Cardellina JH II, McMahon JB, Currens MJ, et al. The<br />

calanolides, a novel HIV-inhibitory class of coumarin derivatives from the tropical rainforest tree,<br />

Calophyllum langericum. J Med Chem 1992; 35:2735–2743.<br />

82. Boyer PL, Currens MJ, McMahon JB, Boyd MR, Hughes SH. Analysis of nonnucleoside drugresistant<br />

variants of human immunodeficiency virus type 1 reverse transcriptase. J Virol 1993;<br />

67:2412–2420.<br />

83. Goldman ME, O'Brien JA, Ruffing TL, Schleif WA, Sardana VV, Bynes VW, et al. A<br />

nonnucleoside reverse transcriptase inhibitor active on human immunodeficiency type 1 isolates<br />

resistant to related inhibitors. Antimicrob Agents Chemother 1993; 37:947–949.<br />

84. Kleim J-P, Bender R, Kirsch R, Meichsner C, Paessens A, Riess G. Mutational analysis of residue<br />

190 of human immunodeficiency virus type 1 reverse transcriptase. Virology 1994; 200:696–701.<br />

85. Smith MBK, Rouzer CA, Taneyhill LA, Smith NA, Hughes SH, Boyer PL, et al. Molecular<br />

modeling studies of HIV-1 reverse transcriptase nonnucleoside inhibitors: Total energy of complexation<br />

as a predictor of drug placement and activity. Protein Science 1995; 4:2203–2222.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_76.html (1 of 2) [4/5/2004 4:53:14 PM]


Document<br />

86. De Clercq E. Antiviral therapy of human immunodeficiency virus infections. Clin Microbiol Rev<br />

1995; 8:200–239.<br />

87. Champoux JJ. Roles of ribonuclease H in reverse transcription. In: Skalka AM, Goff SP, eds.<br />

Reverse Transcriptase. Plainview, New York: Cold Spring Harbor Laboratory Press, 1993:103–117.<br />

88. Telesnitsky A, Goff SP. Strong-stop strand transfer during reverse transcription. In: Skalka AM,<br />

Goff SP eds. Reverse Transcriptase. Plainview, New York: Cold Spring Harbor Laboratory Press,<br />

1993:49–83.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_76.html (2 of 2) [4/5/2004 4:53:14 PM]


Document<br />

89. Hughes SH, Arnold E, Hostomsky Z. RNase H of retroviral reverse transcriptases. Plainview, New<br />

York: Cold Spring Harbor Laboratory Press, 1996; in press.<br />

90. Schatz O, Mous J, Le Grice SFJ. HIV-1 RT-associated ribonuclease H displays both endonuclease<br />

and 3' 5' exonuclease activity. EMBO J 1990; 9:1171–1176.<br />

91. Tanese N, Goff SP. Domain structure of the Moloney murine leukemia virus reverse transcriptase:<br />

mutational analysis and separate expression of the DNA polymerase and RNase H activities. Proc Natl<br />

Acad Sci USA 1988; 85:1777–1781.<br />

92. Tanese NT, Telesnitsky A, Goff SP. Abortive reverse transcription <strong>by</strong> mutants of Moloney murine<br />

leukemia virus deficient in the reverse transcriptase-associated RNase H function. J Virol 1991;<br />

65:4387–4397.<br />

93. Tisdale M, Schulze T, Larder BA, Moelling K. Mutations within the RNase H domain of human<br />

immunodeficiency virus type 1 reverse transcriptase abolish virus infectivity. J Gen Virol 1991;<br />

72:59–66.<br />

94. Davies JF, Hostomska Z, Hostomsky Z, Jordan SR, Matthews DA. Crystal structure of the<br />

ribonuclease H domain of HIV-1 reverse transcriptase. Science 1991; 252:88–95.<br />

95. Chattopadhyay D, Finzel BC, Munson SH, Evans B, Sharma SK, Strakalattis NA, et al.<br />

Crystallographic analysis of an active HIV-1 ribonuclease H domain show structural features that<br />

distinguish it from the inactive form. Acta Crystallogr 1993; D49:423–427.<br />

96. Prasad VR, Goff SP. Linker insertion mutagenesis of the human immunodeficiency virus<br />

transcriptase expressed in bacteria: definition of the minimal polymerase domain. Proc Natl Acad Sci<br />

USA 1989; 86:3104–3108.<br />

97. Hizi A, Hughes SH, Shaharabany M. Mutational analysis of the ribonuclease H activity of human<br />

immunodeficiency virus 1 reverse transcriptase. Virology 1990; 175:575–580.<br />

98. Hizi A, Tal R, Hughes SH. Mutational analysis of the DNA polymerase and ribonuclease H<br />

activities of human immunodeficiency virus type 2 reverse expressed in Escherichia coli. Virology<br />

1991; 180:339–346.<br />

Page 77<br />

99. Boyer PL, Ferris AL, Clark P, Whitmer J, Frank P, Tantillo C, et al. Mutational analysis of the<br />

fingers and palm subdomains of human immunodeficiency virus type-1 (HIV-1) reverse transcriptase. J<br />

Mol Biol 1994; 243:472–483.<br />

100. Ding J, Jacobo-Molina A, Tantillo C, Lu X, Nanni RG, Arnold E. Buried surface analysis of HIV-1<br />

reverse transcriptase p66/p51 heterodimer and its interaction with dsDNA template/primer. J Mol<br />

Recognition 1994; 7:157–161.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_77.html (1 of 2) [4/5/2004 4:53:17 PM]


Document<br />

101. Boyer PL, Ferris AL, Hughes SH. Cassette mutagenesis of the reverse transcriptase of human<br />

immunodeficiency virus type 1. J Virol 1992; 66:1031–1039.<br />

102. Boyer PL, Ferris AL, Hughes SH. Mutational analysis of the fingers domain of human<br />

immunodeficiency virus type-1 reverse transcriptase. J Virol 1992; 66:7533–7537.<br />

103. Crouch RJ, Dirksen ML, Ribonuclease H. In: Linn SM, Roberts RJ, eds. Nuclease. Plainview, New<br />

York: Cold Spring Harbor Laboratory Press, 1982:211–241.<br />

104. Kanaya S, Kohara A, Miura Y, Sekiguchi A, Iwai S, Inoue H, et al. Identification of the amino acid<br />

residues involved in an active site of Eschericia coli ribonuclease H <strong>by</strong> site-directed mutagenesis. J Biol<br />

Chem 1990; 265:4615–4621.<br />

105. Smith JS, Roth MJ. Purification and characterization of an active human immunodeficiency virus<br />

type 1 RNase H domain. J Virol 1993; 67:4037–4049.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_77.html (2 of 2) [4/5/2004 4:53:17 PM]


Document<br />

106. Cirino NM, Cameron CE, Smith JS, Rausch JW, Roth MJ, Benkovic SJ, et al. Divalent cation<br />

modulation of the ribonuclease functions of human immunodeficiency virus reverse transcriptase.<br />

Biochemistry 1995; 34:9936–9943.<br />

Page 78<br />

107. Schatz O, Cromme FV, Gruninger-Leitch F, Le Grice SFJ. Point mutations in conserved amino<br />

acid residues within the C-terminal domain of HIV-1 reverse transcriptase specifically repress RNase H<br />

function. FEBS Lett 1989; 257:311–314.<br />

108. Volkmann S, Wohrl BM, Tisdale M, Moelling K. Enzymatic analysis of two HIV-1 reverse<br />

transcriptase mutants with mutations in carboxyl-terminal amino acid residues conserved among<br />

retroviral ribonucleoside H. J Biol Chem 1993; 268:2674–2683.<br />

109. Mizrahi V, Brooksbank RL, Nkabinde NC. Mutagenesis of the conserved aspartic acid 443,<br />

glutamic acid 478, asparagine 494, and aspartic acid 498 residues in the ribonuclease H domain of<br />

p66/p51 human immunodeficiency virus type 1 reverse transcriptase. J Biol Chem 1994;<br />

269:19245–19249.<br />

110. Yang W, Hendrickson WA, Crouch RJ, Satow Y. <strong>Structure</strong> of ribonuclease H phased at 2 Å<br />

resolution <strong>by</strong> MAD analysis of the selenomethionyl protein. Science 1990; 249:1398–1405.<br />

111. Katayanagi K, Miyagawa M, Matsushima M, Ishikawa M, Kanaya S, Ikehara M, et al. Threedimensional<br />

structure of ribonuclease H from E. coli. Nature 1990; 347:306–309.<br />

112. Katayanagi K, Okumura M, Morikawa K. Crystal structure of Escherichia coli RNase HI in<br />

complex with Mg 2+ at 2.8 Å resolution. Proteins 1993; 17:337–346.<br />

113. Loya S, Tal R, Kashman Y, Hizi A. Illimaquinone, a selective inhibitor of the RNase H activity of<br />

human immunodeficiency virus type 1 reverse transcriptase. Antimicrob Agents Chemother 1990;<br />

34:2009–2012.<br />

114. Loya S, Hizi A. The interaction of illimaquinone, a selective inhibitor of the RNase H activity, with<br />

the reverse transcriptases of human immunodeficiency and murine leukemia retroviruses. J Biol Chem<br />

1993; 268:9323–9328.<br />

115. Restle T, Muller B, Goody RS. Dimerization of human immunodeficiency virus reverse<br />

transcriptase. J Biol Chem 1990; 265:8986–8988.<br />

116. Muller B, Restle T, Kuhnel H, Goody RS. Expression of the heterodimeric form of human<br />

immunodeficiency virus type 2 reverse transcriptase in Escherichia coli and characterization of the<br />

enzyme. J Biol Chem 1991; 266:14709–14713.<br />

117. Restle T, Muller B, Goodys RS. RNase H activity of HIV reverse transcriptase is confined<br />

exclusively to the dimeric forms. FEBS Letters 1992; 300:97–100.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_78.html (1 of 2) [4/5/2004 4:53:19 PM]


Document<br />

118. Divita G, Rittinger K, Geourjon C, Deleage G, Goody RS. Dimerization kinetics of HIV-1 and HIV-<br />

2 reverse transcriptase: A two step process. J Mol Biol 1994; 145:508–521.<br />

119. Wang J, Smerdon SJ, Jager J, Kohlstaedt LA, Rice PA, Friedman JM, et al. Structural basis of<br />

asymmetry in the human immunodeficiency virus type 1 reverse transcriptase heterodimer. Proc Natl<br />

Acad Sci USA 1994; 91:7242–7246.<br />

120. Tuerk C, MacDougal S, Gold L. RNA pseudoknots that inhibit human immunodeficiency virus<br />

type 1 reverse transcriptase. Proc Natl Acad Sci USA 1992; 89:6988–6992.<br />

121. Green L, Waugh S, Binkley JP, Hostomska Z, Hostomsky Z, Tuerk C. Comprehensive chemical<br />

modification interference and nucleotide substitution analysis of an RNA pseudoknot inhibitor to HIV-1<br />

reverse transcriptase. J Mol Biol 1995; 247:60–68.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_78.html (2 of 2) [4/5/2004 4:53:19 PM]


Document<br />

122. Marshall WS, Beaton G, Stein CA, Matsukura M, Caruthers MH. Inhibition of human<br />

immunodeficiency virus activity <strong>by</strong> phosphorodithioate oligodeoxycytidine. Proc Natl Acad Sci USA<br />

1992; 89:6265–6269.<br />

123. Marshall WS, Caruthers MH. Phosphorodithioate DNA as a potential therapeutic drug. Science<br />

1993; 259:1564–1570.<br />

124. Loya S, Hizi A. The inhibition of human immunodeficiency virus type 1 reverse transcriptase <strong>by</strong><br />

avarol and avarone derivatives. FEBS Lett 1990; 269:131–134.<br />

Page 79<br />

125. Loya S, Rudi A, Tal R, Kashman Y, Loya Y, Hizi A. 3,5,8-Trihydroxy-4-quinolone, a novel natural<br />

inhibitor of the reverse transcriptase of human immunodeficiency viruses type 1 and type 2. Arch<br />

Biochem Biophys 1994; 309:315–322.<br />

126. Sheridan RP Rusinko A, Nilkantan R, Venkatraghavan R. Searching for pharmacophores in large<br />

coordinate data bases and its use in drug design. Proc Natl Acad Sci USA 1989; 86:8165–8169.<br />

127. Marshall GR, Barry CD, Bosshard HE, Dammkoehler RA, Dunn DA. The conformational<br />

parameter in drug design: the active analog approach. In: Olson EC, Christoffersens RE eds. Computer-<br />

Assisted <strong>Drug</strong> <strong>Design</strong>. Honolulu, Hawaii: American Chemical Society, 1979:205–226.<br />

128. Marshall GR. Computer-aided drug design. Ann Rev Pharmacol Toxicol 1987; 27:193–213.<br />

129. Najera I, Holguin A, Quinones-Mateu ME, Munoz-Fernandez MA, Najera R, Lopez-Galindez C, et<br />

al. Pol gene quasispecies of human immunodeficiency virus: mutations associated with drug resistance<br />

in virus from patients undergoing no drug therapy. J Virol 1995; 69:23–31.<br />

130. Erickson JW. The not-so-great escape. Nature Struct Biol 1995; 2:523–529.<br />

131. Arnold E, Das K, Ding J, Yadav PNS, Hsiou, Y, Boyer PL, Hughes SH. Targeting HIV reverse<br />

transcriptase for anti-AIDS drug design. <strong>Drug</strong> <strong>Design</strong> and Discovery 1996; 13:29–47.<br />

132. Larder BA, Kemp SD, Harrigan PR. Potential mechanism for sustained antiretroviral efficacy of<br />

AZT-3TC combination therapy. Science 1995; 269:696–699.<br />

133. Dueweke TJ, Pushkarskaya T, Poppe SM, Swaney SM, Zhao JQ, Chen ISY, et al. A mutation in<br />

reverse transcriptase of bis(heteroaryl) piperazine-resistant human immunodeficiency virus type 1 that<br />

confers increased sensitivity to other nonnucleoside inhibitors. Proc Natl Acad Sci USA 1993;<br />

90:4713–4717.<br />

134. Larder BA. AZT resistance suppressed <strong>by</strong> a mutation conferring HIV-1 resistance to nonnucleoside<br />

reverse transcriptase inhibitors. Antimicrob Agents Chemother 1992; 36:1171–1174.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_79.html (1 of 2) [4/5/2004 4:53:21 PM]


Document<br />

135. Mazzulli T, Rusconi S, Merrill DP, D'Auilla RT, Moonis M, Chou T-C, et al. Alternating verse<br />

continuous drug regimens in combination chemotherapy of human immunodeficiency virus type 1<br />

infection in vitro. Antimicrob Agents Chemother 1994; 38:656–661.<br />

136. Smith MS, Brian EL, De Clercq E, Pagano JS. Susceptibility of human immunodeficiency virus<br />

type 1 replication in vitro to acyclic adenosine analogs and synergy of the analogs with 3'-azido-3'deoxythymidine.<br />

Antimicrob Agents Chemother 1989; 33:1482–1486.<br />

137. Dornsife RE, St. Clair MH, Huang AT, Panella TJ, Koszalka GW, Burns CL, et al. Anti-human<br />

immunodeficiency virus synergism <strong>by</strong> zidovudine (3'-azidothymi<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_79.html (2 of 2) [4/5/2004 4:53:21 PM]


Document<br />

dine) and didanosine (dideoxyinosine) contrasts with their additive inhibition of normal human<br />

marrow progenitor cells. Antimocrob Agents Chemother 1991; 35:322–328.<br />

138. Antonelli G, Dianzani F, Bellarosa D, Turriziani O, Riva E, Gentile A. <strong>Drug</strong> combination of AZT<br />

and ddI: synergism of action and prevention of appearance of AZT-resistance. Antiviral Chem<br />

Chemother 1994; 5:51–55.<br />

139. Balzarini J, Karlsson A, Perez-Perez MJ, Camarasa MJ, Tarpley WG, De Clercq E. Treatment of<br />

human immunodeficiency virus type 1 (HIV-1)-infected cells with combinations of HIV-1-specific<br />

inhibitors results in a different resistance pattern than does treatment with single-drug therapy. J Virol<br />

1993; 67:5353–5359.<br />

140. Balzarini J, Perez-Perez M-J, Velazquez S, San-Felix A, Camarasa M-J, De Clercq E, et al.<br />

Suppression of the breakthrough of human immunodeficiency virus type 1 (HIV-1) in cell culture <strong>by</strong><br />

thiocarboxanilide derivatives when used individually or in combination with other HIV-1-specific<br />

inhibitors (i.e., TSAO derivatives). Proc Natl Acad Sci USA 1995; 92:5470–5474.<br />

141. Fletcher RS, Arion D, Borkow G, Wainberg MA, Dmitrienko GI, Parniak MA. Synergistic<br />

inhibition of HIV-1 reverse transcriptase DNA polymerase activity and virus replication in vitro <strong>by</strong><br />

combinations of carboxanilide nonnucleoside compounds. Biochemistry 1995; 34 10106–10112.<br />

142. Baba M, Ito M, Shigeta S, Tanaka H, Miyasaka T, Ubasawa M, et al. Synergistic inhibition of<br />

human immunodeficiency virus type 1 replication <strong>by</strong> 5-ethyl-1-ethoxymethyl-6-(phenylthio)uracil (E-<br />

EPU) and azidothymidine in vitro. Antimocrob Agents Chemother 1991; 35:1430–1433.<br />

143. Richman DD, Rosenthal AS, Skoog M, Echner RJ, Chou T-C, Sabo JP, et al. BIRG-587 is active<br />

against zidovudine-resistant human immunodeficiency virus type 1 and synergistic with zidovudine.<br />

Antimicrob Agents Chemother 1991; 35:305–308.<br />

Page 80<br />

144. Buckheit RW Jr., White EL, Germany-Decker J, Allen LB, Ross LJ, Shannon WM, et al. Cell<strong>based</strong><br />

and biochemical analysis of the anti-HIV activity of combinations of 3'-azido-3'-deoxythymidine<br />

and analogues of TIBO. Antiviral Chem Chemother 1994; 5:35–42.<br />

145. Chong K-T, Pagano PJ, Hinshaw RR. Bis(heteroacyl)piperazine reverse transcriptase inhibitor in<br />

combination with 3'-azido-3'-deoxythymidine or 2',3'-dideoxycytidine synergistically inhibits human<br />

immunodeficiency virus type 1 replication in vitro. Antimicrob Agents Chemother 1994; 38:288–293.<br />

146. Brennan TM, Taylor DL, Bridges CG, Leyda JP, Tyms AS. The inhibition of human<br />

immunodeficiency virus type 1 in vitro <strong>by</strong> a non-nucleoside reverse transcriptase inhibitor MKC-442<br />

alone and in combination with other anti-HIV compounds. Antiviral Res 1995; 26:173–187.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_80.html (1 of 2) [4/5/2004 4:53:24 PM]


Document<br />

147. Johnson VA, Walker BD, Barlow MA, Paradis TJ, Chou T-C, Hirsch MS. Synergistic inhibition of<br />

human immunodeficiency virus type 1 and type 2 replication in vitro <strong>by</strong> castanospermine and 3'-azido-3'deoxythymidine.<br />

Antimicrob Agents Chemother 1989; 33:53–57.<br />

148. Connell EV, Hsu M-C, Richman DD. Combinative interaction of a human immunodeficiency virus<br />

(HIV) tat antagonist with HIV reverse transcriptase inhibitors and an HIV protease inhibitor. Antimicrob<br />

Agents Chemother 1994; 38:348–352.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_80.html (2 of 2) [4/5/2004 4:53:24 PM]


Document<br />

149. Craig JC, Duncan IB, Whittaker L, Roberts NA. Antiviral synergy between inhibitors of HIV<br />

proteinase and reverse transcriptase. Antiviral Chem Chemother 1993; 4:161–166.<br />

150. Emini EA, Graham DJ, Gotlib L, Condra JH, Byrnes VW, Schleif WA. HIV and multidrug<br />

resistance. Nature 1993; 364:679.<br />

151. Larder BA, Kellam P, Kemp SD. Convergent combination therapy can select viable multidrug<br />

resistant HIV-1 in vitro. Nature 1993; 365:451–453.<br />

Page 81<br />

152. Shafer RW, Kozal MJ, Winters M, Iversen AKN, Katenstein DA, Ragni MV, et al. Combination<br />

therapy with zidovudine and didanosine selects for drug-resistant human immunodeficiency virus type 1<br />

strains with unique patterns of pol gene mutations. J Infect Dis 1994; 169:722–729.<br />

153. Shirasaka T, Kavlick MF, Ueno T, Gao W-Y, Kojima E, Alcaide ML, et al. Emergence of human<br />

immunodeficiency virus type 1 variants with resistance to multiple dideoxynucleosides in patients<br />

receiving therapy with dideoxynucleosides. Proc Natl Acad Sci USA 1995; 92:2398–2402.<br />

154. Gu Z, Gao Q, Fang H, Salomon H, Parniak MA, Goldberg E, et al. Identification of a mutation of<br />

codon 65 in the IKKK motif of reverse transcriptase that encodes human immunodeficiency virus<br />

resistance to 2',3'-dideoxycytidine and 2',3'-dideoxy-3'-thiacytidine. Antimicrob Agents Chemother<br />

1994; 38:275–281.<br />

155. Lander BA, Kemp SD. Multiple mutations in HIV-1 reverse transcriptase confer high-level<br />

resistance to zidovudine (AZT). Science 1989; 246:1155–1158.<br />

156. Fitzgibbon JE, Howell RM, Haberzettl CA, Sperber SJ, Gocke DJ, Dubin DT. Human<br />

immunodeficiency virus type 1 pol gene mutations which cause decreased susceptibility to 2',3'dideoxycytidine.<br />

Antimicrob Agents Chemother 1992; 36:153–157.<br />

157. Prasad VR, Lowy I, Santos TDL, Chiang L, Goff SP. Isolation and characterization of a<br />

dideoxyguanosinetriphosphate-resistant mutant of human immunodeficiency virus reverse transcriptase.<br />

Proc Natl Acad Sci USA 1991; 88:11363–11367.<br />

158. Havlir D, Murphy R, Saag M, Kaul I, Johnson V, Richman DD. Nevirapine: further dose escalation<br />

of monotherapy (600 mg/daily) and combination therapy with zidovudine. In: The First National<br />

Conference on Human Retroviruses and Related Infections, Washington D.C., 1993:101.<br />

159. Saag MS, Sommadossi JP, Rainey D, Myers M, Cort S, Hall D, et al. A pharmacokinetic and<br />

antiretroviral activity study of nevirapine in combination with zidovudine plus zalcitabine (ZDV/ddC),<br />

zidovudine plus didanosine (ZDV/ddI), or didanosine (ddI) Alone. In: The First National Conference on<br />

Human Retroviruses and Related Infections, Washington D.C., 1993:102.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_81.html (1 of 2) [4/5/2004 4:53:26 PM]


Document<br />

160. Gu Z, Gao Q, Li X, Parniak MA, Wainberg MA. Novel mutation in the human immunodeficiency<br />

virus type 1 reverse transcriptase gene that encodes cross-resistance to 2',3'-dideoxyinosine and 2',3'dideoxycytidine.<br />

J Virol 1992; 66:7128–7135.<br />

161. Tisdale M, Kemp SD, Parry NR, Larder BA. Rapid in vitro selection of human immunodeficiency<br />

virus type 1 resistant to 3'-thiacytidine inhibitors due to a mutation in the YMDD region of reverse<br />

transcriptase. Proc Natl Acad Sci USA 1993; 90:5653–5656.<br />

162. Larder BA, Coates KE, Kemp SD. Zidovudine-resistant human immunodeficiency virus selected<br />

<strong>by</strong> passage in cell culture. J Virol 1991; 65:5232–5236.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_81.html (2 of 2) [4/5/2004 4:53:26 PM]


Document<br />

163. Lacey SF, Larder BA. A novel mutation (V75T) in the HIV-1 reverse transcriptase confers<br />

resistance to 2',3'-didehydro-2',3'-dideoxythymidine (D4T) in cell culture. Antimicrob Agents<br />

Chemother 1994; 38:1428–1432.<br />

164. Byrnes VW, Sardana VV, Schleif WA, Condra JH, Waterbury JA, Wolfgang JA, et al.<br />

Comprehensive mutant enzyme and viral variant assessment of human immunodeficiency virus type 1<br />

reverse transcriptase resistance to nonnucleoside inhibitors. Antimicrob Agents Chemother 1993;<br />

37:1576–1579.<br />

Page 82<br />

165. Vasudevachari MB, Battista C, Lane HC, Psallidopoulos MC, Zhao B, Cook J, et al. Prevention of<br />

the spread of HIV-1 infection with nonnucleoside reverse transcriptase inhibitors. Virology 1992;<br />

190:269–277.<br />

166. Balzarini J, Karlsson A, Perez-Perez M-J, Vrang L, Walbers J, Zhang H, et al. HIV-1-specific<br />

reverse transcriptase inhibitors show differential activity against mutant strains containing different<br />

amino acid substitutions in the reverse transcriptase. Virology 1993; 192:246–253.<br />

167. Balzarini J, Velazquez S, San-Felix A, Karlsson A, Perez-Perez M-J, Camarasa M-J, et al. Human<br />

immunodeficiency virus type 1-specific [2',5'-bis-O-[tert-butyldimethylsilyl] - β - D - ribofuranosyl] -3' -<br />

spiro-5'' -(4'' -amino - 1",2" - oxathiole-2", 2'-dioxide)-purine analogues show a resistance spectrum that<br />

is different from that of the human immunodeficiency virus type-1-specific nonnucleoside analogues.<br />

Mol Pharmacol 1993; 43:109–114.<br />

168. Bacolla A, Shih C-K, Rose JM, Piras G, Warren TC, Grygon CA, et al. Amino acid substitutions in<br />

HIV-1 reverse transcriptase with corresponding residues from HIV-2. J Biol Chem 1993;<br />

268:16571–16577.<br />

169. Demeter L, Resnick L, Nawaz T, Timpone JG Jr., Batts D, Reichman RC. Phenotypic and<br />

genotypic analysis of ateviridine (ATV) susceptibility of HIV-1 isolates obtained from patients receiving<br />

ATV monotherapy in a phase I clinical trial (ACTG 187): comparison to patients receiving combination<br />

therapy with ATV and zidovudine. In: Third Workshop on Viral Resistance. Gaithersburg, Maryland,<br />

1993.<br />

170. Tisdale M, Parry NR, Cousens D, St. Clair MH and Boone LR. Anti-HIV activity of (lS-4R)-4-[2amino-6-cyclopropylamino-9H-purin-9-yl]-2-cyclopentene-1-methanol<br />

(1592U89). In: Abstracts of the<br />

34th Interscience Conference on Antimicrobial Agents and Chemotherapy. Orlando, Florida, 1994:92.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_82.html [4/5/2004 4:53:28 PM]


Document<br />

3<br />

Retroviral Integrase: <strong>Structure</strong> as a Foundation for <strong>Drug</strong> <strong>Design</strong><br />

Alison B. Hickman and Fred Dyda<br />

National Institutes of Health, Bethesda, Maryland<br />

I. Introduction<br />

A. Retroviral Lifecycle<br />

Page 83<br />

The human immunodeficiency virus (HIV) is one of only a few retroviruses known to infect humans. It<br />

is estimated that approximately twenty-two million people are now infected worldwide [1]. With only a<br />

tiny number of exceptions, infection ultimately leads to the development of the lethal condition of<br />

acquired immunodeficiency syndrome, or AIDS. To date, only a handful of drugs have been shown to<br />

have any effect on the course of the disease. These are, in general, relatively ineffective at significantly<br />

prolonging life, and drug resistance develops rapidly. Equally discouraging, vaccines have not yet been<br />

developed to prevent infection.<br />

The retroviral lifecycle presents several steps that can be targeted as possible sites of intervention <strong>by</strong><br />

inhibitors. As shown in Figure 1, when a retrovirus encounters a host cell, specific recognition between<br />

proteins on the surface of the virus and receptors on the host cell surface leads to membrane fusion. The<br />

viral core then enters the cell cytoplasm where the process of reverse transcription begins. The<br />

requirement of the conversion of viral RNA to double-stranded DNA is a feature unique to retroviruses.<br />

With the recent exception of the protease inhibitor saquinavir, ritonavir, and indinavir, the drugs<br />

approved to date <strong>by</strong> the U.S. Food and <strong>Drug</strong> Administration (FDA) for the treatment of HIV infection<br />

have been nucleoside analogs targeted against the viral enzyme that carries out this conversion, reverse<br />

transcriptase.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_83.html [4/5/2004 4:53:30 PM]


Document<br />

Figure 1<br />

Retroviral lifecycle as summarized in Reference 67. Reprinted <strong>by</strong><br />

permission of Springer-Verlag Publishing Co., New York, NY.<br />

Page 84<br />

Although details of the timing of reverse transcription, nuclear localization, and integration are not yet<br />

clear, it is generally recognized that the movement of double-stranded viral DNA across the nuclear<br />

membrane is followed <strong>by</strong> insertion, or integration, of the viral genome into a host-cell chromosome. The<br />

viral DNA moves as part of a larger “preintegration complex,” a high-molecular-weight aggregate<br />

whose composition has not yet been completely defined.<br />

The end result of integration is the incorporation of the viral DNA into the DNA of the host cell. Once<br />

there, the provirus can serve as a template for the production of mRNA, allowing for the synthesis of<br />

viral proteins. These are assembled at the cell membrane to produce new viral particles, which then bud<br />

off to seek out new cells to infect. The integrated viral DNA is also necessarily copied whenever the<br />

host cell undergoes cell division. The insidious nature of<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_84.html (1 of 2) [4/5/2004 4:54:27 PM]


Document<br />

the virus arises because, once integrated, the viral DNA can no longer be distinguished from host cell<br />

DNA and has become a permanent fixture of the host cell genome.<br />

B. Rationale for <strong>Drug</strong> <strong>Design</strong> Against Integrase to Fight HIV and AIDS<br />

Page 85<br />

It has been demonstrated that the chemical steps that comprise DNA integration are carried out <strong>by</strong> the<br />

viral protein, integrase (IN). Integrase is encoded <strong>by</strong> the 3' end of the viral pol gene, which also codes<br />

for two other viral enzymes, the protease and reverse transcriptase. These three enzymes are initially<br />

synthesized as part of a larger polyprotein that is subsequently cleaved <strong>by</strong> the protease to the individual<br />

proteins.<br />

Why is integrase a good target for drug-design efforts to prevent infection <strong>by</strong> halting the viral replication<br />

cycle? First, integration is required for replication. In the absence of integration, the virus is unable to<br />

continue to make copies of itself. Secondly, the enzyme that carries out integration is virally encoded,<br />

and when the viral genome is disrupted so that functional integrase is no longer made, sustained viral<br />

replication does not occur [2]. This demonstrates that if viral integrase can be effectively inhibited, there<br />

is no protein encoded <strong>by</strong> the host cell that can replace it and carry out viral integration. Finally, since<br />

mammalian cells do not have enzymes capable of integrating HIV DNA, there are no vital host cell<br />

analogs of integrase carrying out essential reactions whose function would be blocked <strong>by</strong> integrase<br />

inhibitors.<br />

Effective inhibition of HIV integrase would add to the number of sites at which the virus replication<br />

cycle can be halted. One can imagine treatment protocols in which a mixture of inhibitors, each aimed at<br />

a different viral protein, could be administered. This is known as divergent combination therapy. As<br />

structural details are a necessary starting point for rational drug design, we present here our recent<br />

results on the high-resolution three-dimensional structure of the catalytic core domain of HIV-1<br />

integrase [3]. We also review the current literature discussing integrase inhibitors and present thoughts<br />

on ways in which knowledge of the chemical reactions carried out <strong>by</strong> integrase and its structure might<br />

direct the development of effective inhibitors.<br />

II. Biochemical Reactions Catalyzed By HIV Integrase<br />

A. In Vivo Integration<br />

Details of the initial chemical reactions that occur during HIV integration are now well understood (for<br />

reviews, see References 4,5). Once linear double-stranded DNA is available for integration, (Figure 2a)<br />

integrase then removes<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_85.html [4/5/2004 4:54:29 PM]


Document<br />

Figure 2<br />

In vivo reactions carried out <strong>by</strong> HIV integrase.<br />

Page 86<br />

two nucleotides from each 3' end of the viral DNA (Figure 2b). The two nucleotides are removed as a<br />

dinucleotide rather than in two individual steps. The specificity for this reaction is conferred <strong>by</strong> the third<br />

and fourth nucleotides from each 3' end, a -CA sequence that is absolutely conserved. Once two<br />

nucleotides have been removed, leaving recessed 3' hydroxyl groups, the next step is the joining of the 3'<br />

ends to target DNA (Figure 2c, d). This process, known as double-ended integration, occurs on opposite<br />

strands such that the joining sites on each of the target DNA strands are separated <strong>by</strong> five base pairs. The<br />

final step in integration is the repair of the single-stranded gaps generated <strong>by</strong> the staggered insertion of<br />

the viral 3' ends on opposite strands; this regenerates an intact double-stranded DNA molecule (Figure<br />

2e and f). Gap repair is probably carried out <strong>by</strong> host cell DNA repair systems.<br />

One necessary consequence of retroviral integration is the duplication of five base pairs of host cell<br />

DNA on either side of the integrated provirus.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_86.html [4/5/2004 4:54:34 PM]


Document<br />

Another is the loss from the ends of the viral DNA of the original two base pairs that preceded the<br />

conserved 3'-CA.<br />

B. In Vitro Assays to Monitor Integrase Activity<br />

Page 87<br />

In contrast to the in vivo reaction, concerted integration in vitro of two HIV DNA ends into a target<br />

DNA molecule separated <strong>by</strong> a 5 base-pair stagger occurs very inefficiently. However, in vitro systems<br />

have been developed [6,7] using recombinant HIV integrase that have allowed the chemistry of the<br />

single-ended integration event to be studied in fine detail. It is possible and routine to use short, doublestranded<br />

synthetic oligonucleotides that mimic the viral ends to monitor the removal of two nucleotides<br />

from 3' ends (denoted 3' processing or cutting) and the subsequent insertion of one 3' processed DNA<br />

molecule into another (known as strand transfer or joining). Typical reactions are depicted in Figure 3.<br />

The stereochemical mechanism of 3' processing and strand transfer has been investigated using DNA<br />

substrates that incorporate phosphorothioate link-ages [8]. For both reactions, the introduced chiral<br />

centers are inverted in the products, implying that the reactions occur via a one-step in-line displacement<br />

mechanism rather than via a covalent intermediate.<br />

A third assay of integrase activity, termed disintegration, has more recently been developed [9] that<br />

monitors the apparent reversal of strand-<br />

Figure 3<br />

Reactions carried out <strong>by</strong> integrase in vitro, using short oligonucleotide substrates.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_87.html [4/5/2004 4:54:38 PM]


Document<br />

Page 88<br />

transfer (Figure 3). While disintegration probably has no physiological significance, it has been useful in<br />

defining aspects of integrase biochemistry.<br />

The three in vitro activities of integrase require divalent metal ions as cofactors. The only two metals<br />

that support these activities are Mn 2+ and Mg 2+. Since quite high metal concentrations must be added to<br />

assays (1–10 mM for optimal activity), it has been presumed that Mg 2+ is the ion used in vivo.<br />

C. Evidence for a Multimer as the Active Unit of Integrase<br />

Several lines of evidence demonstrate that the active unit of integrase is a multimer. It is clear, as an<br />

isolated protein in solution, that integrase forms dimers [6,10–12], and it has been shown <strong>by</strong><br />

sedimentation equilibrium studies that Rous sarcoma virus (RSV) integrase exists in reversible<br />

equilibrium between monomeric, dimeric, and tetrameric forms [13]. Protein-protein cross-linking<br />

studies of HIV-1 [14] and RSV [15] integrases confirm the existence of protein dimers and tetramers in<br />

solution, and in vivo, the yeast GAL4 two-hybrid system has demonstrated that HIV-1 integrase can<br />

interact with itself [16].<br />

Complementation studies using mutant proteins in vitro provide compelling evidence that the active<br />

form of integrase must be at least a dimer [14, 17]. This can be inferred from the result that when certain<br />

inactive forms of integrase—generated either <strong>by</strong> truncation or point mutation—are mixed, robust<br />

activity can be reconstituted. This indicates that different monomers in a multimer are capable of<br />

providing different essential functions in the context of an active complex.<br />

Collectively, these studies suggest that integrase acts as a multimer. This would also seem the most<br />

straight-forward model to explain the observation that viral integration requires two coordinated cutting<br />

and joining reactions on the target DNA during strand transfer. However, physical studies have not yet<br />

addressed what form of integrase actually binds to DNA and carries out the chemical reactions of<br />

integration.<br />

III. Properties of HIV-1 Integrase<br />

A. Domain <strong>Structure</strong> of Retroviral Integrases<br />

A consistent view of the domain structure of retroviral integrases has emerged <strong>by</strong> combining the results<br />

from biochemical studies using deletion and site-specific mutants, limited proteolysis experiments, and<br />

sequence comparisons among the family of retroviral integrases. The organization of the domains of<br />

integrase is shown schematically in Figure 4.<br />

The central domain of HIV-1 integrase, consisting approximately of residues 50 to 200, is largely<br />

conserved among retroviral integrases, and forms<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_88.html [4/5/2004 4:54:40 PM]


Document<br />

Figure 4<br />

Schematic domain structure of HIV-1 integrase as adapted from<br />

Engelman et al. [19]. <strong>Structure</strong>s of two domains, the catalytic core<br />

extending from residues 50 to 212 [3] and the nonspecific<br />

DNA-binding domain from residues 220 to 270 [28,29], have<br />

recently been determined <strong>by</strong> x-ray crystallography and NMR<br />

spectroscopy, respectively.<br />

Page 89<br />

the protease-resistant core of the protein [18,19]. Within this domain are three invariant residues that<br />

comprise the “D,D-35-E motif” (see alignment in Figure 5). These are residues Asp64, Asp116, and<br />

Glu152. Even conservative substitution of any of these residues leads to loss of all three in vitro<br />

activities of integrase in parallel [19–21]. The D,D-35-E motif is also observed in retrotransposons and<br />

some prokaryotic transposases. A truncated form of HIV-1 integrase consisting of residues 50 to 212 is<br />

capable of disintegration [22], implying that the catalytic site is contained within this domain. These<br />

observations and the absolute requirement for metals for in vitro activity have led to the proposal that<br />

the three acidic residues constitute a divalent metal-binding site capable of binding one or two Mg 2+ or<br />

Mn 2+ ions to form a catalytically active enzyme. As will be seen in later sections, the three-dimensional<br />

structure of the core domain of HIV-1 integrase is consistent with this hypothesis. The catalytic<br />

mechanism may be, therefore, similar to the one proposed <strong>by</strong> Beese and Steitz for the 3'–5' exonuclease<br />

of E. coli DNA polymerase I [23]. It is proposed that for phosphate bond cleavage, one metal ion helps<br />

form the attacking hydroxide ion while the other stabilizes a pentacovalent intermediate around the<br />

phosphorus.<br />

The C-terminus of HIV-1 integrase, consisting approximately of residues 210 to 288, includes the<br />

dominant nonspecific DNA binding domain [24, 25], which has been more finely mapped to residues<br />

220–270 [26]. The C-terminus is the least conserved region of retroviral integrases; only one residue,<br />

Trp235, is absolutely invariant. However, it has been reported that removal of only five amino acids<br />

from the C-terminus of HIV-1 integrase is enough to severely reduce its 3' processing and strand transfer<br />

activities [27]. One notable feature of the C-terminus is its high proportion of positively charged<br />

residues. As discussed in Section IV.E, the structure of part of this region has recently been determined<br />

using NMR spectroscopic methods [28,29].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_89.html (1 of 2) [4/5/2004 4:54:43 PM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_90.html [4/5/2004 4:54:55 PM]<br />

Page 90


Document<br />

Figure 5<br />

Alignment of amino acid sequences of retroviruses. See Engelman et al. [19] for details.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_91.html [4/5/2004 4:55:13 PM]<br />

Page 91


Document<br />

Page 92<br />

The role of the N-terminus of integrase, residues 1 to 50, is still unclear. Within this region are four<br />

strictly conserved amino acids: two His and two Cys residues. In HIV-1 integrase, the spacing is His-X 3-<br />

His-X 23-Cys-X 2-Cys. This cluster of His and Cys residues is reminiscent of a zinc-binding motif, and it<br />

has been demonstrated that the full-length protein binds Zn 2+ [22,30], and that the separately expressed<br />

domain consisting only of residues 1 to 55 also binds Zn 2+ stoichiometrically [31]. However, it has not<br />

been shown that either the structural integrity or the enzymatic activities of integrase require Zn 2+. While<br />

truncation of residues from the N-terminus of HIV integrase results in loss of 3' processing and strand<br />

transfer activities [22,25], in the case of RSV integrase, the N-terminal region can be replaced <strong>by</strong><br />

unrelated sequences, and the enzyme is still capable of all three in vitro activities [32].<br />

B. Biophysical Properties of Full-Length Recombinant HIV-1 Integrase<br />

It has been known for some time that recombinant HIV-1 integrase is a particularly poorly behaved<br />

protein in solution. Its solubility in most usual buffers is limited to approximately 1 mg/mL, and even<br />

then only in the presence of high concentrations of NaCl. At ~1 mg/mL, HIV-1 integrase slowly<br />

precipitates out of solution, revealing one of its characteristic features, a tendency towards aggregation.<br />

These properties of the protein are not unreasonable, since in its viral environment integrase is probably<br />

never required to be a soluble protein. To maintain the integrity of preintegration complexes, it may<br />

even be advantageous for the protein to have the properties of being rather insoluble and sticking to<br />

itself, nucleic acid, and perhaps other proteins.<br />

C. Properties of Truncated Versions of HIV-1 Integrase<br />

It has been our approach to protein structure determination <strong>by</strong> x-ray crystallography that it is imperative<br />

to begin with well-characterized and well-behaved protein. In particular, it is important that the protein<br />

be reasonably soluble and monodisperse in solution. Unfortunately, as discussed above, recombinant<br />

HIV-1 integrase satisfies neither of these conditions. One approach we and others have taken to<br />

circumvent these problems has been to examine truncated versions of HIV-1 integrase to determine if<br />

removal of amino acids from either terminus or both affects solubility and aggregation properties.<br />

Although we observed that two proteins we constructed, IN 213–288 and IN 50–288, were more soluble than<br />

the full-length HIV-1 integrase, IN 1–288 [33, and unpublished observations], our first target protein for<br />

crystallization efforts was the core domain of HIV-1 integrase consisting of residues 50 to 212, IN 50–212.<br />

We reasoned that this protein domain was likely to be compact and<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_92.html [4/5/2004 4:55:15 PM]


Document<br />

well folded since it is relatively resistant to proteolysis. As it is also active for disintegration, we<br />

concluded that it contained the enzyme active site.<br />

Page 93<br />

We exploited the convenience of histidine-tag (HT) technology to develop methods to purify large<br />

quantities of IN 50–212 [12]. A 20-amino-acid histidine-containing tag was added to the N-terminus of<br />

HIV-1 IN 50–212 [22] to allow rapid purification on nickel affinity columns. It was subsequently removed<br />

<strong>by</strong> thrombin cleavage. Biophysical studies showed that although buffer conditions could be identified<br />

where the protein was soluble to ~ 4 mg/mL, under these conditions the protein was highly aggregated<br />

(unpublished observations). Although the aggregation problem could be largely avoided <strong>by</strong> the addition<br />

of high concentrations of the zwitterionic detergent CHAPS, conditions could not be identified under<br />

which protein crystals formed in the presence of CHAPS.<br />

D. Systematic Mutation of Hydrophobic Residues to Improve Protein Solubility<br />

As it became clear that IN 50–212 was crystallographically challenged, a condition readily understood in<br />

terms of its aggregation problems and low solubility, a more radical approach was undertaken to try and<br />

improve its biophysical properties. Hydrophobic residues in the catalytic core were targeted for sitespecific<br />

mutation according to the following criteria: where two or more hydrophobic residues were<br />

encountered close together in the primary amino acid sequence, they were each changed to an alanine<br />

residue. When a hydrophobic residue stood alone, it was mutated to lysine. In this way, 29 different<br />

mutant proteins of IN 50–212 were rapidly generated using the overlapping polymerase chain reaction<br />

(PCR) and screened for improved solubility properties [34]. Three mutated proteins were identified that<br />

were more soluble at lower NaCl concentration than the unmutated core (V165K, F185K, and the<br />

double mutation of W131A/W132A). However, one of these in which Phe185 was mutated to Lys had<br />

dramatically improved solubility and was ultimately crystallized and its three-dimensional structure<br />

determined [3]. The remarkable biophysical properties of this single point mutant of IN 50–212 have<br />

recently been described [34].<br />

IV. <strong>Structure</strong> Of The Catalytic Core Domain Of HIV-1 Integrase<br />

A. Description of the <strong>Structure</strong><br />

The Overall Protein Fold<br />

The three-dimensional structure of the catalytic core domain of HIV-1 integrase is centered on a mixed<br />

five stranded β sheet flanked <strong>by</strong> several helices forming<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_93.html [4/5/2004 4:55:17 PM]


Document<br />

Figure 6<br />

Molscript stereo figure of the three-dimensional structure of the catalytic core of<br />

HIV-1 integrase. The two catalytically essential aspartic acid residues (D64 and D116)<br />

visible in the x-ray structure are highlighted.<br />

Page 94<br />

an α-β meander sandwich (see Figure 6) [35]. In the crystal structure, interpretable electron density<br />

starts at Cys56, leading to a short loop. The first β strand starts at Gly59 and runs until Val68. The two<br />

central residues of a type I' reverse β turn, Glu69 and Gly70, change the polypeptide chain direction to<br />

form the second β strand between residues Lys71 and His78, which runs antiparallel with the first<br />

strand. A type I'β turn follows, with Val79 and Ala80 changing the chain direction again to form the<br />

third β strand between Ser81 and Ile89, which runs antiparallel with the second strand. A short loop<br />

between Pro90 and Glu92 leads to the first α helix (helix A) between Thr93 and Trp108. This helix<br />

packs against the bottom face of the sheet formed <strong>by</strong> the first three antiparallel strands <strong>by</strong> several<br />

hydrophobic interactions. A short loop (Pro109 and Val110) leads to the fourth β strand between Lys111<br />

and His114. This strand is parallel with the first. A short loop starting at Thr115 leads to helix B, a one<br />

turn helix between Gly118 and Thr122, followed <strong>by</strong> helix C between Ser123 and Ala133. This helix<br />

runs parallel to and packs against helix A. The residues Gly134 and Ile135 form a short loop prior to the<br />

fifth and last β strand of the structure between Lys136 and Ala138. This short strand is parallel with the<br />

first and the fourth. There is no interpretable electron density due to disorder between Gly140 and<br />

Met154. At Met154, the fourth α helix (helix D) starts and runs until Ala169 on the top face of the sheet<br />

formed <strong>by</strong> the first three β strands. The residue Glu170 leads into the next helix (helix E) running<br />

between His171 and Lys186, the first residue of a short basic sequence (Lys186, Arg187, and Lys188).<br />

Together with<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_94.html (1 of 2) [4/5/2004 4:55:35 PM]


Document<br />

Gly193 and Tyr194, Lys188 and Gly189 form two short antiparallel β strands separated <strong>by</strong> a turn of<br />

three residues (Gly190, Ile191, and Gly192) not involved in main chain hydrogen bonds. The first<br />

residue of the last helix, (helix F), which runs until Asp212, is Ser194.<br />

Three Conserved Acidic Residues at the Enzyme Active Site<br />

Page 95<br />

There are four amino acids in the core domain sequence that are absolutely conserved among retroviral<br />

integrases: Asp64, Asp116, Glu152, and Lys159. The three acidic residues form the conserved D,D-35-<br />

E motif and have been shown to be essential for catalysis (see Section III.A). The role of Lys159 in<br />

retroviral integrases is not obvious; its replacement with Val does not abolish catalytic activity, although<br />

there is a decrease in strand transfer activity [20].<br />

The first essential acidic residue, Asp64, is located in the middle of the first β strand, while Asp116 is in<br />

a loop region right after the fourth β strand. These two residues define the active-site area and they are<br />

right next to each other three-dimensionally with their α-carbons separated <strong>by</strong> only 6.7 Å. The closest<br />

approach is 3.4 Å between Oδ1 of Asp64 and Cβ of Asp116. These residues are on the surface of the<br />

molecule, not part of any obvious substrate-binding cleft. The third catalytically essential acidic residue,<br />

Glu152, is in the disordered and hence crystallographically invisible region between Gly140 and<br />

Met154. Its location therefore must be inferred from other parts of the structure and from available threedimensional<br />

structures of related proteins. The location of Met154, the residue only two positions<br />

upstream from Glu152, is known because of interpretable electron density. The distance between the αcarbons<br />

of Glu152 and Met154 cannot be larger than about 7.3 Å, which constrains Glu152 to the<br />

neighborhood of the two other essential carboxylates, allowing it to contribute to the formation of a<br />

divalent metal-binding site.<br />

More recently, a crystal structure of the avian sarcoma virus (ASV) integrase core domain was solved<br />

[36]. Within this domain, ASV integrase has 24% sequence identity to the HIV-1 integrase core and, as<br />

expected, its three-dimensional structure is remarkably similar. The ASV integrase core in its native<br />

form has much better solution properties than the HIV-1 integrase core, and did not require any point<br />

mutations to render it crystallizable. Due to this fact and perhaps also due to its different crystal packing<br />

interactions, the crystal lattice of the ASV integrase core domain is somewhat more ordered than that of<br />

HIV-1. The two three-dimensional structures can be aligned quite well, using 74 α-carbons, with an rms<br />

deviation of only 1.4 Å in these α-carbon positions. The most remarkable difference between the two<br />

structures is that in the ASV structure the electron density is interpretable in all parts of the molecule.<br />

This is not to say, however, that serious disorder is not present. For example, in one particular loop, the<br />

temperature factors are above 70 Å 2 for the α carbons, indicating larger than 1 Å mean displacement<br />

value for these atoms. The corresponding<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_95.html [4/5/2004 4:55:37 PM]


Document<br />

Page 96<br />

region in the HIV-1 structure is the uninterpretable stretch from residues 140 to 153, showing that<br />

beyond the three-dimensional similarity, the molecules also share a similar disorder pattern despite their<br />

different crystal packing interactions. It is clear that in the apoenzyme (metal-free) form of the HIV-1<br />

integrase core, disorder is present in parts of the active site. However, in the holo form, structural<br />

stability must be necessary to form a metal-binding site.<br />

Position of the Third Essential Carboxylate<br />

Does the structure of ASV integrase give us a hint about the likely conformation of the polypeptide<br />

chain around Glu152 in the holoenzyme form of the HIV-1 integrase core? The answer is probably yes,<br />

considering the overall similarity of the structures. The first residue after the disordered part in the HIV-<br />

1 integrase core is Met154, which is also the first residue in helix C. The corresponding helix in the<br />

ASV core is longer, running between Gln153 and Gly175. The residue analogous to Glu152 is Glu157<br />

in the ASV integrase core structure, located a half-turn upstream from Ala159, which corresponds to<br />

Met154 in the HIV-1 integrase structure. It is plausible to assume, therefore, that the polypeptide chain<br />

in the holoenzyme form of HIV-1 integrase would also be in a helical conformation around Glu152, and<br />

its location would be very close to the one that can be inferred from the location of Glu157 in the ASV<br />

integrase core. Secondary structure prediction also supports this assumption, assigning α-helical<br />

structure around Glu152. Why does this helical turn show significant disorder in the HIV-1 integrase<br />

structure? The answer might be found in the amino acid sequence: Pro145 is a highly conserved residue<br />

among retroviral integrases, the only exception being ASV integrase where it is substituted with a Ser.<br />

Since the main chain nitrogen of a proline is not capable of participating in hydrogen bonding, it is very<br />

rarely found in helices. It is likely that if the polypeptide chain around Glu152 were helical in the<br />

holoenzyme form of the HIV-1 integrase core, then this helix would start after Pro145. There is no such<br />

restriction in the ASV integrase core, and it is possible that this is why helix C is longer in ASV than in<br />

HIV. This may also explain the disorder around Glu152 in the HIV-1 integrase core, since it is closer to<br />

the end of the helix and more susceptible to disordering effects. A longer helix and therefore a more<br />

ordered active site in the apo form may be a unique feature of the ASV integrase core.<br />

B. Similarity to Other Polynucleotidyl Transferases<br />

Overall Protein Folds<br />

The catalytic core domain of HIV-1 integrase has a topologically identical fold with the RNase H<br />

domain of HIV-1 reverse transcriptase [37], the RuvC Holli-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_96.html [4/5/2004 4:55:39 PM]


Document<br />

Figure 7<br />

<strong>Structure</strong>s of the catalytic core of HIV-1 integrase, HIV-1 RNase<br />

H, RuvC, and the core domain of MuA transposase demonstrating<br />

similarities in folding topology. The catalytically essential residues<br />

are highlighted.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_97.html (1 of 2) [4/5/2004 4:55:55 PM]<br />

Page 97


Document<br />

day junction resolving enzyme [38], and also the core domain of the phage MuA transposase [39] (see<br />

Figure 7). In the case of HIV-1 integrase, RNase H, and RuvC, definite three-dimensional similarity<br />

extends only to the ends of the last β strand; from this point, the structures diverge. In HIV-1 RNase H,<br />

there is only one more α helix corresponding to helix D in the HIV-1 integrase core but in a 40°<br />

different orientation. In RuvC there are three more helices, with the last one running parallel to helix D<br />

of HIV-1 integrase, but in the opposite direction and also 4.6 Å closer to the β sheet. In contrast, the<br />

homology between HIV-1 inte-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_97.html (2 of 2) [4/5/2004 4:55:55 PM]


Document<br />

grase and the MuA transposase extends until the carboxyl termini of their respective catalytic domains<br />

with three very similarly positioned and oriented α helices.<br />

Page 98<br />

Limited three-dimensional alignment between the four molecules can be accomplished <strong>by</strong> identifying<br />

structurally homologous stretches along the polypeptide chains if this search is restricted to between the<br />

first well-ordered amino terminal residue and the end of the last β strand. For the core domain of HIV-1<br />

integrase, this corresponds to the region between Ile60 and Gln137. Using the corresponding region in<br />

the MuA transposase, the two structures can be aligned with an rms deviation of 1.7 Å over 69 α-carbon<br />

positions. The main differences between these structures are two insertions in the transposase core: an<br />

11-residue β-stranded extension replacing the turn between the first and the second strand in the HIV-1<br />

integrase core, and a 15-residue extension before helix B with no secondary structure. Both of these<br />

extensions interact with the downstream nonspecific DNA-binding domain of the transposase. For HIV-<br />

1 RNase H, the alignment results in a rms deviation of 2.0 Å over 48 alignable α-carbon positions. The<br />

position of helix A is significantly different in RNase H, as it shifts more than 5 Å toward the β sheet.<br />

There is also an additional 2.5 turn helix following the fourth β strand and a 5-residue loop after this<br />

helix. For RuvC the alignment yields an rms deviation of 2.0 Å over 50 alignable α-carbon positions. In<br />

this case, the differences are mostly the result of longer secondary structure elements in RuvC. Of all the<br />

molecules compared, the HIV-1 integrase core is the smallest, with the most compact design in the<br />

region where these alignments were performed. For comparison, let us mention again that the<br />

homologous ASV integrase core can be aligned with an rms deviation of 1.4 Å over 74 α-carbon<br />

positions in this region.<br />

Both topological similarity and three-dimensional homology with the MuA transposase was expected<br />

<strong>based</strong> on the similarity of the reactions the enzymes catalyze, but the relationship with RNase H and<br />

RuvC was a surprise. This discovery led to the proposal of a new polynucleotidyl transferase<br />

superfamily. All the members of the superfamily are divalent metal ion-dependent endonucleases, and<br />

they all leave 3'-OH and 5'-phosphate groups at the site of cleavage. All the members of the superfamily<br />

display their catalytically essential acidic residues at the same general location. There are three such<br />

residues in HIV-1 integrase, RNase H, and the MuA transposase, while there are four in RuvC. Two of<br />

these residues are always located on the same three-dimensional structural elements, while the location<br />

of the third varies. The Asp64 residue HIV-1 integrase corresponds to Asp443 in HIV-1 RNaseH, Asp7<br />

in RuvC, and Asp269 in the MuA transposase. Based on the three-dimensionally aligned structures, the<br />

α-carbon positions of these residues cluster quite well around that of HIV-1 integrase, with an rms<br />

deviation of 0.84 Å. All these residues are located in the middle of first β strand. The side chain torsion<br />

angle, Chi 1, is<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_98.html [4/5/2004 4:55:56 PM]


Document<br />

-62° for HIV-1 integrase, a frequently observed rotamer. For the other three molecules this Chi 1 value<br />

varies between -142° and -156°, a common range for rotamer angles. The preference for the first<br />

rotamer of the HIV side chain is probably due to the 2.73 Å-long hydrogen bond between Oο°2 of<br />

Asp64 and Nε2 of Gln62. There is no such interaction in the other molecules. The different rotamer<br />

causes a 2.46 Å rms separation of the Cγ positions around the HIV-1 Cγ, compared to only 1.63 Å<br />

around the Cγ of Asp443 in HIV-1 RNase H, indicting the carboxylate of Asp64 in HIV-1 integrase as<br />

the outlier.<br />

Position of the Second Carboxylate Residue<br />

Page 99<br />

In all the analogous structures, the second essential carboxylate resides just after the end of the fourth β<br />

strand. The main-chain atoms are not involved in strand-forming direct hydrogen bonds, therefore the<br />

chain diverts from running parallel with the first strand, forming a small cleft. The equivalent residues<br />

are Asp116 in HIV-1, Asp498 in HIV-1 RNase H, and Glu66 in RuvC. The clustering is weaker than for<br />

Asp64; the rms deviation is 1.77 Å in α-carbon position around the HIV-1 integrase residue.<br />

Interestingly, <strong>by</strong> including the structurally otherwise highly homologous ASV integrase core, the rms<br />

deviation increases to 2.15 Å due to the 3 Å distance between the Cα of Asp116 of HIV-1 integrase and<br />

that of the corresponding residue, Asp121 of ASV integrase. The rms separation between the ASV<br />

position and the rest of the cluster (now excluding HIV-1 integrase) is 2.2 Å, which is rather high,<br />

identifying the ASV residue as the outlier. For the Chi 1 torsion angles, all three preferred rotamers are<br />

present: Chi 1 is -86° for HIV-1 integrase, 73° for HIV-1 RNase H, and -173° for the MuA transposase.<br />

The RuvC Chi 1 value is not included in this comparison because it has a Glu in this position. The<br />

different Chi 1 values combined with the variation in α-carbon positions leads to a somewhat more<br />

scattered Cγ (or Cδ for Glu66 in RuvC) position with an rms deviation of 2.71 Å around Cγ of Asp116<br />

of HIV-1 integrase. By including the ASV molecule, the scatter increases to 3.82 Å due to the 6 Å<br />

distance between Cγ of Asp116 in HIV-1 integrase and Cγ of Asp121 of ASV integrase.<br />

The Third Essential Carboxylate<br />

The location of the third essential catalytic carboxylate varies between different members of the<br />

superfamily. For HIV-1 integrase, Glu152 is in a disordered region with no interpretable electron<br />

density. Based on the location of the equivalent residue in the ASV integrase, its position is assumed to<br />

be on helix D, as discussed above. For RNase H, Glu478 is located on helix A, with its side chain<br />

pointing toward the other two carboxylates to complete the divalentmetal-binding site. Such metal<br />

binding has been observed crystallographically<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_99.html [4/5/2004 4:51:54 PM]


Document<br />

Page 100<br />

[37]. For RuvC, four residues have been shown to be essential for catalysis [40]. The third of these,<br />

Asp138, is at the amino end of the last helix of the structure. In the three-dimensionally aligned<br />

structures, this helix is parallel with the helix D in HIV-1 integrase, although it is running in the opposite<br />

direction. The position of helix D in RuvC is also significantly different, mainly due to a 13 Å shift<br />

along its axis toward the active site, placing Asp138 close to the other two carboxylates. The fourth<br />

essential residue, Asp141, is on the first turn of the same helix, very close in the aligned structures to the<br />

essential Glu157 of ASV integrase, their α carbons separated <strong>by</strong> only 1.6 Å. For the MuA transposase,<br />

its third essential carboxylate, Glu392, is in a loop region, just one residue upstream from the amino end<br />

of a helix, the topological equivalent of helix D. Unexpectedly, this residue turns away from the region<br />

defined <strong>by</strong> the two other carboxylates to a position where it clearly cannot contribute to the formation of<br />

a metal-binding site. It is likely, therefore, that the conformation of the polypeptide chain around Glu392<br />

in the transposase core observed in the crystal structure belongs to an inactive form. In this case, a<br />

conformational change upon transposase tetramer assembly or perhaps upon substrate binding is<br />

required for activity.<br />

Significance of the Disordered Region<br />

From the point of view of HIV-1 integrase, it is interesting to note that the apparently flexible part of the<br />

MuA transposase structure is topologically equivalent with the disordered and uninterpretable part of the<br />

integrase. Similarly, in the crystal structure of the isolated RNase H domain of HIV-1 reverse<br />

transcriptase, a five-residue loop in a topologically equivalent location is disordered and therefore<br />

uninterpretable. In the ASV integrase core, the corresponding loop is visible but with rather high<br />

mobility. It seems that some kind of disorder or flexibility in this region is a common feature of the<br />

superfamily. Crystal structures of enzyme-substrate or enzyme-inhibitor complexes will tell us the<br />

functional significance of this flexibility as well as the exact configuration of the active site.<br />

C. The Dimer Interface<br />

HIV-1 integrase is active as a multimer, and the catalytic core domain alone forms dimers in solution,<br />

even at low protein concentration (see Section II.C). In the crystal structure, a roughly spherical dimer of<br />

about 45 Å diameter was observed, formed <strong>by</strong> a crystallographic two-fold axis. The dimer has a large<br />

solvent-excluded surface of 1300 Å 2 per monomer. This area is close to what is expected for dimers in<br />

this molecular weight range [41]. Therefore, we are convinced that in the crystal structure the authentic<br />

dimer is present. This was subsequently confirmed <strong>by</strong> the structure of the ASV integrase core. Although<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_100.html [4/5/2004 4:51:56 PM]


Document<br />

Page 101<br />

crystallized under different conditions, forming crystals that are in a different space group with different<br />

crystal-packing interactions, the dimer observed for the ASV core is essentially identical with that of<br />

HIV-1 integrase, although the solvent-excluded surface is smaller (only 740 Å 2). This difference is<br />

largely due to the absence of helix F in the ASV structure.<br />

The core domain dimer of HIV-1 integrase is held together <strong>by</strong> several hydrophobic and polar<br />

interactions. Hydrophobic interactions dominate the interface between helix E from one monomer and<br />

helices A and B from the other. There is a buried salt bridge between Glu87 of the third β strand and<br />

Lys103 on helix A. There are also some water-mediated polar interactions between these two secondary<br />

structure elements. There are direct hydrogen bonds between residues on helix A and residues on helix E<br />

across the interface including one between Lys185 (the substitution responsible for the improved<br />

solubility and therefore crystallizability) and the main-chain carbonyl on Ala105. In the ASV core,<br />

His198 is in this position, forming a very similar hydrogen bond with the carbonyl oxygen of Ala110.<br />

There are about 10 water molecules buried in the interface, all involved in hydrogen bonds. The part of<br />

the solvent-accessible surface of the monomer which becomes buried upon dimer formation displays a<br />

high degree of shape compatibility with itself: <strong>by</strong> rotating it 180° around the crystallographic two-fold<br />

axis, the resulting surface will fit the original one without forming large pockets. It is possible that the<br />

core domain of HIV-1 integrase has evolved to optimize this compatibility in order to increase its<br />

stability. It would be interesting to see the effect on protein activity of site-directed mutations aimed at<br />

disrupting this interface and hence the dimer (or possibly the higher order multimers in the context of<br />

the full-length protein).<br />

D. Implications of Crystallographic Dimer for the Chemistry of Catalysis<br />

The nearly spherical nature of the dimer formed <strong>by</strong> two monomers of the integrase catalytic core places<br />

active sites on respective monomers on opposite sides of the dimer: approximately 35 Å separates the<br />

carboxylate oxygens of Asp64 of each monomer. While we are convinced that the observed dimer is not<br />

an artifact or consequence of crystallization, it would seem difficult to reconcile this distance with the<br />

observation that, during in vivo strand transfer, cuts on the target DNA occur with a separation of five<br />

base pairs, corresponding to 15–20 Å in B-form DNA. How can a single dimer accomplish this? One<br />

possibility is that the cuts do not occur simultaneously. One end of the viral DNA could be joined <strong>by</strong> a<br />

reaction at one active site, followed <strong>by</strong> carefully controlled movement of DNA and protein relative to<br />

one another such that the second active site is now positioned five base pairs away from the initial site of<br />

strand transfer. It has been proposed, in a variation on this theme, that the first strand-transfer<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_101.html [4/5/2004 4:51:57 PM]


Document<br />

Page 102<br />

reaction is followed <strong>by</strong> DNA relaxation (unwinding, rotation, etc.), resulting in the site on the target<br />

DNA for the second transfer reaction site now being located close to the active site of the second<br />

monomer [42]. An alternate possibility is that a multimer larger than a dimer is responsible for the<br />

coordinated cutting and strand-transfer reactions. For example, two contacting dimers can be modeled<br />

such that two active sites of the resulting tetramer are located 15–20 Å apart. It is also possible that even<br />

higher order multimers are involved.<br />

There is, as yet, no convincing evidence in support of any one model. The observation that RSV<br />

integrase cuts target DNA with a six-base-pair stagger rather than the five observed for HIV correlates<br />

intriguingly with the apparently longer distance (~ 38 Å vs. 35 Å) between active sites in the RSV<br />

dimer. However, understanding the coordinated cutting and joining reaction awaits three-dimensional<br />

information on the arrangement of monomers within an integrase multimer binding to DNA.<br />

E. Three-Dimensional <strong>Structure</strong>s of Other Domains of HIV-1 Integrase<br />

Three-dimensional structural information has not yet been obtained for a full-length integrase protein. In<br />

its absence, attempts have been made to determine the structure of the smaller domains consisting of the<br />

separately expressed N- and C-termini that flank the core whose structure is now known.<br />

The Amino Terminus of Integrase<br />

While the N-terminus of HIV-1 integrase, consisting of residues 1 to 55, has been separately expressed,<br />

purified, and biophysically characterized [31], structural data has not yet been obtained. This protein<br />

domain binds metal ions such as Zn 2+, Co 2+, and Cd 2+ stoichiometrically, and is monomeric at low<br />

protein concentrations. Dramatic changes in helix content (from 14% to 32%) are observed in the<br />

circular dichroism (CD) spectrum upon addition of metal. Analysis of CD spectral features led<br />

researchers to conclude that it is highly probably that integrase contains a zinc finger that folds in much<br />

the same way as the TFIIIA-like DNA binding proteins, with two His residues located on an α helix and<br />

two cysteines part of a β sheet [31]. However, confirmation of such a model awaits structure<br />

determination <strong>by</strong> x-ray crystallography or NMR spectroscopy.<br />

The Carboxy Terminus of Integrase<br />

When the C-terminal domain is expressed as a separate polypeptide, IN 213–288 can be purified from the<br />

initial soluble fraction from cell lysates [33]. This small protein fragment, therefore, was an attractive<br />

target for structure determination.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_102.html [4/5/2004 4:51:59 PM]


Document<br />

Figure 8<br />

Molscript stereo figure of the structure of the nonspecific DNA-binding domain<br />

of HIV-1 integrase, IN 220–270 , determined <strong>by</strong> heteronuclear NMR spectroscopy [28].<br />

Page 103<br />

The structure of this domain is of particular interest as it represents the dominant nonspecific DNAbinding<br />

region of integrase. Gel filtration and sedimentation equilibrium results indicated that purified<br />

IN 213–288 partitioned between dimers and highly aggregated material (unpublished observations).<br />

However, a smaller domain consisting of residues 220 to 270 maintains the DNA-binding properties of<br />

the longer C-terminal domain and is better behaved in solution. Recently, two groups have reported the<br />

structure of this smaller fragment, IN 220–270, determined using multidimensional heteronuclear NMR<br />

spectroscopic methods [28,29].<br />

As shown in Figure 8, the overall structure of IN 220–270, is that of a β sandwich formed <strong>by</strong> two threestranded<br />

β sheets. As anticipated <strong>by</strong> biophysical studies, the polypeptide is a dimer in solution. The<br />

interface between monomers is formed <strong>by</strong> the antiparallel interaction of three β strands from each<br />

subunit and is stabilized predominantly <strong>by</strong> hydrophobic interactions. There is a long loop between<br />

strands β1 and β2 which, in the dimer, defines the sides of a cleft that is of the appropriate dimensions<br />

(about 24 × 24 × 12 Å) to accommodate double-stranded DNA. The folding topology is very similar to<br />

that of SH3 domains that are found in several proteins involved in signal transduction, despite the lack<br />

of significant sequence homology. This is rather unusual since SH3 domains are generally involved in<br />

protein binding rather than interactions with DNA.<br />

V. Prospects for Inhibitors<br />

A. Overview of Inhibitor Studies to Date<br />

The investigation of HIV integrase inhibitors has been largely restricted to testing available compounds<br />

that inhibit other enzymes with similar substrates or<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_103.html (1 of 2) [4/5/2004 4:52:07 PM]


Document<br />

Table 1 Representative Inhibitors of HIV-1 Integrase and IC 50 values of compounds that inhibit 3' processing and<br />

strand transfer activities of HIV-1 integrase<br />

Compound<br />

I aurintricarboxylic acid<br />

monomer<br />

IC 50 of 3' processing<br />

(μM)<br />

IC 50 of strand<br />

transfer (μM)<br />

Page 104<br />

Active against<br />

IN 50–212 ? Reference<br />

10–50 n.d. n.d. 44<br />

II cosalane n.d. 25 n.d. 45<br />

III DHNQ 5.7 2.5 yes 47<br />

IV primaquine 15 3.6 n.d. 46<br />

V CAPE 220 19 yes * 46<br />

VI quercetin 24 14 n.d. 47<br />

VII quercetagetin 0.8 0.1 yes 47<br />

VIII AG1717 0.4 0.16 yes 50<br />

IX β-conidendrol 0.5 0.5 n.d. 51<br />

X suramin 0.25 0.11 n.d. 53<br />

XI curcumin 95 40 yes 55<br />

XII (neocuproine) 2-Cu + 3 3 yes 56<br />

XIII AZT-monoP i 100–150 100–150 yes 57<br />

XIV GT 17-mer 0.092 0.046 n.d. 59<br />

XV HCKFWW 2 2 yes 43<br />

The third column indicates if these compounds also inhibit disintegration <strong>by</strong> IN 50–212 . n.d. = not determined. * = only if<br />

pre-incubated.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_104.html (1 of 2) [4/5/2004 4:52:09 PM]


Document<br />

proposed mechanisms. For example, as will be seen, many topoisomerase II inhibitors also inhibit HIV<br />

integrase. While screening of chemical databases is likely underway at several pharmaceutical<br />

companies, results have either not been made available or are discouraging (see below). One foray into<br />

integrase inhibitor design—rather than discovery—has recently been described using a peptide<br />

combinatorial library approach [43]. We summarize below published reports to date (March 1996) in<br />

which compounds have been identified that inhibit integrase in in vitro assays with IC 50 values of 100<br />

μM or less (IC 50 is the concentration at which the measured activity is inhibited <strong>by</strong> 50%). In vitro<br />

inhibition data is compiled in Table 1; structures of selected compounds are shown in Figure 9.<br />

An Effective Pharmacophore: Multiple Hydroxyl Groups on Aromatic Rings<br />

The first report of a class of compounds that inhibits HIV integrase appeared in 1992 [44].<br />

Aurintricarboxylic acid (I) and its derivatives, known to inhibit other enzymes that process nucleic acids,<br />

were determined to inhibit 3' processing with moderate IC 50 values of 10–50 μM. As shown in Figure 9,<br />

a recurring structural theme was established early on in which integrase inhibitors often possess<br />

aromatic rings with multiple hydroxy substituents that are either located<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_104.html (2 of 2) [4/5/2004 4:52:09 PM]


Document<br />

Figure 9<br />

Chemical structures of reported inhibitors of HIV-1 integrase. (I) aurintricarboxylic<br />

acid monomer; (II) cosalane; (III) dihydronaphthoquinone or DHNQ; (IV)<br />

primaquine; (V) caffeic acid phenethyl ester or CAPE; (VI) quercetin; (VII)<br />

quercetagetin; (VIII) AG1717; (IX) β-conidendrol; (X) suramin; (XI) curcumin.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_105.html (1 of 2) [4/5/2004 4:52:14 PM]<br />

Page 105


Document<br />

Page 106<br />

on the same ring or can potentially be positioned close together in three-dimensional space if rings stack<br />

on top of each other. More recently it was shown that cosalane (II), a steroid-substituted derivative of<br />

(I), was no better in inhibiting integrase in vitro than the parent compound, but showed promise in HIV<br />

cytopathicity cell-culture assays [45]. Although cosalane and a number of related analogues inhibit both<br />

HIV protease and integrase in vitro, the primary site of action is believed to be inhibition of gp120<br />

binding to CD4 receptors.<br />

In 1993, the effects of selected topoisomerase II inhibitors, antimalarial agents, DNA binders,<br />

naphthoquinones, and various other agents on integrase activity in vitro were investigated [46].<br />

Although certain effective topoisomerase inhibitors are also good HIV integrase inhibitors, this is not a<br />

generalizable correlation. Since many topoisomerase inhibitors also bind DNA, it is difficult to assess<br />

whether the observed in vitro effects result from specific interaction with integrase or from the<br />

sequestering or distorting of the DNA substrate. However, several compounds were identified that are<br />

not known to be DNA binders but that inhibited integrase with reasonable IC 50 values. These included<br />

dihydroxynaphthoquinone (III), primaquine (IV), caffeic acid phenethyl ester (CAPE, V), and quercetin<br />

(VI).<br />

Motivated <strong>by</strong> the structural similarities between compounds III–VI, a more intensive structure-function<br />

study of flavones was undertaken in which approximately 50 related compounds were screened for<br />

inhibition of in vitro integrase activity [47]. Flavones are planar compounds containing three aromatic<br />

rings substituted with various polar groups such as hydroxy substituents. General trends were observed<br />

relating structure to inhibitory effectiveness; for example, inhibition required at least three hydroxy<br />

groups, the most favorable arrangement being when they were located ortho to one another. The most<br />

effective compound, quercetagetin (VII), is a potent topoisomerase II inhibitor and a known DNA<br />

intercalator. It was noted <strong>by</strong> the authors that many flavones are not integrase-specific; rather, they inhibit<br />

a broad range of enzymes including DNA polymerases, ATPases, and NAPDH-monooxygenases. They<br />

are also, in general, capable of DNA intercalation. It has not been established that their inhibitory effects<br />

are due to direct interactions with integrase.<br />

A subsequent detailed structure-activity relationship study of CAPE (V) revealed that while the ortho<br />

hydroxys were important for in vitro integrase inhibition, both the caffeic acid and phenethyl moieties<br />

could be substantially modified [48]. Ortho hydroxyl groups in the context of other classes of<br />

compounds also confer anti-integrase potency. For example, several semisynthetic compounds derived<br />

from arctigenin, a topoisomerase I inhibitor that itself is not active against integrase, have been shown to<br />

inhibit HIV integrase [49]. More compelling evidence of the importance of ortho hydroxyl groups was<br />

provided <strong>by</strong> the tyrphostins, a group of synthetic compounds that are tyrosine kinase inhibitors. Several<br />

of these also inhibit integrase in the submicromolar range<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_106.html [4/5/2004 4:52:16 PM]


Document<br />

(for example, compound VIII, aka AG1717). In cell-<strong>based</strong> screening assays, AG1717 demonstrated<br />

some antiviral activity [50].<br />

Page 107<br />

Another study that demonstrated a role for ortho hydroxyl groups in in vitro integrase inhibition<br />

identified β-conidendrol (IX) via random screening as an inhibitor with an IC 50 value of less than 1 μM<br />

[51]. Although β-conidendrol did not inhibit several other nucleic-acid processing enzymes, indicating<br />

some specificity for integrase, it was not active in cell-<strong>based</strong> antiviral assays at concentrations as high as<br />

100 μM.<br />

Other Classes of Integrase Inhibitors<br />

Several compounds and their derivatives that do not contain adjacent hydroxy groups on a phenyl ring<br />

have recently been identified as HIV integrase inhibitors. These include suramin (X), curcumin (XI),<br />

phenanthroline-Cu + complexes (XII), and 3'-azido-3'-deoxythymidylate (AZT) monophosphate (XIII).<br />

Suramin (X) is a known inhibitor of DNA and RNA polymerases, retroviral reverse transcriptases, and<br />

topoisomerase II. It has also been shown to prevent the infection of T lymphocytes <strong>by</strong> HIV in vitro [52].<br />

Its six sulfonic acid groups confer a strong negative charge, and it was reasoned that there might be an<br />

inhibitory electrostatic interaction with the positive residues of the HIV C-terminus domain. Suramin<br />

was shown to be an effective inhibitor of 3' processing and strand transfer, with IC 50 values of 0.25 μM<br />

and 0.11 μM, respectively [53]. It was not demonstrated, however, that the mechanism of inhibition does<br />

involve binding to the C-terminus of integrase, although this could be readily addressed using Cterminal<br />

truncated mutants active for disintegration.<br />

Curcumin (XI), the coloring dye in the spice turmeric, is structurally related to CAPE (V). It has also<br />

been shown to inhibit HIV replication <strong>by</strong> inhibiting p 24 antigen production and tat-mediated transcription<br />

[54]. As shown in Table 1, it also has moderate integrase inhibitory properties [55]. Although its two-<br />

OH groups are neither adjacent to each other nor on the same phenyl ring, its conformations can be<br />

modeled to bring the hydroxy groups into close proximity <strong>by</strong> stacking the two phenyl rings on each<br />

other.<br />

Several tetrahedral cuprous phenanthroline complexes, known inhibitors of transcription, were tested<br />

against integrase and shown to be reasonably effective inhibitors [56]: IC 50 values in the range of 1–10<br />

μM were determined (for example, the neocuproine-Cu + complex, XII). Analyses of the mode of<br />

inhibition demonstrated that these compounds act noncompetitively, and that inhibition does not<br />

correlate with inhibition of DNA binding. Thus, it has been proposed that these metal chelates may act<br />

at a site distant from the active site, or perhaps in the context of an enzyme-DNA complex.<br />

3'-Azido-3'-deoxythymidine, or AZT, is a nucleoside analog approved for use to treat AIDS. Its<br />

metabolites, the mono-, di- and triphosphate forms, accu-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_107.html (1 of 2) [4/5/2004 4:52:18 PM]


Document<br />

Page 108<br />

mulate during treatment; in particular, AZT-monoP i (XIII) accumulates in cells to millimolar levels.<br />

These metabolites were investigated as possible integrase inhibitors and it was shown that all three<br />

phosphate derivatives inhibit with IC 50 values of 100–150 μM, although AZT itself is not inhibitory [57].<br />

These results suggest that despite the weak inhibition <strong>by</strong> these particular compounds, nucleoside analogs<br />

may serve as lead compounds for the development of integrase inhibitors.<br />

It was recently observed that oligonucleotides that form guanosine quartet structures inhibit HIV<br />

replication [58]. In light of the AZT results that suggested that there may be a nucleotide binding site on<br />

integrase—and because integrase binds DNA—these oligonucleotides (for example, 5'-<br />

GTGGTGGGTGGGTGGGT-3', XIV) were investigated as possible inhibitors [59]. These compounds<br />

have the lowest inhibition constants reported to date (see Table 1), and suggest an exciting new avenue<br />

for integrase inhibitor development.<br />

In a completely different approach to integrase inhibitors, a synthetic peptide combinatorial library was<br />

used to select a hexapeptide capable of inhibiting integrase proteins [43]. The first two amino acids were<br />

selected using a library of 400 dipeptides, and the remaining amino acids selected one-<strong>by</strong>-one in an<br />

iterative process. The optimal hexapeptide, HCKFWW (XV), inhibits HIV-1 integrase with an IC 50 of 2<br />

μM. The peptide does not compete for DNA binding to viral DNA, nor does it represent a sequence<br />

present in integrase itself. Although it is not expected that a peptide consisting of L-amino acids would<br />

be a suitable drug in itself, the use of D-amino acids or peptidomimetic backbones may be fruitful<br />

directions to pursue.<br />

<strong>Summary</strong><br />

Many of the compounds identified to date that inhibit HIV integrase in vitro have common structural<br />

features as illustrated in Figure 9. Most notably, these include hydroxy-substitued phenyl rings.<br />

However, these compounds may inhibit in vitro activity for reasons unrelated to enzyme binding. For<br />

example, it is difficult to know what to make of inhibition studies where the compounds added to the<br />

assays are known to bind DNA. Do the compounds affect in vitro activity because they bind and<br />

sequester the substrate? It is possible that they distort the DNA or intercalate between base pairs,<br />

preventing appropriate binding to the enzyme. While this is itself is a valid basis for the design of<br />

inhibitors against HIV infection, particularly if it targets DNA specific to the virus such as the LTR<br />

sequences [60], it is not an approach that builds on knowledge of the three-dimensional structure of the<br />

protein. To this end, valuable information could be obtained from studies in which direct binding of<br />

compounds to integrase is measured. This is also true for those inhibitors that do not contain the hydroxysubstituted<br />

phenyl pharmacophore. Binding studies could be carried<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_108.html [4/5/2004 4:52:19 PM]


Document<br />

Page 109<br />

out, for example, using radiolabeled inhibitors or possibly <strong>by</strong> monitoring any UV-spectral shifts in the<br />

case of aromatic compounds.<br />

B. What We Need to Know before We can Start <strong>Structure</strong>-Based <strong>Drug</strong> <strong>Design</strong><br />

The review of known integrase inhibitors presented in the previous section demonstrates the paucity of<br />

effective inhibitors reported in the literature. A limited number of structural types have been<br />

investigated, focused heavily on compounds with aromatic ring systems and hydroxy substituents. Most<br />

inhibitors reported to date are only moderately effective in in vitro assays, with IC 50 values residing in<br />

the low μM range (see Table 1).<br />

To build on this knowledge base, and to use known molecules as lead compounds for the development<br />

of more effective inhibitors, it would be extremely valuable to obtain a high-resolution crystal structure<br />

of an integrase-inhibitor complex. We and others are vigorously pursuing this goal. It is not clear if the<br />

lack of success to date is because the inhibitors identified so far manifest their effects in in vitro assays<br />

predominantly at the level of the DNA, or if some aspect of the structure of the HIV-1 integrase core<br />

itself—for example, its mobility in certain regoins—prevents the formation of a tight complex. It is also<br />

possible that binding is weaker to the HIV-1 core domain than to the full-length protein because of<br />

missing enzyme-inhibitor contacts. Co-crystallization attempts would benefit from in vitro studies to<br />

determine relative binding constants as a guide in selecting the most tightly bound inhibitors. It would<br />

also be useful to obtain information on the effect of variables, such as Mn 2+ or Mg 2+, on binding<br />

constants.<br />

It may be futile at this stage to attempt to model the binding of known inhibitors to the catalytic core<br />

domain of HIV-1 integrase in the absence of more complete information. It cannot a priori be assumed<br />

that the site of action of all these inhibitors is the enzyme active site identified <strong>by</strong> the constellation of<br />

conserved acidic residues. For example, certain very effective nonnucleoside inhibitors of HIV reverse<br />

transcriptase bind not to the enzyme active site, but rather to a small pocket adjacent to it. There are no<br />

obvious structural features of the integrase core—such as the deep trough surrounding active site<br />

residues in the case of HIV protease [61,62]—that can be readily identified as a potential inhibitor<br />

binding site. Furthermore, since part of the region defining the active site of HIV-1 integrase is<br />

disordered in our crystal structure, this prevents a surface rendering of the region around the conserved<br />

acidic residues. It also seems unlikely, given the known differences in active-site geometries between<br />

the HIV and ASV integrase (see Section IV.B), that the structure of the related ASV integrase core<br />

domain <strong>by</strong> itself would be particularly useful in this regard. As the active form of the enzyme<br />

presumably binds divalent metal ions, it will be important to determine how or if the structure of HIV<br />

integrase changes<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_109.html [4/5/2004 4:52:21 PM]


Document<br />

when metals are bound. Finally, there is no evidence to rule out the possibility that the two termini<br />

contribute to part of the active site, occlude some of it, or restrict access to it in as yet undetermined<br />

ways. For this and other obvious reasons, three-dimensional structures of larger versions of HIV-1<br />

integrase, such as IN 1–212, IN 50–288, and the full-length protein, IN 1–288, will be required.<br />

Page 110<br />

We also lack a clear picture of how the enzyme substrates, the viral DNA ends and the target DNA, bind<br />

to integrase. The DNA must at some point approach the region defined <strong>by</strong> the three conserved acidic<br />

residues so that bond cleavage and joining can occur. However, the dominant DNA binding domain is<br />

defined <strong>by</strong> residues in the C-terminus. It would be extremely valuable to determine the relative<br />

orientation of these domains in the context of a larger version of integrase. Even more revealing would<br />

be the structure of the full-length protein with bound DNA. Once we possess this information, it should<br />

be possible to rapidly progress with structure-<strong>based</strong> drug design.<br />

C. Possible Approaches to the <strong>Design</strong> of Effective Integrase Inhibitors<br />

There are a variety of approaches to the design of integrase inhibitors that are obvious and do not depend<br />

on knowing its three-dimensional structure. However, the rational implementation and refinement of<br />

these approaches will require high-resolutional structural data, much of which, as indicated above, is not<br />

yet available. It still may be useful to discuss here different classes of inhibitors that can be envisioned.<br />

Preventing DNA Binding<br />

One approach to the inhibition of integrase would be to prevent binding of the DNA substrate.<br />

Unfortunately, we do not yet know how or where DNA binds. There are likely to be several sites on the<br />

enzyme that contact DNA, including the C-terminus and the region around the active site. In the absence<br />

of the structure of an integrase-DNA complex, structures of related enzymes (RNase H, the MuA<br />

transposase, and RuvC) with their DNA substrates would be useful guides in suggesting ways in which<br />

DNA could interact with integrase. However, there is no guarantee that modes of DNA binding are<br />

conserved among members of this polynucleotidyl transferase family. The overall structure of the Cterminus<br />

fragment suggests that it should be possible to develop compounds that bind specifically in the<br />

cleft formed <strong>by</strong> dimers of IN 220–270 (see Section IV.E) which may prevent DNA binding.<br />

Inhibition at the Active Site<br />

It may be possible to inhibit integrase <strong>by</strong> preventing the binding of the required metal cofactor(s) to the<br />

active site acidic residues that are presumed to provide<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_110.html [4/5/2004 4:52:23 PM]


Document<br />

Page 111<br />

coordinating ligands to the metal(s). This could be accomplished <strong>by</strong> sterically blocking access to the<br />

active site or <strong>by</strong> specifically binding the acidic residues themselves. Such a mechanism has been<br />

suggested to explain the inhibition of integrase <strong>by</strong> curcumin [55], which could bind to the active site<br />

aspartates or glutamates via its hydroxy groups. Compounds could also be devised that chelate the<br />

metal(s) once bound, preventing access of the active site to the substrate or distorting the active site<br />

geometry preventing phosphate bond cleavage. An intriguing approach to disrupting metal binding has<br />

recently been reported for the Zn 2+-binding HIV nucleocapsid (NC) protein [63]. In this case,<br />

compounds were developed that specifically eject Zn 2+ from the zinc-finger region of NC, interfering<br />

with viral replication. Such an approach might be applied to Mn 2+ or Mg 2+ binding at the active site,<br />

although the Zn 2+-binding domain at the N terminus of integrase also suggests itself as a target.<br />

It is curious that approaches have not been devised for mechanism-<strong>based</strong> inhibitors, particularly since<br />

the stereochemical mechanism of integration has been understood for some time now [8].<br />

Interfering with Multimerization<br />

The structures of the domains of HIV-1 integrase determined to date both reveal dimers [3,28,29,36]. It<br />

may be possible to develop compounds that bind specifically to the dimer interfaces, preventing<br />

interactions between monomers that may be necessary for activity. The success of this approach<br />

presumes that during the retroviral lifecycle there is a point where the monomer surfaces are accessible.<br />

It is not clear that integrase in the context of preintegration complexes is in equilibrium between the<br />

monomeric and higher order forms. This approach need not be restricted to preventing dimer formation<br />

if higher order interactions (e.g. formation of a tetramer) are also mechanistically relevant. To this end,<br />

the structure of an integrase tetramer, such as that formed <strong>by</strong> the full-length protein, would be useful in<br />

identifying dimer-dimer interface(s).<br />

Other Ways to Confound Integrase<br />

There are other parts of the retroviral life cycle involving integrase that could be targeted for inhibition.<br />

For example, integrase can bind other proteins such as human Ini1 [64], and most likely interacts with<br />

the viral proteins that are part of the preintegration complex [65]. Although it is not known if these<br />

interactions are essential for viral replication, preventing protein-protein binding would provide another<br />

site at which to attack integrase. It is possible that interactions between integrase and other proteins in<br />

the preintegration complex are crucial for maintaining the integrity of this complex and its ability to<br />

migrate to the cell nucleus. Interfering with these protein-protein or protein-nucleic acid interactions<br />

may be another approach to halting viral replication. For example, prein-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_111.html [4/5/2004 4:52:24 PM]


Document<br />

cubation of phosphotyrosine with integrase has been shown to inhibit interaction with the MA protein<br />

[65]. Therefore, phosphotyrosine analogs could be a unique approach to antiviral development.<br />

D. Concluding Statement<br />

Page 112<br />

It is known from drug design studies with HIV reverse transcriptase and protease that the virus is able to<br />

escape from the pressures of inhibitors <strong>by</strong> mutation of the drug targets [66]. Although integrase is, a<br />

priori, a reasonable target for drug-design efforts, it must be anticipated that integrase will also be able<br />

to rapidly mutate and there<strong>by</strong> avoid inhibition.<br />

By analogy with recent approaches to reverse transcriptase inhibitors, it may be possible to design a set<br />

of integrase inhibitors that act at slightly different binding sites and from which the virus cannot<br />

simultaneously escape. That is, the combination of mutations required to avoid inhibition may be severe<br />

enough to prevent integrase from carrying out its required chemistry. (In the absence of direct structural<br />

information on the sites of inhibitor binding to integrase, the generation of escape mutants in vitro and<br />

their subsequent sequencing may be an indirect way to identify inhibitor binding sites.) It will also be<br />

important to determine if the virus will be able to simultaneously mutate reverse transcriptase, integrase,<br />

and the protease in response to a combination of inhibitors targeted against all three pol gene products.<br />

The answers to these questions will require the development of suitable inhibitors and the beginning of<br />

in vitro testing. To this end—while large-scale screening and the development of combinatorial<br />

chemistry methods should continue—the structure of the catalytic core domain of HIV-1 integrase is a<br />

starting point for the rational design of integrase inhibitors. There is much more structural information<br />

that must be obtained for the full-length protein and its complexes with metals, inhibitors, and<br />

substrates. We and others are aggressively pursuing results on these fronts.<br />

E. Recent Developments<br />

Several studies published since March 1996 have expanded the list of in vitro integrase inhibitors<br />

effective at IC 50 values below 100 μM. These include two dicaffeoylquinic acids obtained from<br />

medicinal plants and a synthetic analog, L-chicoric acid [68], the HIV protease inhibitors NSC 117027<br />

and NSC 158393 [69], certain anthraquinone derivatives [70], coumermycin, and pyridoxal phosphate<br />

[71]. In addition to exhibiting in vitro inhibition, the dicaffeoylquinic acids effectively inhibited HIV-1<br />

replication in T-lymphoblastoid cell lines [68].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_112.html [4/5/2004 4:52:26 PM]


Document<br />

Page 113<br />

Follow-up studies were also reported for two previously identified classes of integrase inhibitors.<br />

Several nucleotides that were more effective inhibitors than the originally tested AZT nucleotides were<br />

identified [71]. For example, the L-enantiomers of 5-fluoro-2',3'-dideoxycytidine monophosphate and<br />

triphosphate inhibit 3' processing and strand transfer with IC 50 values of ~40 μM. A structure-function<br />

study on GT-containing oligonucleotides showed that both the number of quartets formed and the loop<br />

sequences between the quartets are important for activity, and that inhibitors of this type may function<br />

<strong>by</strong> interacting with the N-terminus of integrase [72].<br />

A particularly important contribution was the demonstration that preintegration complexes isolated from<br />

HIV-infected lymphoid cells can be used in assays to screen for inhibition of integration [70].<br />

Intriguingly, many compounds previously identified as in vitro inhibitors of 3' processing and strand<br />

transfer had no effect on integration carried out <strong>by</strong> either crude or partially purified preintegration<br />

complexes. Thus, such an assay may be a valuable method of screening out “false positives” identified<br />

using in vitro oligonucleotide assays, or corroborating the evidence that a particular compound is indeed<br />

active against integrase.<br />

Acknowledgments<br />

Our work described here was carried out in the laboratories of R. Craigie and D. R. Davies. We express<br />

our gratitude to our co-workers who, over the years, participated in the effort to determine the structure<br />

of HIV integrase. In particular, we acknowledge the contributions of F. D.Bushman, M. Carmichael, A.<br />

Engelman, S. Hosseini, T. Jenkins, K. Mizuuchi, I. Palmer, P. Rice, P. Sun, and P. Wingfield. We would<br />

also like to thank D. R. Davies and T. Jenkins for their comments on the manuscript and A. Mazumder<br />

for alerting us to the most recent work on integrase inhibitors and for his contributions to Section V.A.<br />

References<br />

1. AIDS WEEKLY Plus. Key KK, ed. Atlanta, Charles Henderson Publisher, 1996: Feb. 5 & 12, p. 14.<br />

2. Cara A, Guarnaccia F, Reitz, Jr. MS, Gallo RC, Lori F. Self-limiting, cell type-dependent replication<br />

of an integrase-defective human immunodeficiency virus type 1 in human primary macrophages but not<br />

T lymphocytes. Virol 1995, 208:242–248.<br />

3. Dyda F, Hickman AB, Jenkins TM, Engelman A, Craigie R, Davies DR. Crystal structure of the<br />

catalytic domain of HIV-1 integrase: Similarity to other polynucleotidyl transferases. Science 1994;<br />

266:1981–1986.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_113.html [4/5/2004 4:52:27 PM]


Document<br />

4. Vink C, Plasterk RHA. The human immunodeficiency virus integrase protein. Trends Genet 1993;<br />

9:433–437.<br />

Page 114<br />

5. Katz RA, Skalka AM. The retroviral enzymes. Annu Rev Biochem 1994; 63:133–173.<br />

6. Sherman PA, Fyfe JA. Human immunodeficiency virus integration protein expressed in Escherichia<br />

coli possesses selective DNA cleaving activity. Proc Natl Acad Sci USA 1990; 87:5119–5123.<br />

7. Bushman FD, Craigie R. Activities of human immunodeficiency virus (HIV) integration protein in<br />

vitro: Specific cleavage and integration of HIV DNA. Proc Natl Acad Sci USA 1991; 88:1339–1343.<br />

8. Engelman A, Mizuuchi K, Craigie R. HIV-1 DNA integration: Mechanism of viral DNA cleavage<br />

and DNA strand transfer. Cell 1991; 67:1211–1221.<br />

9. Chow SA, Vincent KA, Ellison V, Brown PO. Reversal of integration and DNA slicing mediated <strong>by</strong><br />

integrase of human immunodeficiency virus. Science 1992; 255:723–726.<br />

10. Grandgenett DP, Vora AC, Schiff RD. A 32,000-dalton nucleic acid-binding protein from avian<br />

retravirus cores possesses DNA endonuclease activity. Virol 1978; 89:119–132.<br />

11. Vincent KA, Ellison V, Chow SA, Brown PO. Characterization of human immunodeficiency virus<br />

type 1 integrase expressed in Escherichia coli and analysis of variants with amino-terminal mutations. J.<br />

Virol 1993; 67:425–437.<br />

12. Hickman AB, Palmer I, Engelman A, Craigie R, Wingfield P. Biophysical and enzymatic properties<br />

of the catalytic domain of HIV-1 integrase. J Biol Chem 1994; 269:29279–29287.<br />

13. Jones KS, Coleman J, Merkel GW, Laue TM, Skalka AM. Retroviral integrase functions as a<br />

multimer and can turn over catalytically. J Biol Chem 1992; 267:16037–16040.<br />

14. Engelman A, Bushman FD, Craigie R. Identification of discrete functional domains of HIV-1<br />

integrase and their organization within an active multimeric complex. EMBO J 1993; 12:3269–3275.<br />

15. Andrake MD, Skalka AM. Multimerization determinants reside in both the catalytic core and C<br />

terminus of avian sarcoma virus integrase. J Biol Chem 1995; 270:29299–29306.<br />

16. Kalpana GV, Goff SP. Genetic analysis of homomeric interactions of human immunodeficiency<br />

virus type 1 integrase using the yeast two-hybrid system. Proc Natl Acad Sci USA 1993;<br />

90:10593–10597.<br />

17. Van Gent DC, Vink C, Oude Groeneger AAM, Plasterk RHA. Complementation between HIV<br />

integrase proteins mutated in different domains. EMBO J 1993; 12:3261–3267.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_114.html (1 of 2) [4/5/2004 4:52:29 PM]


Document<br />

18. Johnson MS, McClure MA, Feng D–F, Gray J, Doolittle RF. Computer analysis of retroviral pol<br />

genes: Assignment of enzymatic functions to specific sequences and homologies with nonviral enzymes.<br />

Proc Natl Acad Sci USA 1986; 83:7648–7652.<br />

19. Engelman A,<br />

Craigie R.<br />

Identification of<br />

conserved amino<br />

acid residues<br />

critical for human<br />

immunodeficiency<br />

virus type 1<br />

integrase function<br />

in vitro. J Virol<br />

1992;<br />

66:6361–6369.<br />

20. Van Gent DC, Oude Groeneger AAM, Plasterk RHA. Mutational analysis of the integrase protein of<br />

human immunodeficiency virus type 2. Proc Natl Acad Sci USA 1992; 89:9598–9602.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_114.html (2 of 2) [4/5/2004 4:52:29 PM]


Document<br />

Page 115<br />

21. Kulkosky J, Jones KS, Katz RA, Mack JPG, Skalka AM. Residues critical for retroviral integrative<br />

recombination in a region that is highly conserved among retroviral/retrotransposon integrases and<br />

bacterial insertion sequence transposases. Mol Cell Biol 1992; 12:2331–2338.<br />

22. Bushman FD, Engelman A, Palmer I, Wingfield P, Craigie R. Domains of the integrase protein of<br />

human immunodeficiency virus type 1 responsible for polynucleotidyl transfer and zinc binding. Proc<br />

Natl Acad Sci USA 1993; 90:3428–3432.<br />

23. Beese LS, Steitz TA. Structural basis for the 3'-5' exonuclease activity of Escherichia coli DNA<br />

polymerase I: a two metal ion mechanism. EMBO J 1991; 10:25–33.<br />

24. Woerner AM, Marcus-Sekura CJ. Characterization of a DNA binding domain in the C-terminus of<br />

HIV-1 integrase <strong>by</strong> deletion mutagenesis. Nucl Acids Res 1993; 21:3507–3511.<br />

25. Vink C, Oude Groeneger AAM, Plasterk RHA. Identification of the catalytic and DNA-binding<br />

region of the human immunodeficiency virus type 1 integrase protein. Nucl Acids Res 1993;<br />

21:1419–1425.<br />

26. Puras Lutzke RA, Vink C, Plasterk RHA. Characterization of the minimal DNA-binding domain of<br />

the HIV integrase protein. Nucl Acids Res 1994; 22:4125–4131.<br />

27. Drelich M, Wilhelm R, Mous J. Identification of amino acid residues critical for endonuclease and<br />

integration activities of HIV-1 IN protein in vitro. Virol 1992; 188:459–468.<br />

28. Lodi PJ, Ernst JA, Kuszewski J, Hickman AB, Engelman A, Craigie R, Clore GM, Gronenborn AM.<br />

Solution structure of the DNA binding domain of HIV-1 integrase. Biochem 1995; 34:9826–9833.<br />

29. Eijkelenboom APAM, Puras Lutzke RA, Boelens R, Plasterk RHA, Kaptein R, Hard K. The DNAbinding<br />

domain of HIV-1 integrase has an SH3-like fold. Nature Struct Biol 1995; 2:807–810.<br />

30. Haugan IR, Nilsen BM, Worland S, Olsen L, Helland DE. Characterization of the DNA-binding<br />

activity of HIV-1 integrase using a filter binding assay. Biochem Biophys Res Commun 1995;<br />

217:802–810.<br />

31. Burke CJ, Sanyal G, Bruner MW, Ryan JA, LaFemina RL, Robbins HL, Zeft AS, Middaugh CR,<br />

Cordingley MG. Structural implications of spectroscopic characterization of a putative zinc finger<br />

peptide from HIV-1 integrase. J Biol Chem 1992; 267:9639–9644.<br />

32. Bushman FD, Wang B. Rous sarcoma virus integrase protein: Mapping functions for catalysis and<br />

substrate binding. J Virol 1994; 68:2215–2223.<br />

33. Engelman A, Hickman AB, Craigie R. The core and carboxyl-terminal domains of the integrase<br />

protein of human immunodeficiency virus type 1 each contribute to nonspecific DNA binding. J Virol<br />

1994; 68:5911–5917.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_115.html (1 of 2) [4/5/2004 4:52:31 PM]


Document<br />

34. Jenkins TM, Hickman AB, Dyda F, Ghirlando R, Davies DR, Craigie R. Catalytic domain of human<br />

immunodeficiency virus type 1 integrase: Identification of a soluble mutant <strong>by</strong> systematic replacement<br />

of hydrophobic residues. Proc Natl Acad Sci USA 1995; 92:6057–6061.<br />

35. Orengo CA, Thornton JM. Alpha plus beta fold revisited: some favoured motifs. <strong>Structure</strong> 1993;<br />

1:105–120.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_115.html (2 of 2) [4/5/2004 4:52:31 PM]


Document<br />

44. Cushman M, Sherman P. Inhibition of HIV-1 integration protein <strong>by</strong> aurintricarboxylic acid<br />

monomers, monomer analogs, and polymer fractions. Biochem Biophys Res Commun 1992; 185:85–90.<br />

45. Cushman M, Golebiewski WM, Pommier Y, Mazumder A, Reymen D, De Clercq E, Graham L,<br />

Rice WG. Cosalane analogues with enhanced potencies as inhibitors of HIV-1 protease and integrase. J<br />

Med Chem 1995; 38:443–452.<br />

46. Fesen MR, Kohn KW, Leteurte F, Pommier Y. Inhibitors of human immunodeficiency virus<br />

integrase. Proc Natl Acad Sci USA 1993; 90:2399–2403.<br />

47. Fesen MR, Pommier Y, Leteurtre F, Hiroguchi S, Yung J, Kohn KW. Inhibition of HIV-1 integrase<br />

<strong>by</strong> flavones, caffeic acid phenethl ester (CAPE) and related compounds. Biochem Pharmacol 1994;<br />

48:595–608.<br />

48. Burke Jr TR, Fesen MR, Mazumder A, Wang J, Carothers AM, Grunberger D, Driscoll J, Kohn K,<br />

Pommier Y. Hydroxylated aromatic inhibitors of HIV-1 integrase. J Med Chem 1995; 38:4171–4178.<br />

49. Eich E, Pertz H, Kaloga M, Schulz J, Fesen MR, Mazumder A, Pommier Y. (-)-Arctigenin as a lead<br />

structure for inhibitors of human immunodeficiency virus type-1 integrase. J Med Chem 1996;<br />

39:86–95.<br />

50. Mazumder A, Gazit A, Levitzki A, Nicklaus M, Yung J, Kohlhagen G, Pommier Y. Effects of<br />

tyrphostins, protein kinase inhibitors, on human immunodeficiency virus type 1 integrase. Biochem<br />

1995; 34:15111–15122.<br />

51. LaFemina RL, Graham PL, LeGrow K, Hastings JC, Wolfe A, Young SD, Emini EA, Hazuda DJ.<br />

Inhibition of human immunodeficiency cirus integrase <strong>by</strong> bis-catechols. Antimicrob Agents Chemother<br />

1995; 39:320–324.<br />

52. Mitsuya H, Popovic M, Yarchoan R, Matsushita S, Gallo RC, Broder S. Suramin protection of T<br />

cells in vitro against infectivity and cytopathic effect of HTLV-III. Science 1984; 226:172–174.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_116.html (2 of 2) [4/5/2004 4:52:33 PM]


Document<br />

Page 117<br />

53. Carteau S, Mouscadet JF, Goulaouic H, Subra F, Auclair C. Inhibitory effect of the polyanionic drug<br />

suramin on the in vitro HIV DNA integration reaction. Arch Biochem Biophys 1993; 305:606–610.<br />

54. Li CJ, Zhang LJ, Dezube BJ, Crumpacker CS, Pardee AB. Three inhibitors of type 1<br />

immunodeficiency virus long terminal repeat-directed gene expression and virus replication. Proc Natl<br />

Acad Sci USA 1993; 90:1839–1842.<br />

55. Mazumder A, Raghavan K, Weinstein J, Kohn KW, Pommier Y. Inhibition of human<br />

immunodeficiency virus type-1 integrase <strong>by</strong> curcumin. Biochem Pharmacol 1995; 49:1165–1170.<br />

56. Mazumder A, Gupta M, Perrin DM, Sigman DS, Rabinovitz M, Pommier Y. Inhibition of human<br />

immunodeficiency virus type 1 integrase <strong>by</strong> a hydrophobic cation: The phenanthroline-cuprous<br />

complex. AIDS Res Human Retro 1995; 11:115–125.<br />

57. Mazumder A, Cooney D, Agbaria R, Gupta M, Pommier Y. Inhibition of human immunodeficiency<br />

virus type 1 integrase <strong>by</strong> 3' -azido-3' -deoxythymidylate. Proc Natl Acad Sci USA 1994; 91:5771–5775.<br />

58. Ojwang J, Elbaggari A, Marshall HB, Jayaraman K, McGrath MS, Rando RF. Inhibition of human<br />

immunodeficiency virus type 1 activity in vitro <strong>by</strong> oligonucleotides composed entirely of guanosine and<br />

thymidine. J Acqu Immune Defic Syn 1994; 7:560–570.<br />

59. Ojwang JO, Buckheit RW, Pommier Y, Mazumder A, de Vreese K, Esté JA, Reymen D, Pallansch<br />

LA, Lackman-Smith C, Wallace TL, de Clercq E, McGrath MS, Rando RF. T30177, an oligonucleotide<br />

stabilized <strong>by</strong> an intramolecular guanosine octet, is a potent inhibitor of laboratory strains and clinical<br />

isolates of human immunodeficiency virus type 1. Antimicrob Agents Chemother 1995; 39:2426–2435.<br />

60. Carteau S, Mouscadet JF, Goulaouic H, Subra F, Auclair C. Inhibition of the in vitro integration of<br />

Moloney murine leukemia virus DNA <strong>by</strong> the DNA minor groove binder netropsin. Biochem Pharmacol<br />

1994; 47:1821–1826.<br />

61. Navia MA, Fitzgerald PMD, McKeever BM, Leu C-T, Heimbach JC, Herber WK, Sigal IS, Darke<br />

PL, Springer JP. Three-dimensional structure of aspartyl protease from human immunodeficiency virus<br />

HIV-1. Nature 1989; 337:615–620.<br />

62. Wlodawer A, Erickson JW. <strong>Structure</strong>-<strong>based</strong> inhibitors of HIV-1 protease. Annu Rev Biochem 1993;<br />

62:543–585.<br />

63. Rice WG, Supko JG, Malspeis L, Buckheit Jr RW, Clanton D, Bu M, Graham L, Schaeffer CA,<br />

Turpin JA, Domagala J, Gogliotti R, Bader JP, Halliday SM, Coren L, Sowder II RC, Arthur LO,<br />

Henderson LE. Inhibitors of HIV nucleocapsid protein zinc fingers as candidates for the treatment of<br />

AIDS. Science 1995; 270:1194–1197.<br />

64. Kalpana GV, Marmon S, Wang W, Crabtree GR, Goff SP. Binding and stimulation of HIV-1<br />

integrase <strong>by</strong> a human homolog of yeast transcription factor SNF5. Science 1994; 266:2002–2006.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_117.html (1 of 2) [4/5/2004 4:56:01 PM]


Document<br />

65. Gallay P, Swingler S, Song J, Bushman F, Trono D. HIV nuclear import is governed <strong>by</strong> the<br />

phosphotyrosine-mediated binding of matrix to the core domain of integrase. Cell 1995; 83:569–576.<br />

66. Coffin JM. HIV population dynamics in vivo: Implications for genetic variation, pathogenesis, and<br />

therapy. Science 1995; 267:483–489.<br />

67. Williams KJ, Loeb LA. Retroviral reverse transcriptases: Error frequencies and mutagenesis. Curr<br />

Top Microbiol Immunol 1992; 176:165–180.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_117.html (2 of 2) [4/5/2004 4:56:01 PM]


Document<br />

68. Robinson Jr WE, Reinecke MG, Abdel-Malek S, Jia Q, Chow SA. Inhibitors of HIV-1 replication<br />

that inhibit HIV integrase. Proc Natl Acad Sci USA 1996; 93:6326–6331.<br />

Page 118<br />

69. Mazumder A, Wang S, Neamati N, Nicklaus M, Sunder S, Chen J, Milne GWA, Rice WG, Burke Jr<br />

TR, Pommier Y. Antiretroviral agents as inhibitors of both human immunodeficiency virus type 1<br />

integrase and protease. J Med Chem 1996; 39;2472–2481.<br />

70. Farnet CM, Wang B, Lipford JR, Bushman FD. Differential inhibition of HIV-1 preintegration<br />

complexes and purified integrase protein <strong>by</strong> small molecules. Proc Natl Acad Sci USA 1996;<br />

93:9742–9747.<br />

71. Mazumder A, Neamati N, Sommadossi J, Gosselin G, Schinazi RF, Imbach J, Pommier Y. Effects of<br />

nucleotide analogues on human immunodeficiency virus type 1 integrase. Mol Pharmacol 1996;<br />

49:621–628.<br />

72. Mazumder A, Neamati N, Ojwang JO, Sunder S, Rando RF, Pommier Y. Inhibition of the human<br />

immunodeficiency virus type 1 integrase <strong>by</strong> guanosine quartet structures. Biochem 1996;<br />

35:13762–13771.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_118.html [4/5/2004 4:56:03 PM]


Document<br />

4<br />

Bradykinin Receptor Antagonists<br />

Donald J. Kyle<br />

Scios Nova Inc., Sunnyvale, California<br />

I. Introduction<br />

Page 119<br />

The term “kinins” is generally made in reference to either the nonapeptide bradykinin (Arg1-Pro2-Pro3- Gly4-Phe5-Ser6-Pro7-Phe8-Arg9) or the decapeptide kallidin (Lys1-Arg2-Pro3-Pro4-Gly5-Phe6-Ser7-Pro8- Phe9-Arg10). In rats another kinin, Ile1-Ser2-Arg3-Pro4-Pro5-Gly6-Phe7-Ser8-Pro9-Phe10-Arg11 (T-kinin) is<br />

produced under certain circumstances and binds to the same receptors as bradykinin [1,2]. A schematic<br />

of the human kinin-kallikrein system is shown in Figure 1.<br />

The release of kinins from precursor proteins (known as kininogens) is mediated <strong>by</strong> enzymes called<br />

kininogenases [3–5]. The predominant enzymes responsible are kallikreins, but others, which include<br />

trypsin, plasmin, and some snake venoms, also release kinins. Kininogens are primarily synthesized in<br />

the liver and represent an abundant source of the precursors that are required for kinin generation. These<br />

proteins are produced from alternative splicing of a single gene product and there are two forms: high<br />

molecular weight kininogen (HMWK) and low molecular weight kininogen (LMWK) [6]. Unlike<br />

HMWK, which exists in the circulation as a complex with plasma pre-kallikrein, LMWK circulates<br />

freely.<br />

During immunological reaction, charged surfaces—which may be derived from bacterial<br />

lipopolysaccharide, oligosaccharides, connective tissue proteoglycans, or damaged basement<br />

membranes—facilitate the conversion of factor XII to factor XIIa. Once factor XIIa is present, prekallikrein<br />

can be cleaved to its active form, known as plasma kallikrein. This enzyme acts upon its<br />

preferred substrate, HMWK, to release bradykinin. Plasma kallikrein is further able to convert inactive<br />

factor XII to active XIIa, there<strong>by</strong> participating in a positive<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_119.html [4/5/2004 4:56:06 PM]


Document<br />

Figure 1<br />

Diagram of the human kinin-kallikrein system including the native ligands for B1 and B2<br />

receptor subtypes.<br />

feedback loop. The cleavage of bradykinin from HMWK is highly localized since pre-kallikrein and<br />

substrate (HMWK) circulate as a complex.<br />

Page 120<br />

Another kinin, Lys-bradykinin (also known as kallidin), is produced via the action of the tissuekallikrein<br />

enzyme on LMWK. This enzyme is found in many tissues, either in the form of a precursor<br />

requiring activation or as an active enzyme. In contrast to plasma kallikrein, which preferentially acts<br />

upon HMWK, tissue kallikrein can release kallidin from either HMWK or LMWK. Through the action<br />

of aminopeptidases, kallidin can subsequently be converted directly into bradykinin. This enzyme is<br />

present in both the plasma and on the surface of epithelial cells.<br />

Both bradykinin and kallidin can be degraded <strong>by</strong> a variety of plasma and cell surface enzymes<br />

(kininases) [7]. The most widely recognized of these enzymes are kininase I, kininase II (angiotensin<br />

converting enzyme, ACE), and carboxypeptidase N. In plasma, kininase I cleaves the C-terminal<br />

arginine from both bradykinin and kallidin to form [des-Arg 9] kinins. These [des-Arg 9] kinins are known<br />

to act as agonists of B1 receptors that are present in some species and have been implicated in the<br />

pathophysiology associated with prolonged inflammation [8–10].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_120.html (1 of 2) [4/5/2004 4:56:27 PM]


Document<br />

Page 121<br />

Nearly all cells express kinin receptors that mediate the activities of both bradykinin and kallidin. The<br />

activation of these G-protein coupled receptors causes relaxation of venular smooth muscle and<br />

hypotension, increased vascular permeability, contraction of smooth muscle of the gut and airway<br />

leading to increased airway resistance, stimulation of sensory neurons, alteration of ion secretion of<br />

epithelial cells, production of nitric oxide, release of cytokines from leukocytes, and the production of<br />

eicosanoids from various cell types [11,12]. Because of this broad spectrum of activity, kinins have been<br />

implicated as an important mediator in many pathophysiologies including pain, sepsis, asthma,<br />

rheumatoid arthritis, pancreatitis, and a wide variety of other inflammatory diseases. Moreover, a recent<br />

report demonstrated that bradykinin B2 receptors on the surface of human fibroblasts were upregulated<br />

three-fold beyond normal in patients with Alzheimer's disease, implicating bradykinin as a participant in<br />

the peripheral inflammatory processes associated with that disease [13].<br />

In contrast to the adverse physiologies associated with bradykinin release, there is a growing body of<br />

literature that implicates bradykinin as a protective agent during periods of cardiac or renal stress<br />

[14–16]. In this regard there is substantial evidence that the cardioprotective effects afforded <strong>by</strong> ACEinhibitor<br />

treatment are a result of metabolically preserving bradykinin and are therefore mediated <strong>by</strong><br />

bradykinin B2 (and possibly B1) receptors [17–18]. These results point to a possible therapeutic role for<br />

a kinin receptor agonist.<br />

Overall, the kinins are an important part of a well-organized physiological system. The various aspects<br />

and interdependencies of the kinin system have been, and continue to be, the focus of intensive research<br />

efforts in many laboratories. Many pharmaceutical companies have identified this system as an ideal site<br />

for therapeutic intervention in many inflammatory diseases. Hence, there have been many diverse<br />

approaches taken toward the discovery of antagonists (peptide and nonpeptide) of B2 and B1 receptors.<br />

This review focuses on the structure-<strong>based</strong> design strategies pursued in our laboratories during the past<br />

several years.<br />

II. Ligand-Based Investigations<br />

A. The Solution Conformation of Bradykinin<br />

In the late 1980s when we began the pursuit of bradykinin receptor antagonists, information of relevance<br />

to medicinal chemists was scarce. For example, not one nonpeptide antagonist of this receptor was<br />

known, nor were any series upon which to base a structure-activity relationship. Moreover, all<br />

publications described bradykinin as being highly flexible in an aqueous environment, such that no<br />

structural mimetics could be rationalized. Of course the receptors had not been cloned at that time so<br />

nothing was known about the primary sequence of the receptor or the three-dimensional structure.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_121.html [4/5/2004 4:56:29 PM]


Document<br />

Page 122<br />

Initially our approach was to complete a detailed examination of bradykinin using two-dimensional<br />

NMR methods in combination with empirical energy calculations [19]. Our strategy was derived on the<br />

basis of spectral data, biological results from conformationally restricted analogs, as well as the<br />

relationship between ordering in bradykinin and the dielectric environment of the solvent. Our guiding<br />

hypothesis was that, although in aqueous solution bradykinin is conformationally random, the<br />

biologically active form of the peptide is likely ordered and stabilized within the lipid bilayer of the cell<br />

membrane prior to binding with its receptor. Alternatively, the receptor binding environment might also<br />

be hydrophobic and there<strong>by</strong> lead to similar conformational biases in the ligand. We presumed that an<br />

appropriate solvent environment should be able to stimulate, at least in terms of hydrophobicity and<br />

dielectric constant, the nature of a cell membrane, and a 90:10 d 8-dioxane-H 2O mixture was selected for<br />

NMR experiments. It was anticipated that under these nonsolvating conditions the conformational<br />

diversity of bradykinin might be severely restricted. The ultimate analysis of the two-dimensional NMR<br />

data collected at 500 MHz supported a single major conformational species. There were five HN-CαH<br />

connectivities, one for each amide. This was confirmed in the 13C NMR spectrum where only nine<br />

carbonyl resonances, one for each amino acid, were present.<br />

In order to provide a starting point for subsequent molecular dynamics simulations the assumption was<br />

made, <strong>based</strong> on multiple observed long-range amide-to-amide nuclear overhauser effects (NOEs), that it<br />

was indeed a single major conformational species. Although bradykinin contains three proline residues,<br />

the absence of any strong CαH i-CαH j+, cross peaks in the nuclear overhauser enhancement<br />

spectroscopy (NOESY) spectrum was taken as proof that all peptide bonds were trans. In total, 35<br />

interproton distances were extracted from the NOESY spectrum and, whenever possible, stereospecific<br />

assignments for pro-R and pro-S hydrogens were made explicitly. A temperature-dependent study of the<br />

chemical shifts of the amide protons resulted in a near-linear dependence suggesting no major<br />

conformational changes were coinciding with the temperature change and there<strong>by</strong> allowing a<br />

comparison of slopes (Δδ/Δt). The lowest values obtained for these slopes corresponded to Phe 8 and<br />

Arg 9 suggesting solvent sequestering for these amides.<br />

Given the high Chou and Fasman probability of β-turns in the sequences Pro 2-Pro 3-Gly 4-Phe 5 and Ser 6-<br />

Pro 7-Phe 8-Arg 9 (3.79 × 10 -4 and 1.99 × 10 -4, respectively), the computational strategy employed was to<br />

begin from two initial structures: (a) an extended β strand, and (b) a structure containing these two<br />

predicted β turns. Utilizing custom routines written using the program CHARMm, version 21 [21], the<br />

interproton distances were incorporated into the potential-energy expression in the form of an additional<br />

potential-energy term. During the 3-ps heating step of the molecular dynamics, the temperature was<br />

raised from 0K to 300K in steps of 20K every 0.2 ps. Since the target distances<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_122.html [4/5/2004 4:56:31 PM]


Document<br />

Page 123<br />

were poorly satisfied in the starting structures, the potential-energy term corresponding to the imposed<br />

NOE data (E NOE) was applied gradually <strong>by</strong> increasing the scale factor in a nonlinear fashion such that it<br />

was 0.0 after 0.2 ps, 0.1 after 1.2 ps, and 1.0 after the full 3.0 ps. Following 15 ps of equilibration, 7 ps<br />

of incremental production dynamics was completed. During this stage the NOE scale factor was raised<br />

from 1.0 to 4.5. By slowly raising the force constants for the NOE restraints as the target distances<br />

became better satisfied, no dramatic increase in temperature was observed. Finally the NOE scale factor<br />

was set to 5.0, 10 ps of production dynamics were completed, and an average structure was extracted<br />

from the last 5 ps of the coordinate trajectory.<br />

Analysis of the two average structures obtained from the two unique starting points demonstrated<br />

convergence to a similar conformational species. In each, the sum of the NOE restraint energy was less<br />

than 4.7 kcal/mol and the RMS deviation from the target distances was below 0.25 Å. Similar results<br />

were obtained for each simulation when they were repeated without the electrostatic term being included<br />

in the total potential-energy function. This important data lends credence to the hypothesis that the final<br />

structures are derived from the NOE restraints and not <strong>by</strong> poorly represented electrostatic interactions.<br />

The average dynamic structures are characterized as having all trans peptide bonds and hydrophobic side<br />

chain groups oriented outward into solution, perhaps ready to interact with the receptor. There is a<br />

possible 1–3 hydrogen-bonded γ turn bridging Phe 5, although it is not explicitly defined <strong>by</strong> the NOE<br />

data set. If present, then the preferred overall geometry would be U shaped and, if absent, an S shaped<br />

geometry is possible <strong>based</strong> on coincident conformational analyses. According to the dihedral angle<br />

values for Pro 7 and Phe 8, a type-II β turn extending from Ser 6 to Arg 9 also exists. A variety of reports<br />

have subsequently appeared that are in agreement with the conformation we described in this work.<br />

A similar C-terminal turn structure was observed in an analogous NMR study of a first-generation kinin<br />

antagonist, NPC 567 (DArg 0-Arg 1-Pro 2-Hyp 3-Gly 4-Phe 5-Ser 6-DPhe 7-Phe 8-Arg 9), although the type of<br />

turn was not the same. Our initial speculation was that this slight structural difference might partially<br />

account for the functional differences of bradykinin and NPC 567. These solution conformations, one of<br />

an agonist and the other of an antagonist, were subsequently used to focus the design and synthesis of<br />

conformationally constrained peptide analogues of NPC 567.<br />

B. Conformationally Constrained Bradykinin Antagonist Peptides<br />

The ligand-<strong>based</strong> approach of conformationally constrained peptides has been widely used. The process<br />

involves the incorporation of conformational constraints into known peptides, either agonist or<br />

antagonist, which enforce a<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_123.html [4/5/2004 4:57:26 PM]


Document<br />

Page 124<br />

predictable geometry. A series of peptides containing these types of constraints can be useful for<br />

extrapolating the steric and electronic environment of a given binding site. This structural information<br />

can be derived regardless of whether or not the constrained peptide binds to the target receptor. Since<br />

peptides can be prepared rapidly, it is typical to establish a structure-activity relationship using them and<br />

then at some later time transpose that information onto a nonpeptide lead molecule in an attempt to<br />

improve its potency.<br />

As part of an expansion upon the hypothesis that a C-terminal β turn was a structural prerequisite to highaffinity<br />

antagonist binding, a novel series of constrained decapeptides was prepared [22–24]. These<br />

peptides are of the sequence DArg 0-Arg 1-Pro 2-Hyp 3-Gly 4-Phe 5-Ser 6-DHype 7-Y 8-Arg 9, where Y is either<br />

tetrahydroisoquinoline-3-carboxylic acid (Tic), or octahydroindole-2-carboxylic acid (Oic). DHype<br />

denotes an organic ether of D-4-hydroxyproline in either the cis or trans geometric form. The C-terminal<br />

portion of a representative member of this class of peptides was shown—first <strong>by</strong> empirical calculation<br />

[22], then <strong>by</strong> NMR at 600 MHz—to adopt a β turn nearly unambiguously (figure 2) [25,26]. Moreover,<br />

it was shown <strong>by</strong> calculation that the turn was adopted regardless of the nature of the ether group (alkyl,<br />

aryl, etc.) or its geometry (cis or trans). Hence, a diverse series of these peptides was initially used as a<br />

tool to probe the steric and electrostatic topology of an antagonist<br />

Figure 2<br />

Lowest 5 kcal mol -1 of the calculated overall potential energy surface for a model<br />

peptide of Ser-DHype(trans propyl)-Oic-Arg. The contour interval is 0.5 Kcalmol -1<br />

and the highest (outermost) and lowest contour energy values are labeled.<br />

Superimposed on the contour plots are values for ψ i+1 and ψ i+2 from each of the<br />

thirty structures generated from the NMR data corresponding to the tetrapeptide<br />

Ser-DHype(trans propyl)-Oic-Arg.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_124.html (1 of 2) [4/5/2004 4:57:43 PM]


Document<br />

Page 125<br />

binding site on the bradykinin B2 receptor in the guinea pig ileum. The cis ethers, in all cases, bound to<br />

the receptor with significantly lower affinity than did the trans. A more complete listing of the peptides<br />

used in the study is shown in Table 1. These results support the hypothesis that the domain of the<br />

receptor that binds these antagonist ligands is partly made up of a hydrophobic cavity about one side of<br />

the C-terminal turn. However, adjacent to the other side of the<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_125.html (1 of 2) [4/5/2004 4:58:44 PM]


Document<br />

Figure 3<br />

Receptor binding curves for the binding of NPC 17410 and NPC 17643 to B2 receptors<br />

from the guinea pig ileum and cloned rat and human B2 receptors. Legends are noted on<br />

the figure.<br />

turn, there appears to be some type of steric interference (or lack of a pocket) that might otherwise<br />

accommodate the ethers of the cis configuration.<br />

Page 126<br />

More recently, bradykinin B2 receptors have been cloned from both rat and human sources [27,28]. In<br />

receptor-binding experiments using these new receptors, selected members of the DHype-containing<br />

decapeptides were used to probe these receptors [24], a representative sample of the data is shown in<br />

Figure 3. Specifically, NPC 17643 (a trans propyl ether of D-4-hydroxyproline at position 7) and NPC<br />

17410 (a cis propyl ether of D-4-hydroxyproline at position 7) were used. Although the trans ethercontaining<br />

decapeptide behaved similarly in binding assays directed toward the bradykinin B2 receptors<br />

in guinea pig, rat, and human, the cis ether-containing decapeptide, NPC 17410, displayed an interesting<br />

pharmacology. In particular, NPC 17410 bound with similar affinity to both the guinea pig and rat<br />

bradykinin B2 receptors, but had an appreciably higher affinity for the human B2 receptor. This result<br />

strongly suggests that there are slight structural differences in the antagonist binding sites of the rat and<br />

human B2 receptors. With clones available for the rat and human bradykinin B2 receptors, this<br />

prompted a systematic search using NPC 17410 binding to rat/human bradykinin B2 receptor chimeras<br />

and point mutations in an attempt to discover residues on the receptor that comprise this antagonist site.<br />

The details and results of the subsequent application of these novel receptor-probing ligands is fully<br />

described later in this chapter.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_126.html [4/5/2004 4:58:51 PM]


Document<br />

Page 127<br />

The des-Arg 9 forms of these peptides have also been shown to have high affinity for the recently cloned<br />

human B1 receptor [24]. An extension of the work described herein would be to use a more complete<br />

series of des-Arg 9 DHype-containing nonapeptides to probe the binding site of this new receptor where<br />

other interesting pharmacological differences are likely to exist since the B1 receptor is only 33%<br />

homologous to the human B2 [29].<br />

In summary, the DHype-containing decapeptides have been useful in many regards. First, they<br />

incorporate a novel β-turn mimetic that was alternatively functionalized and used to probe the unknown<br />

topology of the guinea pig, rat, and human bradykinin B2 receptors. In this role, one of these tools<br />

showed differential pharmacology between rat and human forms of the receptor. This tool was used in a<br />

synergistic fashion with subsequent molecular biological and computational procedures in the<br />

elucidation of an antagonist binding site. Second, these peptides, together with another potent<br />

decapeptide antagonist with similar conformational constraints [30,31], provide the first strong<br />

experimental evidence that high-affinity decapeptide bradykinin receptor antagonists adopt a C-terminal<br />

β turn in the receptor-bound conformation. Third, certain members of this series of decapeptides contain<br />

alkyl ethers of D-4-hydroxyproline at position seven. In this regard, they are the very first examples of<br />

decapeptide bradykinin receptor antagonists that do not contain a D-aromatic amino acid at the seventh<br />

position as had been previously deemed to be essential. Commercially, this renders the series patentably<br />

distinct from all other known bradykinin receptor antagonists. Finally, several members of the series<br />

(i.e., NPC 17731, NPC 17761, NPC 17974) are among the most potent antagonists for this receptor yet<br />

reported. Hence, there may be applications for these compounds as human therapeutics.<br />

Several “second generation” decapeptide antagonists have been reported, but the prototype from the<br />

class, which was first to be reported, is HOE 140 (DArg0-Arg1-Pro2-Hyp3-Gly4-Thi5-Ser6-DTic7-Oic8- Arg9) [30,31]. This decapeptide has also been shown to preferentially adopt a C-terminal β turn,<br />

consistent with the previous discussion [26,32,33]. The side chain of DTic at position seven is, however,<br />

flexible. While side-chain rotational movement is not allowed, the saturated six-membered ring easily<br />

undergoes an endo/exo ring-flipping motion. Hence, the β turn predominates about the backbone<br />

dihedral angles, but the side chain of DTic could be either endo or exo ring flipped in the receptor-bound<br />

conformation. In the absence of an appropriate x-ray crystallographic structure, there is no definitive<br />

means of establishing which possibility is correct. This type of ring-flipping conformational change<br />

serves to orient the bulky hydrophobic side chain of the DTic residue to either one side of the β turn, or<br />

the other. The data collected from the decapeptides containing either cis or trans ethers of D-4hydroxyproline<br />

at the analogous position in the<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_127.html [4/5/2004 4:58:53 PM]


Document<br />

Page 128<br />

sequence (discussed above) support a hypothesis that the DTic is exo ring flipped in the receptor-bound<br />

state.<br />

There are two factors that must be considered when applying structure-activity-relationship (SAR)<br />

information from a series of peptides toward the design of nonpeptide mimetics and putative library<br />

scaffolds. One is in regard to the backbone conformation that primarily serves as a structural scaffold<br />

upon which the various functionalities (side chains) are attached. The other factor is the side chains<br />

themselves whose spatial positions are primarily dictated <strong>by</strong> the backbone structure. Usually, the threedimensional<br />

arrangement of these differing chemical groups are responsible for affinity and triggering of<br />

the receptor. Knowledge of the relative importance of the individual side chains and amide bonds for<br />

receptor affinity is therefore a critical aspect of small molecule design from a peptidic structure-activity<br />

relationship.<br />

Conformationally constrained derivatives of HOE 140 have been prepared in continuing efforts to<br />

elucidate the ideal backbone conformation peptide antagonists must adopt for bradykinin B2 receptor<br />

interaction. One such series made use of Cα- or N-methyl substituted amino acids, incorporated at<br />

position(s) Gly4, Phe5, or both, in the peptide D-Arg0-Arg1-Pro2-Hyp3-Gly4-Phe5-Ser6-D-Tic7-Oic8-Arg9 (NPC 18545) [34]. An N-methyl substitution in the backbone of an L-amino acid is known to disfavor<br />

helical, or twisted, backbone conformations while favoring an extended backbone. The contrasting Cα methyl modification tends to favor a helical (twisted), rather than extended, conformation [35,36]. These<br />

conformational preferences apply only to the backbone φ, ψ angles (where φi and ψi correspond to<br />

backbone dihedral angles for residue i, defined <strong>by</strong> the four adjacent amino acid backbone atoms C i-1-N i-<br />

Cαi-Ci band Ni-Cαi-Ci-Ni+1, respectively) of the amino-acid residues bearing the modification. Receptor<br />

binding assays were performed in membrane preparations of the guinea pig ileum, a source of B2<br />

receptors, wherein these constrained peptides were evaluated for their abilities to compete with<br />

bradykinin binding.<br />

With the exception of the C α-methyl-Phe 5-containing peptide (NPC 18540), each conformational<br />

constraint caused a significant, at least 1000-fold, loss in binding affinity with respect to the<br />

unconstrained parent peptide, NPC 18545. There are several factors that could contribute to the poor<br />

receptor affinities measured for these peptides. In addition to the possible induction of an adverse<br />

conformation via the N-methyl substitution, this modification also eliminates an amide proton that might<br />

be an important hydrogen-bond donor during ligand-receptor interaction. Furthermore, the N-methyl<br />

substitution enhances the likelihood of trans-cis amide bond isomerization, which could also disrupt an<br />

optimal ligand-receptor interaction <strong>by</strong> altering the spatial display of the local side chains. The C α-methyl-<br />

Phe 5 substitution of NPC 18540 is well<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_128.html [4/5/2004 4:58:58 PM]


Document<br />

Page 129<br />

tolerated <strong>by</strong> the receptor as evidenced <strong>by</strong> only a seven-fold loss in receptor affinity (K i=0.54 nM) with<br />

respect to the parent of the series, NPC 18545. This implies that the φ, ψ backbone dihedral angles about<br />

Phe 5 are on the order of -60°, -60° in the biologically active conformation. This combination of dihedral<br />

angles represents a helical twist or “kink” in the midsection of the peptide.<br />

Since the original submission of the manuscript describing these constrained linear and cyclic peptides,<br />

bradykinin B2 receptors have been cloned from other species including human [27,28]. Toward the goal<br />

of designing a nonpeptide antagonist as a human therapeutic agent, it would be interesting to evaluate<br />

these analogues against these newly reported receptor homologues. This would likely be fruitful and<br />

valuable given that there is evidence, including that presented in this chapter, showing that guinea pig<br />

and rat B2 receptors differ structurally from the human B2 receptor at the antagonist binding site. Hence,<br />

a structure-activity relationship established against the guinea pig or rat receptors could be misleading in<br />

the context of potential human therapeutics.<br />

A systematic study of the relative importances of amides and side chains in a prototypical second<br />

generation antagonist, NPC 18545 (DArg0-Arg1-Pro2-Hyp3-Gly4-Phe5-Ser6-DTic7-Oic8-Arg9) has<br />

recently been described [37,38]. The D-Arg0 and Ser6-DTic7-Oic8-Arg9 segments were left intact in all<br />

peptides on the assumption that N-terminal positive charge(s) and a hydrophobic C-terminal β turn are<br />

minimally required for binding. In a systematic fashion, the amino acids in the core of the peptide (Arg1- Pro2-Hyp3-Gly4-Phe5) were substituted with glycine, an amino acid bearing no chirality or side chain.<br />

Binding assays, either in membranes from the guinea pig ileum or in membranes from a stable cell line<br />

expressing the human B2 receptor, were performed on each peptide and the results compared with the<br />

parent, NPC 18545, which has a Ki against [ 3H]-bradykinin of 0.08 nM. The elimination of all chirality<br />

and sidechain moieties in the segment Arg1-Pro2-Hyp3-Gly4-Phe5 via replacement <strong>by</strong> Gly1-Gly2-Gly3- Gly4-Gly5 (NPC 18152), led to a peptide that no longer binds the receptor. This demonstrated that one or<br />

more of the side chains in this segment are critical during ligand-receptor interaction. Incorporation of<br />

either the Arg1 or Phe5 side chains led to improved potency (285 nM and 483 nM, respectively).<br />

Constructing a peptide with side chains of both Arg1 and Phe5 in place (NPC 18149) yielded a peptide<br />

with good affinity, Ki of 13.7 nM. Overall, this study shows that to maintain potency in the low<br />

picomolar range, peptides in this series require Arg1, Phe5, and either Pro2 or Hyp3 (but not both). The Nterminal<br />

charged moieties and the hydrophobic C-terminal β turn are also required. Potencies in the low<br />

nanomolar range are attainable without including the side chains or chirality associated with Pro2 and<br />

Pro3. These data have<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_129.html [4/5/2004 4:59:04 PM]


Document<br />

subsequently been successfully applied toward the design and synthesis of several nopeptide scaffolds<br />

and mimetics. Ultimately these mimetics were assembled in a combinatorial fashion as discussed in<br />

Section IV.<br />

Page 130<br />

In a related study, wherein NPC 18149 (DArg 0-Arg 1-Gly 2-Gly 3-Gly 4-Phe 5-Ser 6-DTic 7-Oic 8-Arg 9; K i =<br />

13.7 nM; Guinea pig ileum) was taken as the lead peptide, the relative contributions to binding affinity<br />

from each amide bond in the segment Arg 1-Gly 2-Gly 3-Gly 4-Phe 5 were examined. Aminovaleric acid was<br />

used in a systematic fashion as a surrogate for any pair of adjacent Gly-Gly residues in the peptide.<br />

Aminovaleric acid is atomically identical to Gly-Gly with the exception that the amide bond linking the<br />

two glycines is replaced <strong>by</strong> two methylenes. The synthesis of Gly 4-Phe 5 required a special Gly-Phe<br />

mimic that has since been reported [39]. Since this substitution introduces flexibility into the peptide, it<br />

is a means of probing the structural role a given amide bond plays during receptor interaction. Potential<br />

hydrogen-bond donor and acceptor groups in the amide bond are removed via this substitution, which<br />

yields additional insights into potential electrostatic interactions that may also be important during<br />

ligand-receptor interactions.<br />

The conclusions drawn from the data are that in terms of structural or electrostatic interactions with this<br />

antagonist site on the receptor, the amide bond linking residues two and three may not be as critical as<br />

those linking residues three to four and four to five.<br />

Each of these investigations was aimed toward an understanding of either the backbone conformation of<br />

this prototypical decapeptide or the relative importance of the functional groups in the side chains that<br />

make significant contributions to receptor affinity. From the former, nonpeptide frameworks and<br />

scaffolds can be imagined. From the latter, insights into which functionality is required for high-affinity<br />

binding is derived. The remaining challenge is to reassemble these fragments onto synthetically feasible<br />

nonpeptide frameworks as potential new lead compounds. Our approach toward addressing this<br />

challenging problem is described later in this chapter.<br />

III. Receptor <strong>Structure</strong>-Based Investigations<br />

A. Elucidation of an Agonist Binding Site on the B2 Receptor<br />

In addition to the deductions one might make about a receptor binding site on the basis of receptor<br />

binding data from conformationally constrained ligands as previously described, models of bradykinin<br />

and bradykinin antagonists bound to their respective sites on the receptor as complimentary aspects of<br />

the overall strategy are also valuable. Unfortunately, due to the nature of the bradykinin<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_130.html [4/5/2004 4:59:07 PM]


Document<br />

receptor, it has not yet been obtained in crystalline form, nor is it likely to be in the near future.<br />

Page 131<br />

The bradykinin receptor is a member of a family of receptors for which an intracellular interaction with<br />

a G-protein is a critical part of the signal transduction pathway following agonist binding. Structurally,<br />

these G-protein-coupled receptors extend from beyond the extracellular boundary of the cell membrane<br />

into the cytoplasm. The tertiary structure is such that the protein crosses the bilayer of the cell membrane<br />

seven times, thus forming three intracellular loops, three extracellular loops, and giving rise to<br />

cytoplasmic C-terminal and extra-cellular N-terminal strands. It is generally presumed that the<br />

transmembrane domains of these receptors exist as a bundle of helical strands. This assumption is<br />

derived primarily from the known structure of the trans-membrane portions of a structurally related<br />

protein, bacteriorhodopsin [40].<br />

G-Protein-coupled receptors do not lend themselves to analysis <strong>by</strong> either NMR or x-ray crystallography<br />

due to their structural dependence on an intact cell membrane. In our laboratories we pursued this<br />

valuable structural information <strong>by</strong> utilizing a combination of structural homology modeling, molecular<br />

dynamics, systematic conformational searching methods, and mutagenesis experiments. The<br />

combination of these techniques led to a proposed model of bradykinin bound to the agonist site on its<br />

receptor [41].<br />

A hydrophobicity (Kyte-Doolittle) calculation [42] on the amino acid sequence of the rat bradykinin<br />

receptor yielded seven segments, each of which were 21 to 25 contiguous residues with predominantly<br />

hydrophobic side chains. These were presumed to be the seven transmembrane portions of the receptor.<br />

Cartesian coordinates of the backbone atoms within each of these seven segments were built <strong>by</strong><br />

structural homology from the cryomicroscopic structure of the analogous segments of<br />

bacteriorhodopsin. Subsequently, side chains were added to these seven segments as appropriate for the<br />

rat bradykinin receptor, and the resulting geometry was optimized via constrained energy minimization<br />

to alleviate bad contacts. Extracellular and intracellular loops were extracted from the Protein Data Bank<br />

library, following a geometric search <strong>based</strong> upon a vector defined <strong>by</strong> terminal alpha carbons in adjacent<br />

helices. The model was subsequently subjected to a series of constrained and unconstrained energy<br />

minimizations as well as molecular dynamics simulations. The resulting structure of the receptor was<br />

used in a novel two-step docking procedure.<br />

Following our hypothesis that bradykinin adopts a C-terminal β turn upon complexation with the<br />

receptor, the φ, ψ backbone dihedral angles in the tetrapeptide corresponding to the C-terminus of<br />

bradykinin (Ser-Pro-Phe-Agr) were constrained in a harmonic fashion (force constant = 15 Kcal Å -1 mol -<br />

1) to values that define a type II' β-turn [43]. This tetrapeptide probe was then systematically translated<br />

about the interior of a theoretical box inscribing the rat<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_131.html [4/5/2004 4:59:09 PM]


Document<br />

Page 132<br />

receptor model. The translations were such that the tetrapeptide probe molecule was incrementally<br />

repositioned within the receptor <strong>by</strong> following a 3 Å × 3 Å × 3 Å grid pattern. At each new position, both<br />

the probe and receptor were reset to their initial conformations, then the geometry of the complex was<br />

optimized using 200 steps of steepest descent followed <strong>by</strong> 500 steps of Adopted-Basis Newton-Raphson<br />

energy minimization. Subsequently, the sum of the steric and electrostatic contributions to the overall<br />

potential energy (interaction energy)—as measured only between the tetrapeptide probe molecule and<br />

the atoms of the receptor—were calculated. Slices through the receptor illustrating the energy of<br />

interaction as grayscale contour lines (darker gray = lower interaction energy) for that portion of<br />

receptor that was sampled <strong>by</strong> the tetrapeptide probe molecule is shown as an edge-on, frontal view in<br />

Figure 4. In this Figure it is qualitatively clear where the transmembrane domains are located (white), as<br />

well as where the most favorable sites of probe interaction are located (black).<br />

Figure 4<br />

Complete group of contour plots showing energy of interaction between probe<br />

and receptor. Each contour plot corresponds to a different horizontal slice as part<br />

of the first stage in the conformational search. Darker gray indicates most favorable<br />

interaction and the light shades represent least favorable interactions.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_132.html [4/5/2004 4:59:26 PM]


Document<br />

Page 133<br />

This initial stage of the docking process was used to reduce the computational difficulties that would be<br />

inherent in “tumbling” a complete bradykinin molecule (which has great flexibility) about the receptor<br />

in a similar fashion. However, following this initial stage, insight into those regions of the receptor<br />

capable of accommodating the C-terminal portion of the bradykinin molecule was obtained. On the basis<br />

of energetics, and as qualitatively shown in Figure 4, those particular regions are clustered in the central<br />

part of the receptor near to the extracellular domain. Using this information as a steering device to limit<br />

the size of the problem, an exhaustive conformational search was performed using the entire nineresidue<br />

sequence of bradykinin as a probe molecule, again enforcing a C-terminal β turn via dihedral<br />

angle constraints. Specifically, 24 unique geometric orientations (eight on each of three axes) of the<br />

bradykinin molecule were sampled at each of 100 grid points identified during the initial stage as likely<br />

zones to bind the tetrapeptide probe molecule. Bradykinin-receptor complexes within the lowest 150<br />

kcal mol -1 interaction energy with respect to the lowest found (17 complexes out of 2400) were grouped<br />

into sets of related conformational families, of which there were five. Computationally, each of the five<br />

complexes were presumed to be equally likely. All of these simulations were accomplished using<br />

custom routines written using the program CHARMm [21].<br />

To guide the selection of which of the five bradykinin-receptor complexes to consider a “lead” model,<br />

supporting experimental evidence was sought from site-directed mutagenesis experiments. This support<br />

was taken primarily from work describing bradykinin binding assays performed of mutant rat B2<br />

receptors [44,45]. The underlying strategy of the mutation studies was <strong>based</strong> on the hypothesis that,<br />

since bradykinin has positive charges at either end of its sequence (Arg 1 and Arg 9), separated <strong>by</strong> a group<br />

of rather hydrophobic amino acids (Pro 2-Pro 3-Gly 4-Phe 5-Ser 6-Pro 7-Phe 8), it was likely that some acidic<br />

residues in the receptor participated during ligand binding. Several mutant receptors were made such<br />

that each contained either a point mutation or a small cluster of point mutations, wherein native residues,<br />

having negatively charged side chains (Asp, Glu), were replaced <strong>by</strong> alanine(s). Table 2 lists the initial<br />

cluster mutations (rat) that were prepared as well as the follow-up single point mutations (rat).<br />

Figure 5 shows a stereoview of the selected ligand-receptor complex chosen on the basis of best<br />

agreement with the results of these mutagenesis studies. None of the other four putative complexes were<br />

in agreement with this experimental data and were not considered further. Of particular significance in<br />

this work was that the trans-membrane residue Glu 49, when mutated to alanine, showed no adverse<br />

effect on bradykinin receptor affinity with respect to rat wild type. A similar result was reported for the<br />

Glu 196 rarrow.gif Ala 196 mutation. These residues are remotely situated with respect to the proposed site<br />

of bradykinin<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_133.html [4/5/2004 4:59:30 PM]


Document<br />

binding and are colored light gray in Figure 5. In contrast, the [Asp175, Glu178,179] rarrow.gif<br />

Ala175,178,179 cluster mutation showed a 12-fold loss in bradykinin binding affinity, and the [Glu282, Asp 286] rarrow.gif Ala 282,286 cluster mutation lost 17-fold with respect to the wild type receptor. The<br />

Page 134<br />

Asp 268 rarrow.gif Ala 268 and Asp 286 rarrow.gif Ala 286 point mutations caused 19-and 28-fold<br />

respective losses in affinity for bradykinin. Close inspection of the bradykinin Arg 1 side chain location<br />

and surrounding receptor interactions led to the suspicion that Asp 286 and Asp 268<br />

Figure 5<br />

Proposed model of bradykinin bound to the rat B2 receptor at the agonist binding site.<br />

Only the upper portion of the receptor is shown as gray helical ribbons. Bradykinin<br />

backbone and side chain atoms are shown as thick white licorice. Positions of point<br />

mutations having no significant adverse effects on bradykinin binding are shown as<br />

light gray spheres. Positions of mutations affecting bradykinin binding are shown as<br />

dark gray spheres.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_134.html (1 of 2) [4/5/2004 4:59:38 PM]


Document<br />

Page 135<br />

might be jointly interacting either with the guanidino group in the side chain of Arg1 or the N-terminal<br />

amino group in bradykinin. Therefore a receptor containing a double mutation (Asp268,286 rarrow.gif<br />

Ala268,286) would be expected to show a much more dramatic loss in affinity for bradykinin than would<br />

receptors containing the individual point mutations. The appropriate double mutation experiment<br />

confirmed this <strong>by</strong> causing a 500-fold loss in affinity for bradykinin, as predicted (Table 2). The<br />

mutagenized residues of this double mutant B2 receptor are colored dark gray in Figure 5. This type of<br />

an ionic interaction is also precedented <strong>by</strong> the body of literature that exists supporting the requirement of<br />

an N-terminal arginine residue and a free N-terminal amino group in both bradykinin peptide agonists<br />

and antagonists for high affinity binding. All of these mutant receptors were demonstrated to be<br />

functional receptors on the basis of bradykinin-induced membrane depolarization in a Xenopus oocyte<br />

expression system [44,45].<br />

The selected agonist site model is characterized <strong>by</strong> an overall twisted S-shape ligand, similar to the<br />

conformation of bradykinin determined previously in a hydrophobic environment <strong>by</strong> NMR [19,46].<br />

Overall, the model suggested that the N-terminal amino and guanidine groups of Arg 1 interact directly<br />

with negatively charged amino acids in extracellular loop three, and the C-terminal end is in a β-turn<br />

conformation buried just below the extracellular boundary of the trans-membrane domain of the<br />

receptor. Noteworthy is the presence of a hydrophobic cavity in our receptor model located adjacent to<br />

Pro 7 of the bradykinin ligand. This cavity is made up, in part, <strong>by</strong> the residues Phe 261, Leu 104, Val 108, and<br />

Ile 112. Given the historical significance of position seven in peptide bradykinin-like ligands, these<br />

residues represent interesting targets for further mutagenesis experiments. One such result, the mutation<br />

of Phe 261 to Ala 261, has already been described, and the results were supportive of this proposed model<br />

[47]. Antibodies to the extracellular loops two and three have also been shown to compete with<br />

bradykinin binding, lending further experimental support for an extracellular domain on this agonist<br />

binding site.<br />

More recently, chemical crosslinking combined with site-directed mutagenesis was used to analyze the<br />

bradykinin binding site in the human B2 bradykinin receptor [48]. Previous studies using the bovine B2<br />

receptor showed that heterobifiunctional reagents reactive to amines and free sulfhydryls crosslink the<br />

bound bradykinin N-terminus to a sulfhydryl(s) in the receptor [49]. To identify this sulfhydryl(s), two<br />

conserved candidate residues in the human B2 receptor—Cys 20 in the N-terminal domain and Cys 277 in<br />

extracellular loop 3—were mutated to serine residues. Single and double mutants were expressed in Cos<br />

7 cells. All mutants bound [ 3H]bradykinin with typical B2 receptor specificity. The heterobifunctional<br />

reagent m-maleimidobenzoyl-N-hydroxysuccinimide ester crosslinked bradykinin to wild-type and<br />

mutants with maximum efficiencies of 35% (wild type), 40% (Ser 20), 20% (Ser 277), and 0%<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_135.html [4/5/2004 4:59:40 PM]


Document<br />

Page 136<br />

(Ser 20, Ser 277). This clearly demonstrated that Cys 20 and Cys 277 are the only sulfhydryls available for<br />

crosslinking receptor-bound bradykinin. These results provided direct biochemical evidence that the Nterminus<br />

of bradykinin, when bound to the B2 receptor, is adjacent to extracellular loop 3 and the Nterminal<br />

domain in the receptor.<br />

Further consideration of the model led to the hypothesis that agonist peptides may minimally require an<br />

intact C-terminal β-turn structure with appropriate side chains in place and N-terminal amino and<br />

guanidine groups for primary electrostatic interaction(s) with Asp286 and Asp268 in extracellular loop 3.<br />

As a test of this hypothesis, the prototypical second generation antagonist, NPC 18545, (DArg0-Arg1- Pro2-Hyp3-Gly4-Phe5-Ser6-DTic7-Oic8-Arg9) was modified such that residues 2–5 were replaced <strong>by</strong> a<br />

simple twelve carbon chain spacer (12-aminotridecanoic acid). The resulting compound, NPC 18325,<br />

contains only the appropriately charged moieties at the N-terminus, separated <strong>by</strong> a simple organic spacer<br />

moiety from a known β turn forming tetrapeptide [25,26]. This pseudopeptide was tested in the human<br />

bradykinin B2 receptor binding assay and found to have a Ki of 44 nM against [ 3H]bradykinin binding<br />

[50]. Functionally, this pseudopeptide was an agonist as measured <strong>by</strong> its ability to stimulate IP<br />

production in a stable CHO cell line expressing the human B2 receptor and in WI-38 cells. Since it was<br />

designed on the basis of an agonist site on the receptor, this result was not completely surprising despite<br />

the incorporation of the DTic-Oic pair at the C-terminus that previously had been shown to be critical in<br />

high affinity antagonists.<br />

Subsequently the length of the linear carbon chain was varied to further explore the hypothesis that the<br />

agonist site on the receptor had two domains, a hydrophilic site in the extracellular loop area and a<br />

hydrophobic domain in the transmembrane area [50]. Presumably, the two terminal portions of NPC<br />

18325 can only simultaneously interact with each putative domain of the receptor binding site when the<br />

carbon chain is 12–13 methylenes long. But if the carbon chain length is shortened too far, this ligand<br />

might be unable to simultaneously interact with both domains, resulting in an affinity loss. This series of<br />

pseudopeptides and their respective human bradykinin B2 receptor affinities are presented in Table 3.<br />

The data are consistent with the hypotyhesis since the receptor affinity decreases as the carbon chain<br />

length is shortened. An alternative explanation of the data is that a certain hydrophobicity profile is<br />

required of the compounds in this series for good receptor affinity. These results indicated that there<br />

may be additional hydrophobic or flexibility prerequisites to binding in this series of pseudopeptides.<br />

One noteworthy observation was that NPC 18325 showed divergent behavior when evaluated against<br />

different species homologues of the bradykinin B2 receptor. Notably, in the guinea pig ileal membrane<br />

preparation assay, the affinity for the receptor was approximately 10-fold less than what had been<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_136.html [4/5/2004 4:59:41 PM]


Document<br />

Page 137<br />

observed for the human B2 [37]. Furthermore, in contrast to the functional activity of NPC 18325 at the<br />

human B2 receptor, the compound is a functional antagonist as measured against bradykinin-induced<br />

contraction of the isolated guinea pig ileum (pA 2 = 5.5). These findings are in agreement with the<br />

concept that as a ligand is made smaller (i.e., fewer contact points possible with the receptor), the subtle<br />

structural differences in the binding sites on species variants of the same receptor become amplified.<br />

This observation further supports a cautionary posture toward developing nonpeptide antagonists for use<br />

in human diseases on the basis of results obtained in some animals including the guinea pig. Taking this<br />

new molecule as a lead structure, together with the receptor model and structure-activity relationship<br />

associated with related peptides including cyclic antagonists, the pursuit of several related<br />

pseudopeptides was undertaken.<br />

B. Elucidation of an Antagonist Site on the B2 Receptor<br />

There have been a variety of single alanine point mutations experimentally introduced into both rat and<br />

human bradykinin B2 receptors. Several of these have been shown to decrease the affinity of bradykinin<br />

to the receptor and have been implicated structurally near the agonist binding site. In contrast, at the<br />

time of this manuscript, there have been no mutations reported that adversely affect the ability of any<br />

peptide antagonists to bind to the receptor. Furthermore, antibodies raised against the certain<br />

extracellular domains of the kinin receptor compete with bradykinin for binding to the receptor but have<br />

no inhibitory<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_137.html [4/5/2004 4:59:52 PM]


Document<br />

Page 138<br />

action on the binding of antagonist peptides. In addition, it has been shown that bradykinin can be<br />

covalently crosslinked to the B2 receptor while antagonists cannot. These observations have fostered the<br />

belief that the agonist and antagonist binding sites of the receptor are not the same. At best, they may be<br />

partially overlapping, although there is no direct evidence for this. The ultimate identification of the<br />

amino acid residues that make up the antagonist site would be another important step toward the goal of<br />

structure-<strong>based</strong> design of novel nonpeptide antagonists.<br />

As described in a previous section of this chapter, characterizations of the bradykinin B2 receptors from<br />

rat and human using NPC 17410 (Figure 3) revealed different pharmacologies. Specifically, it showed a<br />

higher affinity for the human B2 receptor than it did for the rat B2 (human IC 50 = 0.95 nM, rat IC 50 =<br />

48.0). This ligand “tool” provided a means for evaluating a series of bradykinin rat/human B2 receptor<br />

chimeras [51–53]. Several different chimeras were prepared in a systematic fashion and the affinity of<br />

NPC 17410 was determined for each. The chimeras are depicted schematically in Figure 6 together with<br />

the IC 50 values determined for NPC 17410. Chimeras I through III sample the N-and C-terminal sections<br />

of the receptor for any contributions to an antagonist binding site. The remaining chimeras sample the<br />

core transmembrane domains of the receptor. Each chimera was shown to induce a membrane<br />

depolarization similar to wild type receptor in response to bradykinin when expressed in Xenopus<br />

oocytes. For each NPC 17410 assay, [ 3H]-NPC 17731 was used as the radioligand.<br />

From this systematic approach, specific groups of contiguous residues within the receptor were<br />

identified as possible contributors to an antagonist binding site. The NPC 17410 binding to chimeras III,<br />

IV, and VIII showed rat-like pharmacology (low NPC 17410 affinity). The NPC 17410 binding to<br />

chimeras I, II, VI, and VII showed human-like NPC 17410 pharmacology (high receptor affinity).<br />

Binding to chimeras V and VIII, however, was similar to rat-like NPC 17410 pharmacology, but the<br />

affinity of the compound was slightly shifted back toward human-like results. Comparisons of rat and<br />

human receptor sequences in the regions sampled <strong>by</strong> the chimeras reveals that only two clusters of<br />

residues differ between rat and human B2 receptors. Specifically, TM2 has the same sequence in rat and<br />

human receptors so it is unlikely that the differential pharmacology associated with NPC 17410 binding<br />

can be attributed to residues there. However, TM3 has a cluster of 3 residues that differ (rat rarrow.gif<br />

human: Thr110 rarrow.gif Ala108, Met111 rarrow.gif Ile109, Tyr113 rarrow.gif Ser111) and TM6 has a<br />

cluster of 5 residues that differ (rat rarrow.gif human: Phe259 rarrow.gif Leu257, Leu256 rarrow.gif<br />

Ile254, Val255 rarrow.gif Ile253, Gly252 rarrow.gif<br />

Leu 250, Ala 249 rarrow.gif Val 247) in rat and human receptors. These differences represent important<br />

targets for follow-up point (and cluster) mutation experiments. Our current thinking is that the largest<br />

effects on NPC 17410 pharmacology, if any, might be derived from the TM3 mutants since<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_138.html [4/5/2004 4:59:54 PM]


Document<br />

Figure 6<br />

Rat and human B2 receptor chimera constructs and affinity data<br />

for binding NPC 17410. Also shown is the affinity NPC 17410 to<br />

rat and human wild type receptors.<br />

Page 139<br />

between rat and human, these are quite diverse. However, it is also possible that the cluster of residues<br />

identified in TM6, while not radically dissimilar, may as a group create different hydrophobic<br />

environments between these species homologues. The most significant individual difference within the<br />

TM6 zone is the Phe 259 (rat) rarrow.gif Leu 257 (human) swap and might therefore be most significant<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_139.html (1 of 2) [4/5/2004 5:00:00 PM]


Document<br />

Figure 7<br />

Schematic of the primary amino sequence of the human B2 receptor. Shown<br />

in black are residues experimentally identified as contributing to an agonist binding site.<br />

The dark gray residues are suspect positions for contributing to an antagonist site. The<br />

residues colored light gray have been mutagenized only in the rat B2 receptor, but they<br />

are conserved in the human. The Thr 263 rarrow.gif Ala mutation interferes with agonist binding<br />

only, while Gln 260 partially interferes with agonist and first generation antagonist<br />

binding.<br />

Page 140<br />

within the context of these TM6 residues. Currently, we have prepared these mutant receptors, but at this<br />

time binding to NPC 17410 remains unfinished.<br />

A summary of the amino acids in the human B2 receptor implicated in comprising either agonist or<br />

antagonist sites are highlighted in Figure 7. Marked in dark black are the residues of extracellular loop 3,<br />

TM 6, and the extracellular N-terminal segment that have been shown to participate in agonist binding.<br />

Marked in dark gray in TM 6 and TM 3 are residues likely to partially comprise an antagonist binding<br />

site <strong>based</strong> primarily on the chimeric receptor studies described previously, although there is no explicit<br />

experimental evidence as yet. Shown in light gray are two residues that are conserved between rat and<br />

human B2 receptors. Mutagenesis experiments have been done on this pair in the rat B2<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_140.html (1 of 2) [4/5/2004 5:00:06 PM]


Document<br />

Page 141<br />

receptor with interesting results [54]. Mutations in Thr 263 only affect agonist binding, not antagonist.<br />

Mutations in Gln 260 affect binding of bradykinin and first generation antagonist peptides. As depicted in<br />

the figure, it is possible that the agonist and antagonist binding sites have domains on opposite sides of<br />

the helix that makes up TM 6, with Gln 260 being situated partly in both.<br />

IV. <strong>Design</strong> and Combinatorial Synthesis of Nonpeptidic Antagonists<br />

As was previously described, a significant body of information was generated that provides insights into<br />

the key structural features of bradykinin receptor binding sites and the residues that participate in ligand<br />

binding. In addition, from the ligand-<strong>based</strong> studies, knowledge about relevant structure-activity<br />

relationships was acquired. Our modular synthetic strategy was <strong>based</strong> primarily upon the recognition<br />

that high-affinity ligands appear to be comprised of three domains. These domains are (1) a positively<br />

charged N-terminal segment, (2) a midsection containing a bend or twist with some hydrophobic<br />

substituent attached and, (3) a C-terminal segment of appropriate hydrophobicity and structurally<br />

simulating a type II' β turn. Models of potent cyclic and linear peptide bradykinin receptor antagonists<br />

(described previously) were used in a comparative fashion to select nonpeptide ring systems from a<br />

database of chemical structures fine chemicals database. For each, some degree of chemical diversity<br />

was achieved <strong>by</strong> altering one of several parameters including, o, m, or p substitution of an aromatic ring<br />

or nature of alkyl substituent(s) as well as point(s) of synthetic attachment [55,56].<br />

Each nonpeptide fragment was designed within the framework of several criterion. First, a given<br />

scaffold must closely match the known SAR and be compatible with the putative ligand binding site<br />

structure. Second, each scaffold must be a relatively simple synthetic target, having readily available<br />

starting material, no chiral centers and having a total synthesis of not more than 4–5 steps. Finally, each<br />

template must have a “C-terminal” carboxylate and an “N-terminal” amino group with no interfering<br />

functionality such that it could be readily used in a solid phase synthetic strategy. Given that each<br />

nonpeptide we identified was a viable surrogate for either the second or third domain of high-affinity<br />

ligands (as described above) our goal was to rapidly explore the receptor affinities of all possible<br />

combinations of these nonpeptide templates at position X and Y of the sequence DArg-Arg-X-Y-Arg,<br />

hence a combinatorial synthetic approach was taken.<br />

In this study, there were four linear aminoalkanoic acids [50], four different cinnamic acids, three<br />

different carbolines, three different phenanthridinones, and five different spirocyclics. The variability in<br />

the phenanthridinione series<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_141.html [4/5/2004 5:00:07 PM]


Document<br />

Page 142<br />

was that the central ring could be opened or cleaved and the amino group could be meta or para<br />

substituted on the latter. In the carboline series, the cyclic amino group was either at the β or γ position<br />

of the cyclohexenyl ring and the methylene chain bearing the “C-terminal” carboxylate could be of<br />

variable length. The spirocyclic series was varied <strong>by</strong> alkyl, cycloalkyl, and aryl substitution on the fivemembered<br />

ring amine nitrogen. The cinnamic acids had two carbon chains that could be of varying<br />

length, one of which had the further possibility of containing a double bond(s).<br />

Rather than perform individual syntheses of all possible combinations of these nonpeptide units,<br />

members of each ring type or scaffold family were pooled in equimolar amounts prior to incorporation<br />

into the sequence DArg-Arg-X-Y-Arg. Since each individual member of a given pool was constructed<br />

on a similar carbocyclic scaffold, the chemical environment of the N-terminal amino group and Cterminal<br />

carboxylate groups were expected to follow similar kinetic and thermodynamic controls during<br />

the attachment of the nonpeptide residue to the growing peptide chain. The use of these smaller, directed<br />

libraries made it readily practical to obtain HPLC and mass spectral data for each and therefore confirm<br />

the composition of the library.<br />

Ultimately, 10 libraries of novel nonpeptidic structures were synthesized following typical solid-phase<br />

methodologies. Each library contained from nine to thirty-six different compounds in approximately<br />

equimolar amounts. Unpurified libraries were tested in a receptor binding assay utilizing membrane<br />

preparations from a stable CHO cell line expressing the human B2 receptor. Each library was tested at<br />

concentrations between 10 nM and 1μM. The ability of each library to inhibit[ 3H]-bradykinin binding<br />

was assessed and the results are presented in Figure 8a. Although this type of screening is highly<br />

qualitative, certain libraries appear in Figure 8a that show higher affinity to the receptor than other<br />

libraries. Library one (of the series DArg-Arg-PH-CN-Arg) was ultimately selected for further<br />

deconvolution. This library was further broken down (decoded) in order to determine which<br />

compound(s) were responsible for the apparent activity. It is important to note that breaking these<br />

libraries down to elucidate the structure of the hit(s) was feasible due to the inherently small size of each<br />

library.<br />

Library one contained 12 different structures (recall that there were originally three different<br />

phenanthridinones and four different cinnamic acids). The first deconvolution step of the approach is<br />

shown in Figure 9. Here, only the CN position was randomized, and the PH moieties were specific. This<br />

led to the preparation of three new libraries of 4 compounds each. Receptor binding was again<br />

performed as before and only one of these three new sublibraries showed affinity for the receptor at 10<br />

μM (Figure 8b). The final step in the process required to elucidate the active component(s) was to<br />

synthesize and purify each of the 4 members of this library as shown in Figure 9. Receptor binding on<br />

these<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_142.html [4/5/2004 5:00:09 PM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_143.html (1 of 2) [4/5/2004 5:00:18 PM]<br />

Page 143


Document<br />

Figure 8<br />

(a) Binding assay results for 10 original nonpeptidic libraries<br />

of the sequence DArg-Arg-X-Y-Arg, where X and Y are defined<br />

as PH = phenanthridinone, CB = carboline, SP = spirocycle,<br />

SC = straight chain, CN = cinnamic acid. Each library was tested at<br />

1 μnM and 10 nM. Results were compared to cold bradykinin<br />

binding, which was tested at two lower concentrations, 0.1 nM<br />

and 1 nM. Panels (b) and (c) correspond to the receptor binding<br />

results obtained using the two breakdown steps from original<br />

library number 1, as shown in Figure 9.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_143.html (2 of 2) [4/5/2004 5:00:18 PM]


Document<br />

Figure 9<br />

Composition of ten original nonpeptidic libraries of the sequence DArg-Arg-X-Y-Arg.<br />

X and Y were selected from the set of scaffolds shown in Table 1. Also shown<br />

are the subsequent breakdown libraries from original library number 1. Two-letter codes<br />

used in the figure correspond to the different nonpeptide moieties described in Table 1.<br />

Specifically, PH = phenanthridinone, CB = carboline, SP = spirocycle, SC = Straight<br />

chain, CN = cinnamic acid.<br />

Page 144<br />

four novel nonpeptidic structures showed that only one of the four had affinity to the receptor (Figure<br />

8c). This new compound, I, was subsequently shown to be an antagonist in a cellular assay measuring<br />

bradykinin-stimulated IP turnover [18]. Overall, there were 285 possible structures to survey due to the<br />

number of structure-<strong>based</strong> scaffolds that were prepared. This was rapidly accomplished via 19 synthetic<br />

couplings, 19 assays, and 4 purifications.<br />

Not surprisingly, compound I showed divergent potency when assayed in different species homologues<br />

of the bradykinin B2 receptor. In particular, in a model of bradykinin-induced hypotension in rats and<br />

rabbits, it showed no activity. Likewise, it did not block bradykinin-induced contraction of the isolated<br />

guinea pig ileum. Since compound I is considerably smaller than previously reported decapeptide<br />

antagonists, subtle structural differences (which are<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_144.html (1 of 2) [4/5/2004 5:00:24 PM]


Document<br />

known to exist in species homologs of the bradykinin B2 receptor) are likely amplified. A more<br />

comprehensive pharmacological analysis of compound I is currently underway.<br />

A. Lead Optimization<br />

Page 145<br />

We have previously reported that the C-terminal guanidinyl moiety of Arg [9] in prototypical peptide<br />

bradykinin antagonists is likely to behave more as an aromatic functional group rather than a hydrogenbond<br />

donor/acceptor. This speculation was <strong>based</strong> on proposed models of the agonist and antagonist<br />

binding sites of this receptor that have been elucidated using molecular biological and computational<br />

procedures. On this premise, the newly discovered lead compound, I, was altered such that the Cterminal<br />

arginine was replaced <strong>by</strong> 3',5'-dimethylpyrimidylornithine in an attempt to increase potency.<br />

This known mimetic of arginine contains an aromatic 3',5'-dimethylpyrimidyl ring in the side chain<br />

rather than the guanidino group on naturally occurring arginine.<br />

The results of the receptor binding assay performed using this compound, IA, are shown in Table 4<br />

where it is clear that affinity to the human B2 receptor is improved with respect to compound I. This<br />

data is supportive of the notion that the C-terminal residue(s) in this new series of bradykinin antagonist<br />

compounds interact with a hydrophobic environment, perhaps within the transmembrane domain of the<br />

receptor as previously suggested.<br />

The discovery of I and IA is significant in many regards. First, they are highly nonpeptidic lead<br />

compounds that could be further modified to improve potency and/or reduce molecular weight. Such<br />

improvements might lead to novel therapeutic agents for the treatment of inflammatory diseases. Thus<br />

far in the kinin antagonist literature there is significant evidence showing that, for compounds containing<br />

a C-terminal arginine residue, removal of that arginine generally yields compounds that are antagonists<br />

of the B1 subtype of the bradykinin receptor. Following a similar strategy with compound I could lead to<br />

the discovery of a novel series of nonpeptidic B1 receptor antagonists, although this remains to be<br />

demonstrated.<br />

V. Conclusions<br />

There has been a significant effort invested toward the discovery of novel bradykinin receptor<br />

antagonists during the past decade. In that time, several generations of peptide antagonists have been<br />

developed and a few are in human clinical trials. The pursuit of nonpeptide antagonists of the human<br />

bradykinin B2 receptor continues and incorporates a wide range of strategic approaches. The approach<br />

described herein is an early and very good example of a combinatorial synthesis of nonpeptide building<br />

blocks that mimic peptide structure,<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_145.html [4/5/2004 5:00:27 PM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_146.html [4/5/2004 5:00:51 PM]<br />

Page 146


Document<br />

Page 147<br />

ultimately tested in a nontagged, solution-phase form. Perhaps more significant is that the success<br />

described here demonstrates a possible synergy between structure-<strong>based</strong> design and combinatorial<br />

methodology. This approach has many merits, but the most significant is the application of structurally<br />

directed libraries toward target binding-site structures which, for one reason or another, may not be fully<br />

characterized. This method serves to aim the combinatorial syntheses in a logical direction, rather than<br />

attempt to prepare libraries of vast diversify (and numbers).<br />

Finally, since the libraries of compounds that were prepared contained few members, it was possible to<br />

analytically characterize each of the pools to assess integrity of their composition. Overall, there are<br />

many important advantages in the paradigm we have adopted that make the strategy generally viable in<br />

the context of structure-<strong>based</strong> lead compound discovery.<br />

References<br />

1. Greenbaum LM. Adv Exp Med Biol 1986; 198A:55.<br />

2. Okamoto H, Greenbaum LM. Biochem Biophys Res Commun 1983; 112:701.<br />

3. Muller-Esterl W. Thromb Haemost 1989; 61:2.<br />

4. Bhoola KD, Figueroa CD, Worthy K. Pharmacol Rev 1992; 44:1.<br />

5. Proud D, Perkins M, Pierce JV, Yates KN, Highet PF, Mangkornkanok/Mark M, Bahu R, Carone F,<br />

Pisano JJ. J Biol Chem 1981; 256:10634.<br />

6. Kitamura N, Takagaki Y, Furoto S. Nature 1983; 305:545.<br />

7. Ward PE. In: Burch RM, ed. Bradykinin Antagonists: Basic and Clinical Research. New York:<br />

Marcel Dekker, 1991:147–170.<br />

8. Perkins MN, Campbell EA, Davis A, Dray A. Br J Pharmacol 1992; 107:237P.<br />

9. Perkins MN, Campbell EA, Dray A. Pain 1993; 53:191.<br />

10. Perkins MN, Kelly D. Br J Pharmacol 1993; 110:1441.<br />

11. Kyle DJ, Burch RM. Curr Opin Invest <strong>Drug</strong>s 1993; 2:5.<br />

12. Kyle DJ, Burch RM. <strong>Drug</strong>s of the Future 1992; 17(4):305.<br />

13. Huang HM, Lin TA, Sun GY, Gibson GE. J Neurochem 1995; 64:761.<br />

14. Rubin LE, Levi R. Circ Res 1995; 76:430.<br />

15. Bao G, Gohlke P, Unger T. J Cardiovasc Pharmacol 1992; 20(Supplement 9):S96.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_147.html (1 of 2) [4/5/2004 5:00:59 PM]


Document<br />

16. Martorana PA, Kettenbach B, Breipol G, Linz W, Scholkens BA. Eur J Pharmacol 1990; 182:395.<br />

17. McDonald KM, Mock J, D' Ajoia M, Parrish T, Hauer K, Francis G, Stillman A, Cohn JN.<br />

Circulation 1995; 91(7):2043.<br />

19. Kyle DJ, Blake PR, Hicks RP. In: Burch RM, ed. Bradykinin Antagonists: Basic and Clinical<br />

Research. New York: Marcel Dekker, 1991:131–146.<br />

20. Chou PY, Fasman GD. Biophys J 1979; 26:367.<br />

18.<br />

Schwieler<br />

JH, Kahan<br />

T,<br />

Nussberger<br />

J,<br />

Hjemdahl<br />

P. Am J<br />

Physiol<br />

1993;<br />

264:E631.<br />

21. (a) Brooks BR, Bruccoleri RE, Olafson BD, States DJ, Swaminathon S, Karplus MJ. CHARMM: a<br />

program for macromolecular energy minimization, and dynamics calculations. J Comp Chem 1983;<br />

4:187–217.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_147.html (2 of 2) [4/5/2004 5:00:59 PM]


Document<br />

(b) Molecular Simulations, Inc., 16 New England Executive Park, Burlington, MA 01803-5297.<br />

22. Kyle DJ, Martin JA, Burch RM, Carter JP, Lu S, Meeker S, Prosser JC, Sullivan JP, Togo J,<br />

Noronha-Blob L, Sinsko JA, Walters RF, Whaley LW, Hiner RN. J Med Chem 1991; 34:2649.<br />

Page 148<br />

23. Kyle DJ, Burch RM. <strong>Drug</strong>s of the Future 1992; 17(4):305.<br />

24. Hiner RA, Chakravarty S, Lu S, et al. J Med Chem 1996; manuscript in preparation.<br />

25. Kyle DJ, Green LM, Blake PR, Smithwick D, Summers MF. Pep Res 1992; 5:206.<br />

26. Kyle DJ, Blake PR, Smithwick D, Green LM, Martin JA, Sinsko JA, Summers MF. J Med Chem<br />

1993; 36:1450–1460.<br />

27. McEachern AE, Shelton ER, Shakta S, Obernolte R, Bach C, Zuppan P, Fujisaki J, Aldrich RW,<br />

Jarnagin K. Proc Natl Acad Sci USA 1991; 88:7724.<br />

28. Hess JF, Borkowski JA, Young GS, Strader CD, Ransom RW. Biochem Biophys Res Commun<br />

1992; 184(1):260.<br />

29. Menke JG, Borowski JA, Bierilo KK, et al. J Biol Chem 1994; 269:21583–21586.<br />

30. Hock FJ, Wirth K, Albus U, Linz W, Gerhards HJ, Wiemer G, Henke S, Breipohl G, Knoig W,<br />

Knolle J, Schölkens BA. Br J Pharmacol 1991; 102:769.<br />

31. Wirth K, Hock FJ, Albus U, Linz W, Alpermann HG, Anagnostopoulos H, Henke S, Breipohl G,<br />

Knoig W, Knolle J, Schölkens BA. Br J Pharmacol 1991; 102:774.<br />

32. Sawutz DG, Salvino JM, Seoane PR, Douty BD, Houck WT, Bobko MA, Doleman MS, Dolle RE,<br />

Wolfe HR. Biochem 1994; 33:2373.<br />

34. Chakravarty S, Wilkens D, Kyle DJ. J Med Chem 1993; 36:2569.<br />

33. HOE SDS MICELLES<br />

35. Momany FA. In: Metzger, RM, ed. Topics in Current Physics. Vol. 26; New York: Springer Verlag,<br />

1981:41–79.<br />

36. Momany FA, Chuman H. Meth Enz 1986; 124:3–17.<br />

37. Kyle DJ. Brazil J Med Bio Res 1994; 27:1757.<br />

38. Chakravarty S, Mavunkel BJ, Lu S, Goehring R, Wu JP, Connolly M, Valentine H, Liu YW, Tam C,<br />

Andy R, Kyle, DJ. 23rd European Peptide Symposium, Braga: Sept 4–10, 1994.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_148.html (1 of 2) [4/5/2004 5:01:01 PM]


Document<br />

39. Mavunkel BJ, Lu Z, Kyle DJ. Tett Lett 1993; 34:2255.<br />

40. Henderson R, Baldwin JM, Ceska TA, Zemlin F, Beckmann E, Downing KH. J Mol Biol 1990;<br />

213:899.<br />

41. Kyle DJ, Chakravarty S, Sinsko JA, Stormann TM. J Med Chem 1994; 37:1347.<br />

42. Kyte J, Doolittle RF. J Mol Biol 1982; 157:105.<br />

43. Rose GD, Gierash LM, Smith JA. Adv Prot Chem 1985; 37:1.<br />

44. Novotny E, Bednar D, Connolly M, Connor J, Stormann T. In: Burch RM, ed. Molecular Biology<br />

and Pharmacology of Bradykinin Receptors. Austin, TX: R. G. Landes Company, 1993:19–30.<br />

45. Novotny EA, Bednar DL, Connolly MA, Connor JR, Stormann TM. BBRC 1994; 201:523.<br />

46. Lee SC, Russell AF, Laidig WD. Int. J Pept Prot Res 1990; 35(5):367.<br />

47. Freedman R, Jarnagin K. Cloning of a B2 Bradykinin Receptor: Recent Progress on Kinins. Basel:<br />

Birkhauser Verlag, 1992:487–496.<br />

48. Herzig MCS, Leeb-Lundberg F, Nash N, Connolly M, Kyle DJ. Kinin '95. International conference<br />

on kallikreins and kinins, Denver, CO, Sept, 1995.<br />

49. Herzig MCS, Leeb-Lundberg F. J Biol Chem 1995; in press.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_148.html (2 of 2) [4/5/2004 5:01:01 PM]<br />

50.<br />

Chakravarty<br />

S, Connolly<br />

MA, Kyle<br />

DJ. Peptide<br />

Research<br />

1995; 8:16.


Document<br />

51. Nash N, Connolly MA, Stormann TM, Kyle DJ. 14th Am Pep Sym, June 18, 1995.<br />

52. Nash N, Connolly MA, Stormann TM, Kyle DJ. Mol Pharm 1996; manuscript in preparation.<br />

53. Burch RM, Kyle DJ, Stormann TM. In: Molecular Biology and Pharmacology of Bradykinin<br />

Receptors. Austin, Texas: R. G. Landes Company, 1993:19–32.<br />

54. Nardone J, Hogan PG. PNAS 1994; 91:4417.<br />

55. Chakravarty S, Mavunkel B, Andy R, Kyle DJ. Network Science 1995; 1:1.<br />

56. Chakravarty S, Mavunkel BJ, Andy R, Kyle DJ. 14th American Peptide Symposium, Columbus,<br />

Ohio, June, 1995.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_149.html [4/5/2004 5:01:02 PM]<br />

Page 149


http://legacy.netlibrary.com/reader/message.asp?message=811&BookID=12640&FileName=Page_150.html<br />

The requested page could not be found.<br />

Return to previous page<br />

http://legacy.netlibrary.com/reader/message.asp?message=811&BookID=12640&FileName=Page_150.html [4/5/2004 5:01:05 PM]


Document<br />

5<br />

<strong>Design</strong> of Purine Nucleoside Phosphorylase Inhibitors<br />

Y. Sudhakara Babu, John A. Montgomery, and Charles E. Bugg<br />

BioCryst Pharmaceuticals, Inc., Birmingham, Alabama<br />

W. Michael Carson, Sthanam V. L. Narayana, and William J. Cook<br />

The University of Alabama at Birmingham, Birmingham, Alabama<br />

Steven E. Ealick<br />

Cornell University, Ithaca, New York<br />

Wayne C. Guida and Mark D. Erion *<br />

Ciba-Geigy Corporation, Summit, New Jersey<br />

John A. Secrist, III<br />

Southern Research Institute, Birmingham, Alabama<br />

I. Introduction<br />

A. Enzymology<br />

Page 151<br />

Purine nucleoside phosphorylase (PNP, E.C. 2.4.2.1) catalyzes the reversible phosphorylysis of<br />

ribonucleosides and 2'-deoxyribonucleosides of guanine, hypoxanthine, and related nucleoside analogs<br />

[1]. It normally acts in the phosphorolytic direction in intact cells, although the isolated enzyme<br />

catalyzes the nucleoside synthesis under equilibrium conditions. Figure 1 shows the chemical reaction.<br />

The enzyme has been isolated from both eukaryotic and prokaryotic organisms [2] and functions in the<br />

purine salvage pathway [1,3]. Purine nucleoside phosphorylase isolated from human erythrocytes is<br />

specific for the 6-oxypurines and many of their analogs [4] while PNPs from other organisms vary in<br />

their specificity [5]. The human enzyme is a trimer with identical subunits and a total molecular mass of<br />

about 97,000 daltons [6,7]. Each subunit contains 289 amino acid residues.<br />

* Current affiliation: Gensia, Inc., San Diego, California.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_151.html [4/5/2004 5:01:07 PM]


Document<br />

B. Pharmacology<br />

Figure 1<br />

The reaction catalyzed <strong>by</strong> PNP.<br />

Page 152<br />

Interest in PNP as a drug target arises from its ability to rapidly metabolize purine nucleosides and from<br />

its role in the T-cell branch of the immune system. Unfortunately, PNP can also cleave certain<br />

anticancer and antiviral agents that are synthetic mimics of natural purine nucleosides, thus interfering<br />

with therapy. One such substance is ddI (2'3'-dideoxyinosine), which the Food and <strong>Drug</strong> Administration<br />

approved as a treatment for AIDS in 1991. Another is the potential anticancer agent 2'-deoxy-6thioguanosine<br />

[8]. Our goal was to develop a compound that when administered with the nucleoside<br />

analogs would inhibit PNP while the anticancer and antiviral agents accomplished their therapeutic<br />

missions. The combination of purine nucleoside analogs and a PNP inhibitor might prove to be a more<br />

effective treatment.<br />

The PNP inhibitors alone have potential therapeutic value <strong>based</strong> on the importance of PNP to the<br />

immune system. Patients lacking PNP activity exhibit severe T-cell immunodeficiency while<br />

maintaining normal or exaggerated B-cell function [9]. We, like other researchers, quickly recognized<br />

that PNP inhibitors might selectively suppress the T-cell proliferation associated with an array of<br />

autoimmune disorders such as rheumatoid arthritis, psoriasis, systemic lupus erythematosus, multiple<br />

sclerosis, and insulin-dependent (juvenile-onset) diabetes [10]. This profile also suggests that PNP<br />

inhibitors might be useful in the treatment of T-cell proliferative diseases—such as T-cell leukemia or Tcell<br />

lymphoma—and in the prevention of organ transplant rejection.<br />

C. <strong>Drug</strong> <strong>Design</strong> Strategy<br />

Recent advances in biotechnology, macromolecular crystallography, computer graphics, and related<br />

fields have led to a new approach in drug discovery called structure-<strong>based</strong> drug design. <strong>Structure</strong>-<strong>based</strong><br />

drug design requires a detailed structural knowledge of the target (enzyme or receptor) and the<br />

interaction of small molecules with it.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_152.html [4/5/2004 5:01:10 PM]


Document<br />

Figure 2<br />

<strong>Structure</strong>-<strong>based</strong> drug design strategy.<br />

Page 153<br />

A tight fit is necessary for potency and specificity. A drug that binds to its target and inactivates it for a<br />

long time can be administered in lower doses than one that rapidly separates from its target. A substance<br />

designed to mesh perfectly with a particular binding site of one target is unlikely to interact well with<br />

any other molecule, minimizing unwanted interactions and side effects.<br />

Having chosen PNP as the target, we followed a systematic strategy for designing inhibitory<br />

compounds. Figure 2 outlines the overall strategy of this approach. To serve as a drug, an inhibitor has<br />

to readily cross cell membranes to the interior of cells, where PNP is located.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_153.html (1 of 2) [4/5/2004 5:01:14 PM]


Document<br />

We determined the structure of human PNP <strong>by</strong> x-ray crystallography and used these results in<br />

combination with computer-assisted molecular modeling to design inhibitor candidates. We examined<br />

how well the shape and chemical<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_153.html (2 of 2) [4/5/2004 5:01:14 PM]


Document<br />

Page 154<br />

structure of a candidate would complement the active site of PNP. We used computational chemistry to<br />

estimate the strength of the attractive and repulsive forces between a candidate and the enzyme.<br />

We synthesized only those candidates suggested <strong>by</strong> chemical intuition and computer simulation to have<br />

high affinity for the target. Then we measured the inhibition of PNP and compared the proposed with the<br />

actual fit. Because modeling programs and expert opinion are imperfect, certain compounds did not<br />

meet expectations. After exploring the reasons for the successes and the failures, we returned to<br />

interactive computer graphics to propose modifications that might increase the effectiveness of drug<br />

candidates.<br />

The resulting compounds were evaluated <strong>by</strong> determination of their IC 50 values (the inhibitor<br />

concentration causing 50% inhibition of PNP) and <strong>by</strong> x-ray diffraction analysis using difference Fourier<br />

maps. This iterative strategy—modeling, synthesis, and structural analysis—led us to a number of highly<br />

potent compounds that tested well in whole cells and in animals.<br />

D. Previously Known Inhibitors<br />

At the beginning of our studies several PNP inhibitors had been reported with Ki values in the 10-6 to 10- 7 range, including 8-aminoguanine [11], 9-benzyl-8-aminoguanine [12], and 5'-iodo-9-deazainosine [13].<br />

Acyclovir diphosphate had been shown to have a Ki near 10-8 if assayed at 1 mM phosphate rather than<br />

the more frequently used value of 50 mM phosphate [14]. During our studies, the synthesis of 8-amino-<br />

9(2-thienylmethyl)guanine was reported with a Ki of 6.7 × 10-8 M [15]. Figure 3 illustrates some of these<br />

structures.<br />

Despite the potential benefits of PNP inhibitors and the large number of PNP inhibitors that had been<br />

synthesized, no compound had reached clinical trials. None of these compounds were potent enough to<br />

be useful for therapy and also capable of crossing the cell membrane intact. Although potencies for the<br />

best compounds had affinities 10–100 fold higher than the natural substrate (K m = 20 μM), it is expected<br />

that T-cell immunotoxicity will only occur with very tight binding inhibitors (K i < 10 nM) due to the<br />

high level of in vivo PNP activity and competition with substrate.<br />

II. Crystallography<br />

At the present time, x-ray crystallography is the preferred technique for obtaining the required atomic<br />

resolution structural data. In the late 1970s when this project was first conceived, determining the<br />

structure of a protein was far from routine. The x-ray structural determination occupied a team of<br />

crystallographers led <strong>by</strong> Steven E. Ealick, then at the University of Alabama at Birmingham through<br />

most of the 1980s.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_154.html [4/5/2004 5:01:16 PM]


Document<br />

Figure 3<br />

Previously known inhibitors.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_155.html (1 of 2) [4/5/2004 5:01:20 PM]<br />

Page 155


Document<br />

Page 156<br />

The stumbling block did not lie with obtaining pure PNP or converting the protein into crystals. Robert<br />

E. Parks, Jr. and Johanna D. Stoeckler of Brown University had already isolated the enzyme from<br />

human cells. They supplied quantities of protein to William J. Cook, who succeeded in preparing the<br />

well-ordered crystals required for x-ray studies [16]. We established that PNP crystals function normally<br />

as a catalyst. Thus crystalline PNP is essentially identical to PNP in the body. If it were profoundly<br />

different, one would have no justification for basing drug design on the crystal structure.<br />

In the early years we had to depend on a relatively low-intensity x-ray source. High-resolution data was<br />

obtained through collaboration with John R. Helliwell and his group at the Daresbury Laboratory<br />

Synchrotron Radiation Source in England. Today greatly improved equipment and more synchrotron<br />

facilities are available for protein crystallography.<br />

The three-dimensional structure was determined <strong>by</strong> multiple isomorphous replacement techniques using<br />

synchrotron radiation [17]. The native and guanine-PNP complex structures have been refined to 2.8 Å<br />

resolution [18,19].<br />

A. <strong>Structure</strong> of the Enzyme<br />

Crystals of human PNP are grown from ammonium sulfate solution and stored in artificial mother liquor<br />

solution made of 60% ammonium sulfate in 0.05 M citrate buffer at pH 5.4. The space group is R32 with<br />

hexagonal cell parameters a=142.9(1) Å and c=165.2(1) Å. The PNP crystals contain about 76% solvent<br />

and diffract to around 2.8 Å resolution.<br />

The x-ray data established that PNP crystals contain a high percentage of water. This feature proved<br />

very useful; proposed drugs could easily be soaked into the active site without disrupting the crystal<br />

packing. Figure 4A shows the<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_156.html [4/5/2004 5:01:21 PM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_157.html (1 of 2) [4/5/2004 5:01:43 PM]<br />

Page 157


Document<br />

Figure 4<br />

<strong>Structure</strong> of PNP as stereo drawings. (A) Crystal packing. The low-resolution<br />

surfaces of six trimers are shown. The top level is related to the bottom level <strong>by</strong> a 2-fold<br />

axis along X. The distance between the 3-fold axes is 143 Å. A drug molecule is shown<br />

in the solvent channel near the entrance to an active site. (B) Ribbon drawing. The<br />

lowermost trimer of Figure 4A is shown. This is the native structure; the guanine and<br />

phosphate are shown to mark the active site. (C) The swinging gate. The trimer of Figure<br />

4B is rotated about 30° counterclockwise in the plane, followed <strong>by</strong> a roughly 90° rotation<br />

about X to view the entrance to the active site. A model of the transition state is shown<br />

as a line drawing. Conformational changes of the protein on binding guanine are shown.<br />

Arrows are drawn from the C α positions in the native structure to their positions in the<br />

complex. (D) Active site of PNP. The orientation is approximately that of Figure 4C, but<br />

enlarged and clipped to focus on the substrate. Key side-chain residues are labeled.<br />

Residue 159'F, in the center of figure toward the viewer, is the only residue from the<br />

adjacent subunit. The guanosine and phosphate are shown with thicker bonds. Oxygen<br />

and sulfur atoms are shown as white spheres, nitrogen and phosphorus as black spheres.<br />

(E) Purine binding pocket. The style is the same as Figure 4D, but the figure is rotated<br />

slightly and enlarged. The key group interacting with bound guanine are highlighted. (F)<br />

The best inhibitor. The style is the same as Figure 4D, but the figure has been enlarged<br />

and rotated to place the phosphate binding site far from the viewer in the upper left.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_157.html (2 of 2) [4/5/2004 5:01:43 PM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_158.html (1 of 2) [4/5/2004 5:02:02 PM]<br />

Page 158


Document<br />

Figure 4<br />

(Continued) Only one nitrogen atom from Arg 84 is visible, which—along<br />

with Ser 220 and His 86—interact with the acetyl group branching from the benzylic<br />

carbon. The chlorinated phenyl group is in the center of the figure, interacting with the<br />

aromatic groups in the sugar binding site. The guanine group interactions are the same as<br />

seen in Figure 4E. Figure prepared with ribbons (http://www.cmc.uab.edu/ribbons).<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_158.html (2 of 2) [4/5/2004 5:02:02 PM]


Document<br />

large solvent channels and the position of the active site. The x-ray analysis confirmed the trimeric<br />

nature of the enzyme, as the subunits are related <strong>by</strong> the crystallographic three-fold axis.<br />

Page 159<br />

A ribbon diagram of the trimer is shown in Figure 4B. Each monomer contains an eight-stranded β sheet<br />

and a five-stranded β sheet that join to form a distorted β barrel. Seven α helices surround this β sheet<br />

structure.<br />

The active site is an irregular indentation on the surface of the enzyme, located from the position of a<br />

tightly bound sulfate ion and various substrate analogs. These investigations revealed the identity of the<br />

exact amino acids constituting the active site region; such detail was a prerequisite to drug design.<br />

Information of greater import emerged from analyses of the complexes formed when synthetic<br />

nucleosides, including previously discovered inhibitors, were diffused into the active site.<br />

B. The Active Site<br />

The structural determinations also yielded a surprise. The shape of the enzyme changes when a purine is<br />

bound. The famous lock-and-key analogy [20] has a fallacy; the shape of the lock is not static, but<br />

flexible. Awareness of these conformational changes critically aided our modeling efforts, allowing<br />

prediction of which parts of PNP could change shape to interact with a proposed inhibitor.<br />

A “swinging gate” consisting of residues 241–260 controls access to the active site (Figure 4C). These<br />

residues in the native structure had poorly defined electron density with high thermal motion. The gate<br />

opens in the native enzyme to accommodate the substrate or inhibitor. The maximum movement caused<br />

<strong>by</strong> substrate or inhibitor binding occurs at His 257, which is displaced outwards <strong>by</strong> several angstroms.<br />

After binding, the electron density becomes well defined. The gate is anchored near the central β sheet<br />

at one end and near the C-terminal helix at the other end. The gate movement is complex and appears to<br />

involve a helical transformation near residues 257–261.<br />

Consequently, initial inhibitor modeling attempts using the native PNP structure were far less successful<br />

than subsequent analyses in which coordinates for the guanine-PNP complex were used. Because of the<br />

magnitude of the changes that occur during substrate binding, it is unlikely that modeling studies <strong>based</strong><br />

on the native structure alone would have accurately predicted the structure of PNP/inhibitor complexes.<br />

The active site is located near the subunit-subunit boundary within the trimer and involves seven<br />

polypeptide segments from one subunit and a short loop from the adjacent subunit (Figure 4D). The<br />

purine binding site employs residues Glu 201, Lys 244, and Asn 243 to form hydrogen bonds with N1,<br />

O6, and N7 of purine. The remainder of the purine binding pocket is largely<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_159.html [4/5/2004 5:02:05 PM]


Document<br />

Page 160<br />

hydrophobic, composed of residues Ala 116, Phe 200, and Val 217. The phosphate binding site uses<br />

residues Ser 33, Arg 84, His 86, and Ser 220 with the phosphate positioned for nucleophilic attack at C1'<br />

of the nucleoside. The sugar binding site is mostly hydrophobic consisting of residues Tyr 88, Phe 200,<br />

His 257 from one subunit and Phe 159 of the adjacent subunit. This hydrophobic pocket orients the<br />

sugar to facilitate nucleophilic attack <strong>by</strong> phosphate and subsequent inversion of C1'.<br />

C. Initial Inhibitor Complexes<br />

In order to understand the interaction of inhibitors with the active site residues, the previously known<br />

inhibitors were obtained and crystallographic analyses were carried out. The most important findings<br />

were (1) 8-amino substituents enhance binding of guanines <strong>by</strong> forming hydrogen bonds with Thr 242<br />

and possibly the carbonyl oxygen atom of Ala 116; (2) substitution <strong>by</strong> hydrophobic groups at the 9position<br />

of a purine enhances binding through interaction with the hydrophobic region of the ribose<br />

binding site; and (3) acyclovir diphosphate is a multisubstrate inhibitor with the acyclic spacer between<br />

the purine N9 and the phosphate of near optimal length to accommodate these two binding sites. Based<br />

on these results, a number of starting compounds were proposed that incorporated these and other<br />

features predicted to enhance inhibitor binding.<br />

III. Molecular Modeling<br />

Structural information in combination with graphical methods for displaying accessible volume,<br />

electrostatic potential, and hydrophobicity of the active site of the target macromolecule greatly<br />

facilitates the drug design process. Accurate prediction of binding affinities and protein conformational<br />

changes are currently not routinely possible, although significant advances are being made.<br />

Proposed compounds were screened <strong>by</strong> modeling the enzyme-inhibitor complex using interactive<br />

computer graphics. Macromodel [21] and AMBER [22] <strong>based</strong> molecular energetics were used along<br />

with Monte Carlo/energy minimization techniques [23] to sample the conformational space available to<br />

potential inhibitors docked into the PNP active site. Methods <strong>based</strong> on the work of Goodford [24]<br />

employing custom software were also used. Qualitative evaluation of the enzyme-inhibitor complexes<br />

<strong>by</strong> molecular graphics and semiquantitative evaluation of the interaction energies between the inhibitors<br />

and the enzyme aided in the prioritization of compounds for chemical synthesis.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_160.html [4/5/2004 5:02:06 PM]


Document<br />

IV. <strong>Drug</strong> <strong>Design</strong> Progression<br />

Page 161<br />

We focused initially on filling the purine binding region of the active site. That done, we planned to fill<br />

the sugar binding region and, finally, the phosphate binding site. We expected that each successive step,<br />

moving the compound closer toward fully occupying the active site, would enhance the affinity of the<br />

drug candidate for the enzyme.<br />

A. The Purine Site<br />

From our crystallographic examinations, we knew that three amino acids in the purine binding pocket of<br />

PNP formed hydrogen bonds with purines and their mimics. Such linkages are among the strongest<br />

reversible chemical bonds that exist. In proposing inhibitor candidates, we concentrated on compounds<br />

that would at least form hydrogen bonds with the same three amino acids. Figure 4e shows a close-up of<br />

the purine site.<br />

We favored exchanging a carbon atom for the nitrogen atom that normally occupies position nine, since<br />

there was no interaction of this nitrogen with the active site and earlier studies showed such a change<br />

promotes binding to PNP. Guanine modified in this way is called 9-deazaguanine. The first structures<br />

selected for synthesis were 9-deazaguanines substituted <strong>by</strong> an arylmethyl group at the 9 position. These<br />

compounds were prepared <strong>by</strong> adaption of a literature procedure [25]. We further expected that attaching<br />

an amino group to the carbon atom in position eight on 9-deazaguanine would enhance affinity, since 8aminoguanine<br />

was the first significant inhibitor of PNP.<br />

Both 8-aminoguanine analogs and 9-deazaguanine analogs are good inhibitors of PNP. However,<br />

introduction of an 8-amino group into the 9-deazaguanine derivatives resulted in decreased potency. To<br />

understand this poor binding, we undertook crystallographic analysis of PNP complexes with four<br />

compounds having the 9-thienyl substituent attached to guanine (G), 8-aminoguanine (8AG), 9deazaguanine<br />

(DG), and 8-amino-9-deazaguanine (8ADG). The results of this analysis are summarized<br />

in Figure 5. These data show one mode of binding for compounds that accept a hydrogen bond from Asn<br />

243 at N7 (G and 8AG) and another for compounds that donate hydrogen to Asn 243 from N7 (DG and<br />

8ADG). The 8AG analogs make use of the Thr 242 side chain to form an additional hydrogen bond,<br />

which improves binding affinity. In the 9-deazaguanine series, where N7 has an attached hydrogen<br />

atom, Asn 243 undergoes a shift that is clearly seen in difference Fourier maps. This shift is caused <strong>by</strong><br />

the formation of the N7-H…OD(243) hydrogen bond. A concomitant shift <strong>by</strong> Thr 242 prevents it from<br />

hydrogen bonding to the 8-amino group of 8ADG. Furthermore, the shift moves the methyl group of Thr<br />

242 towards the 8-amino<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_161.html [4/5/2004 5:02:11 PM]


Document<br />

Figure 5<br />

Comparison of 8-amino and 9-deaza substitutions on guanine. Details are extensively<br />

discussed in the text. Cross-hatching schematically indicates the enzyme. Ball and<br />

stick diagrams show the inhibitor and key side-chain residues. Nitrogens are dark<br />

gray spheres, oxygens are light gray. Arrows indicate hydrogen bonding, with the<br />

arrow size showing relative strength.<br />

amino group, generating a hydrophobic environment for the group and decreasing binding affinity.<br />

Page 162<br />

The carbon-for-nitrogen switch in the 9-deaza variant favors association with PNP <strong>by</strong> substituting a<br />

strong hydrogen bond for the relatively weak one occurring between Asn 243 and guanine. Formation of<br />

a simple 8-aminoguanine variant leads to tight binding <strong>by</strong> giving rise to an extra hydrogen bond between<br />

the purine derivative and Thr 242.<br />

The combination of the two “improvements”—the carbon-for-nitrogen substitution and the addition of<br />

the amino group to position eight—was counterproductive because the carbon in position nine prevented<br />

the amino group at<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_162.html (1 of 2) [4/5/2004 5:02:21 PM]


Document<br />

position eight from forming the extra bond with Thr 242. In fact, it set up an unfavorable, repulsive<br />

clash between the threonine and the added amino group.<br />

Page 163<br />

In the absence of detailed structural information, it would have been extremely difficult to explain why<br />

affixing the amino group to the carbon in position eight proved unhelpful. But crystallography quickly<br />

provided the explanation. 9-Deazaguanine itself would be a better choice for the purine component of an<br />

inhibitor. This experience underscores the wonderful economy of the structure-<strong>based</strong> approach. Without<br />

crystallographic data, we might have pursued a logical but unproductive avenue of research much longer<br />

than we did.<br />

B. Ribose Site<br />

The next task was to fill the sugar binding site. The sugar in a nucleoside does not attach to PNP<br />

primarily <strong>by</strong> forming hydrogen bonds, but through hydrophobic attractions. The sugar binding pocket of<br />

PNP consists of three hydrophobic amino acids: Phe 200 and Tyr 88 from the same monomer that binds<br />

guanine and Phe 159 from the adjacent monomer. Several known inhibitors carried a benzene group<br />

attached to position 9 of the purine in place of the sugar in the nucleoside. An initial series of<br />

compounds was synthesized to exploit the hydrophobic region in the ribose binding site.<br />

A number of 9-substituted 9-deazapurine analogs were prepared with various aromatic, heteroaromatic,<br />

and cycloaliphatic substituents. The first 9-deazaguanine derivatives synthesized, such as 9-benzyl-9deazaguanine,<br />

were three to six times more potent than the most potent known inhibitor, 8-amino-9-(2thienylmethyl)guanine.<br />

The optimum spacer between the purine base and the aromatic substituent<br />

proved to be a single methylene group. Crystallographic data showed that generally the planes of the<br />

aromatic rings tend to orient in a reproducible conformation. The aromatic groups optimize their<br />

interaction with Phe 159 and Phe 200, which results in the classic “herringbone” arrangement reported<br />

in a variety of aromatic systems [26].<br />

Inhibitors with cycloaliphatic substituents at N9 of deazaguanine were also as potent as the aromatic<br />

analogs. The cycloaliphatic substituents occupied the same general volume as the aromatic groups. As<br />

with the aromatic series, the optimum spacer between the 9-deazaguanine and the hydrophobic<br />

substituent is one carbon atom. X-ray analysis of the PNP complexes of 9-cyclohexyl-9-deazaguanine, a<br />

relatively poor inhibitor, and the complex of 9-cyclohexylmethyl-9-deazaguanine, a potent inhibitor,<br />

showed the two cyclohexyl groups occupy approximately the same space in the active site with the<br />

purine base pulled out of its optimal position in the former.<br />

The chemistry is more straightforward with the aromatic series. From modeling studies, we saw that the<br />

sugar binding pocket could be filled more<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_163.html [4/5/2004 5:02:22 PM]


Document<br />

completely <strong>by</strong> adding any of several chemical groupings to the benzene ring. The best fit came from<br />

adding a chlorine atom to position 3 of the benzene ring.<br />

C. Phosphate Site<br />

Page 164<br />

The final step added a group that would interact with the phosphate binding site either directly or via<br />

electrostatic interactions. We could not use phosphate itself, because phosphate-containing compounds<br />

are not metabolically stable and have difficulty passing through cell membranes intact. Acyclovir<br />

diphosphate, which is not membrane permeable and is subject to extracellular metabolism, is a good<br />

example.<br />

Our results suggested that an ideal PNP inhibitor in the 9-deazapurine series would contain an aromatic<br />

group and a substituent with affinity for the phosphate site interlinked <strong>by</strong> spacers with optimum lengths.<br />

Crystallographic and modeling studies suggested a two-to-four-atom spacer. Initial modeling studies<br />

encouraged us to prepare several structures, but they failed to improve the binding affinity of our twopart<br />

structure.<br />

Crystallographic analysis of a number of PNP inhibitor complexes revealed significant displacement of<br />

the inhibitors. These displacements appear to be the result of close contacts between the inhibitor and<br />

the ion in the phosphate binding site. Sulfate ions occupy the phosphate site in PNP crystals as they are<br />

grown from ammonium sulfate solution. These inhibitors were more potent when the binding was<br />

measured in 1 mM phosphate solution rather than in 50 mM phosphate. Kinetic studies showed that<br />

these inhibitors were competitive not only with inosine but also with phosphate, in keeping with the<br />

above observation.<br />

These results, summarized in Table 1, show that the IC 50 (50 mM) is equal to or larger than the IC 50 (1<br />

mM), in some cases <strong>by</strong> as much as 100-fold. The ratio and the dimension of the 9-substituent show some<br />

correlation. Compounds such as 8-aminoguanosine and 8-amino-9-(2-thienylmethyl)guanine show no<br />

difference. Since the concentration of phosphate in intact cells is 1 mM, we routinely used this assay<br />

condition for all PNP inhibitors.<br />

Starting with a model of the 9-benzyl-9-deazaguanine/PNP complex, we concluded that two of the<br />

positions on the 9-benzyl group, namely the 2-position of the phenyl ring and one of the benzylic sites,<br />

appeared to be oriented so that a group attached to either one could interact favorably with the phosphate<br />

binding site. The first compound made in this series, 9-[2-(3-phosphonopropoxy)benzyl]guanine, turned<br />

out to be a poor PNP inhibitor. Subsequent crystallographic analysis revealed that the plane of the<br />

aromatic ring had rotated approximately 90° from its optimum position in the hydrophobic pocket. This<br />

reorientation of the ring was presumably necessary to accommodate the four atom spacer between the<br />

phenyl ring and the phosphonate group. A compound<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_164.html [4/5/2004 5:02:25 PM]


Document<br />

Table 1 Inhibition Data for Selected PNP Inhibitors <strong>by</strong> Increasing IC 50 Value<br />

R 2a<br />

R 2<br />

IC 50, μM Ratio d<br />

50 mM<br />

phosphate b<br />

1 mM<br />

phosphate c<br />

(S)-3-Chlorophenyl CH 2CO 2H 0.031 0.0059 5.3<br />

3-Chlorophenyl CH 2CN 1.8 0.010 180<br />

2-Tetrehydrothienyl H 0.22 0.011 20<br />

3,4-Dichlorophenyl H 0.25 0.012 21<br />

3-Thienyl H 0.08 0.020 4<br />

3-Trifluoromethylcyclohexyl H 0.74 0.020 37<br />

Cyclopentyl H 1.8 0.029 62<br />

Cycloheptyl H 0.86 0.030 29<br />

Pyridin-3-yl H 0.20 0.030 7.3<br />

2-(Phosphonoethyl)phenyl e H 0.45 0.035 13<br />

Cyclohexyl H 2.0 0.043 47<br />

2-Furanyl H 0.31 0.085 3.6<br />

(R)-3-Chlorophenyl CH 2CO 2H 0.90 0.16 5.6<br />

2-Phosphonopropoxyphenyl e H 42 1.0 42<br />

a Compounds with R2 not equal to H are racemic mixtures unless the R or S isomer is designated.<br />

b Calf spleen PNP assayed in 50 mM phosphate buffer.<br />

c Calf spleen PNP assayed in 1 mM phosphate buffer.<br />

d IC50 at 50 mM phosphate divided <strong>by</strong> IC 50 at 1 mM phosphate.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_165.html (1 of 2) [4/5/2004 5:02:28 PM]<br />

Page 165


Document<br />

e Guanine base.<br />

Source: Ref. 27.<br />

with a two-carbon spacer was a much better PNP inhibitor; however, it was clear from x-ray analysis<br />

that the aromatic ring was unable to form the ideal “herringbone” packing interaction.<br />

Alternatively, compounds were modeled in which the spacer to the phosphate binding site branched<br />

from the benzylic carbon, thus placing no restrictions on the tilt of the aromatic ring. Examination of the<br />

9-benzyl-9-deazaguanine/PNP complex indicated that of the two benzylic positions, one (pro-R) pointed<br />

into a sterically crowded area within the active site, whereas the other (pro-S) pointed into a relatively<br />

empty space adjacent to the phosphate binding site. This analysis led to the synthesis of racemic 9-[1-(3chlorophenyl)-2-carboxyethyl]-9-deazaguanine.<br />

This compound adds an acetate group (CH 2COO-) to<br />

the methylene carbon atom that joined 9-deazaguanine to the chlorinated benzene ring. This compound<br />

was resolved into its (S) and (R) enantiomers.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_165.html (2 of 2) [4/5/2004 5:02:28 PM]


Document<br />

Page 166<br />

As predicated, the (S) acid was a 30-fold more potent inhibitor of PNP than the (R) form. X-ray<br />

crystallographic analysis of the complexes revealed that the (S) acid was oriented properly for optimal<br />

interactions with all three subsites (Figure 4F), whereas the (R) acid was not. This series of compounds<br />

contains the most potent membrane-permeable inhibitors of PNP yet reported [27].<br />

V. <strong>Summary</strong><br />

Recently, scientists at BioCryst have successfully completed a project to design and synthesize potent<br />

inhibitors of the enzyme Purine Nucleoside Phosphorylase (PNP) using the three-dimensional structure<br />

of the active site. Crystallographic and modeling methods have been combined with organic synthesis to<br />

produce inhibitors. Our experience in creating a set of potential drugs—one of which (BCX-34) is now<br />

in human trials for treating psoriasis and a form of T-cell lymphoma—illustrates the process and the<br />

power of structure-<strong>based</strong> design.<br />

This structure-<strong>based</strong> inhibitor design approach led to a number of inhibitors more than 100 times more<br />

potent than any membrane-permeable inhibitor available at the beginning of this project. During the two<br />

and half years of this project, about 60 active compounds were synthesized. This is a remarkably small<br />

number compared with the extensive synthesis programs generally involved in drug discovery <strong>by</strong> trial<br />

and error techniques. The large number of active compounds and the enhancement of inhibitor potency<br />

stand as proof that crystallographic and modeling techniques are now capable of playing a critical role in<br />

the rapid discovery of novel therapeutic agents. The entire protocol, from choosing the target to creating<br />

a drug suitable for clinical trials, can probably be accomplished today in two or three years.<br />

A. Obstacles Encountered and Lessons Learned<br />

Crystallographic analysis was <strong>based</strong> primarily on the results of difference Fourier maps in which the<br />

interactions between residues in the active site and the inhibitor could be characterized. During these<br />

studies, about 35 inhibitor complexes were evaluated <strong>by</strong> x-ray crystallographic techniques. It is<br />

noteworthy that the resolution of the PNP model extends to only 2.8 Å and that all of the difference<br />

Fourier maps were calculated at 3.2 Å resolution, much lower than often considered essential for drug<br />

design. Crystallographic analysis was facilitated <strong>by</strong> the large solvent content that allowed for free<br />

diffusion of inhibitors into enzymatically active crystals.<br />

Initial inhibitor modeling attempts using the native PNP structure were far less successful than<br />

subsequent analyses in which coordinates for the guanine-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_166.html [4/5/2004 5:02:29 PM]


Document<br />

Page 167<br />

PNP complex were used, mainly because of the magnitude of the changes that occur during substrate<br />

binding. We found that computer modeling required significant tuning in order to provide useful results.<br />

Crystallographic results were useful in testing and modifying modeling parameters. The most useful<br />

modeling results were achieved after incorporation of the conformational searching techniques described<br />

earlier and when the coordinates for the PNP-guanine complex model were used. Visual inspection and<br />

chemical intuition were very important.<br />

B. Perspectives of Treating Targeted Disease<br />

One of the inhibitors designed during the drug discovery process, 9-(3-pyridylmethy)-9-deazaguanine<br />

(BCX-34), was selected for initial clinical development. Current clinical trials utilize both topical and<br />

oral formulations of the drug.<br />

Researchers at the University of Alabama at Birmingham and Washington University School of<br />

Medicine have recently completed small Phase II clinical trials of two indicated applications, cutaneous<br />

T-cell lymphoma (CTCL) and psoriasis, using a topical formulation of BCX-34. Although patients<br />

showed improvement in both trials, the duration of each was too short (six weeks) to adequately assess<br />

the efficacy of the drug. Subsequently 80% of the patients from the CTCL trial (24 patients) entered an<br />

open label trial for treatment of their disease for up to twelve months. At the end of the first six months<br />

of treatment, seven of the patients were in complete remission (verified <strong>by</strong> biopsy), two patients showed<br />

a clinical complete response, and nine patients had shown definite improvement. The other six patients<br />

had shown no change or progression of disease. No serious, drug-related adverse events were reported<br />

during the study.<br />

The process of structure-<strong>based</strong> drug design helped to ensure that the inhibitor would be highly selective<br />

for the PNP enzyme, and thus far no other targets for the drug have been identified. The mechanism of<br />

action of BCX-34 appears to be entirely related to its effect on the proliferation of human T-cells. This<br />

high degree of specificity probably also contributes to the high safety profile of the drug. Although longterm<br />

studies in more patients will be necessary to substantiate these results, it appears likely that BCX-<br />

34 will have a significant clinical effect on at least some T-cell mediated diseases.<br />

Based on the results from these three trials, BioCryst has initiated a multicenter Phase III trial for the<br />

treatment of CTCL, as well as a large, multicenter Phase II trial for psoriasis. In addition to the two<br />

clinical trials using the topical formulation, a Phase I clinical trial in CTCL and T-cell<br />

lymphoma/leukemia has begun using an oral formulation of BCX-34. In the future, a number of other Tcell<br />

mediated diseases or processes are possible targets for BCX-34, including rheumatoid arthritis,<br />

multiple sclerosis, inflammatory bowel disease, and organ transplant rejection.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_167.html [4/5/2004 5:02:31 PM]


Document<br />

References<br />

1. Parks RE Jr., Agarwal RP. In: Boyer PD, ed. The Enzymes. 3rd Ed., New York: Academic, 1972;<br />

7:483–514.<br />

Page 168<br />

2. Stoeckler JD. In: Glazer, RE, ed. Developments in Cancer Chemotherapy. Florida: CRC, Baco Raton,<br />

1984; 35–60.<br />

3. Friedkin M, Kalckar H. In: Boyer PD, Lardy H, Myrback K, eds. The Enzymes. 2nd Ed. New York:<br />

Academic, 1961; 5:237–55.<br />

4. Agarwal KC, Agarwal RP, Stoeckler JD, Parks RE Jr. Purine nucleoside phosphorylase.<br />

Microheterogeneity and comparison of kinetic behavior of the enzyme from several tissues and species.<br />

Biochemistry 1975; 14:79–84.<br />

5. Bzowska A, Kulikowska E, Shugar D. Properties of purine nucleoside phosphorylase (PNP) of<br />

mammalian and bacterial origin. Z Naturforschung C Biosci 1990; 45:59–70.<br />

6. Stoeckler JD, Agarwal RP, Agarwal KC, Schmid K, Parks RE Jr. Purine nucleoside phosphorylase<br />

from human erythrocytes: physiocochemical properties of the crystalline enzyme. Biochemistry 1978;<br />

17:278–83.<br />

7. Williams SR, Goddard JM, Martin DW Jr. Human purine nucleoside phosphorylase cDNA sequence<br />

and genomic clone characterization. Nucleic Acids Res 1984; 12:5779–87.<br />

8. LePage GA, Junga IG, Bowman B. Biochemical and carcinostatic effects of α'-deoxythiguanosine<br />

Cancer Res 1964; 24:835–40.<br />

9. Giblett ER, Ammann AJ, Wara DW, Sandman R, Diamond LK. Nucleoside-phosphorylase deficiency<br />

in a child with severely defective T-cell immunity and normal B-cell immunity. Lancet 1975; 1:1010–3.<br />

10. Otterness I, Bilven M. In: Rainsford K, Velo G, eds. New Developments in Antirheumatic Therapy.<br />

Inflammation and <strong>Drug</strong> Therapy Series. Norwell, MA: Kluwer Academic, 1989; 2:277–304.<br />

11. Stoeckler JD, Cambor C, Kuhns V, Chu SH, Parks RE Jr. Inhibitors of purine nucleoside<br />

phosphorylase, C(8) and C(5') substitutions. Biochemical Pharmacology 1982; 31:163–71.<br />

12. Shewach DS, Chern JW, Pillote KE, Townsend LB, Daddona PE. Potentiation of 2'-deoxyguanosine<br />

cytotoxicity <strong>by</strong> a novel inhibitor of purine nucleoside phosphorylase, 8-amino-9-benzylguanine. Cancer<br />

Res 1986; 46:519–23.<br />

13. Stoeckler JD, Ryden JB, Parks RE Jr, Chu MY, Lim MI, Ren WY, Klein RS. Inhibitors of purine<br />

nucleoside phosphorylase: effects of 9-deazapurine ribonucleosides and synthesis of 5'-deoxy-5'-iodo-9deazainosine.<br />

Cancer Res 1986; 46:1774–8.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_168.html (1 of 2) [4/5/2004 5:02:33 PM]


Document<br />

14. Tuttle JV, Krenitsky TA. Effects of acyclovir and its metabolites on purine nucleoside<br />

phosphorylase. J Biol Chem 1984; 259:4065–9.<br />

15. Gilbertsen RB, Scott ME, Dong MK, Kossarek LM, Bennett MK, Schrier DJ, Sircar JC. Preliminary<br />

report on 8-amino-9-(2-thienylmethyl) guanine (PD 119,229), a novel and potent purine nucleoside<br />

phosphorylase inhibitor. Agents and Actions 1987; 21:272–4.<br />

16. Cook WJ, Ealick SE, Bugg CE, Stoeckler JD, Parks RE Jr. Crystallization and preliminary X-ray<br />

investigation of human erythrocytic purine nucleoside phosphorylase. J Biol Chem 1981; 256:4079–80.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_168.html (2 of 2) [4/5/2004 5:02:33 PM]


Document<br />

17. Ealick SE, Rule SA, Carter DC, Greenhough TJ, Babu YS, Cook WJ, Habash J, Helliwell JR,<br />

Stoeckler JD, Parks RE Jr, Chen SF, Bugg CE. Three-dimensional structure of human erythrocytic<br />

purine nucleoside phosphorylase at 3.2 Å resolution. J Biol Chem 1990; 265:1812–20.<br />

Page 169<br />

18. Narayana SVL, Bugg CE, Ealick SE. Refined structure of purine nucleoside phosphorylase at 2.75 Å<br />

resolution. Acta Cryst D 1996; accepted.<br />

19. Babu YS, Refined structure of guanine: purine nucleoside phosphorylase at 2.8 Å resolution. In<br />

preparation.<br />

20. Koshland DE Jr. The key-lock theory and the induced fit theory. Angew Chem Int Ed Engl 1994;<br />

33:2375–8.<br />

21. Mohamadi F, Richards NGJ, Guida WC, Liskamp R, Lipton M, Caufield C, Change G, Hendrickson<br />

T, Still WC. MacroModel—An integrated software system for modeling organic and bioorganic<br />

molecules using molecular mechanics. J Comput Chem 1990; 11:440–67.<br />

22. Weiner SJ, Kollman PA, Case DA, Singh UC, Ghio C, Alagona S, Profeta S, Weiner P. New force<br />

field for molecular mechanical calculations simulations of proteins and nucleic acids. J Am Chem Soc<br />

1984; 106:765–84.<br />

23. Chang G, Guida WC, Still WC. An internal coordinate monte carlo method for searching<br />

conformational space. J Am Chem Soc 1989; 111:4379–86.<br />

24. Goodford P. A computational procedure for determining energetically favorable binding sites on<br />

biologically important macromolecules. J Med Chem 1985; 28:849–57.<br />

25. Montgomery JA, Niwas S, Rose JD, Secrist JA 3d., Babu YS, Bugg CE, Erion MD, Guida WC,<br />

Ealick SE. <strong>Structure</strong>-<strong>based</strong> design of inhibitors of purine nucleoside phosphorylase. 1. 9-(arylmethyl)<br />

derivatives of 9-deazaguinine. J Med Chem 1993; 36:55–69.<br />

26. Burley SK, Petsko GA. Aromatic-aromatic interaction: a mechanism of protein structure<br />

stabilization. Science 1985; 229:23–8.<br />

27. Ealick SE, Babu YS, Bugg CE, Erion MD, Guida WC, Montgomery JA, Secrist JA 3d. Application<br />

of crystallographic and modeling methods in the design of purine nucleoside phosphorylase inhibitors.<br />

PNAS USA 1991; 88:11540–4.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_169.html [4/5/2004 5:02:34 PM]


http://legacy.netlibrary.com/reader/message.asp?message=811&BookID=12640&FileName=Page_170.html<br />

The requested page could not be found.<br />

Return to previous page<br />

http://legacy.netlibrary.com/reader/message.asp?message=811&BookID=12640&FileName=Page_170.html [4/5/2004 5:02:36 PM]


Document<br />

6<br />

Structural Implications in the <strong>Design</strong> of Matrix-Metalloproteinase<br />

Inhibitors<br />

John C. Spurlino<br />

3-Dimensional Pharmaceuticals, Inc., Exton, Pennsylvania<br />

I. Matrix-Metalloproteinases<br />

Page 171<br />

The matrix metalloproteinases (MMPs) are a family of ubiquitous enzymes that are involved in<br />

extracellular matrix degradation and remodeling. They are critical for the processes of morphogenesis<br />

and wound healing, but are also implicated in many human diseases including arthritis, metastasis, and<br />

cancer tumor growth [1, 2]. This family includes matrilysin, fibroblast collagenase (HFC), neutrophil<br />

collagenase (HNC), stromelysin 1 (HFS), stromelysin-2, stromelysin-3, gelatinases A and B, collagenase-<br />

3, and the membrane type MMP. In addition to the destructive involvement in diseases, MMPs play a<br />

critical role in the remodeling of the extracellular matrix [3].<br />

The MMP enzyme family is part of the superfamily of metzincins. The metzincin superfamily is<br />

distinguished <strong>by</strong> a conserved zinc binding motif for the catalytic zinc and a Met-turn region [4]. The<br />

MMPs are unique in that they also contain a second structural zinc, however this zinc may be absent in<br />

the intact full-length enzyme [5]. The presence of one to four structural calcium ions has been detected<br />

in the MMPs that have been characterized to date. The importance of the zinc ions and at least one of the<br />

structural calcium ions to enzymatic activity has been proven [6].<br />

The MMPs are secreted as inactive proenzymes, which are activated <strong>by</strong> proteolytic cleavage. Once<br />

activated they are subject to control <strong>by</strong> tissue inhibitors of metalloproteinases (TIMPs). It is the<br />

imbalance between the active enzymes and the TIMPs that leads to destructive tissue degradation that<br />

potent directed pharmaceuticals can overcome. These enzymes have been the target of<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_171.html [4/5/2004 5:02:38 PM]


Document<br />

drug design since the late 1970s [7]. Batimastat, [{4-N-hydroxyamino}-2R-isobutyl-3S-{thienylthiomethyl}<br />

succinyl]-L-phenyl-alanine-N-methylamide, a potent nonspecific MMP inhibitor from<br />

British Bio-Tech, is now in Phase III trials. The information gained from current studies indicate that<br />

there is some efficacy in the treatment of disease states <strong>by</strong> MMP inhibitors [8–10].<br />

Page 172<br />

There is still much debate about whether a broad spectrum or directed MMP inhibitor is the best course<br />

of treatment for a variety of diseases, partly because the exact role of individual MMPs is still unclear.<br />

Both collagenase-3 and HFC are suspected to cause osteoarthritis [11]. It is currently believed that<br />

gelatinase and collagenase-3 have a role in breast cancer [12]. Gelatinase A and B have been implicated<br />

in hemorrhagic brain injury [13]. Gelatinase A and B, HFC, and stromelysin may all be involved in<br />

gastric cancer [14]. Matrilysin may be implicated in colon cancer [15]. Increased gelatinase A and B<br />

activity has also been seen in response to beta-amyloid production [16]. The role that individual MMPs<br />

play in causing these diseases, however, remains unclear. This uncertainty underscores the need to<br />

develop selective inhibitors of individual MMPs to ferret out the roles each play in the development of<br />

specific disease states.<br />

The MMPs consist of one or more structural domains (Figure 1). The first domain, the propeptide<br />

domain, confers a self-inhibitory action on the full-length MMP. The second domain contains the active<br />

site residues and is referred to as the catalytic domain. The catalytic domain is characterized <strong>by</strong> the<br />

conserved zinc-binding sequence (HEXGHXXGXXHS), which also contains the glutamate residue that<br />

is essential for activity [17]. The MMPs are activated <strong>by</strong> cleavage of the prodomain. All MMPs contain<br />

these first two domains. Matrilysin, the simplest of the MMPs, is an example of a two-domain enzyme,<br />

where the active enzyme consists solely of the catalytic domain.<br />

The remainder of the MMPs also contain a hemopexin-like domain connected to the catalytic domain <strong>by</strong><br />

a proline-rich linker. This domain is involved in the interaction of the collagenases and stromelysins<br />

with collagen and, in the case of the collagenases, is essential for activity against collagen [18]. The<br />

cleavage of the proline-rich linker region in HFC and HNC is another route to control collagenase<br />

activity. The hemopexin domain of the gelatinases is not necessary for collagen binding, but may be<br />

involved in receptor recognition [19].<br />

The gelatinases also contain a fibronectin-like insert in the catalytic domain that is involved in binding<br />

collagen [20]. The fibronectin domain has also been shown to be essential for elastolytic activity [21]. A<br />

structural picture of these additional domains is essential for an understanding of the mode of action for<br />

these larger MMPs, but is not necessary for a structure-<strong>based</strong> drug design strategy. Differences in the<br />

catalytic domains of the MMPs can be used to drive a targeted drug discovery effort.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_172.html [4/5/2004 5:02:39 PM]


Document<br />

Figure 1<br />

A ribbon model of the full-length collagenase structure (1fbl.pdb). The<br />

prodomain would precede and include the portion labeled in the figure. The<br />

catalytic domain is shown, with a highlighted region where the fibronectin-like<br />

domain of the gelatinases is inserted.<br />

II. 3-Dimensional <strong>Structure</strong> of MMPs<br />

Page 173<br />

Catalytic domain structures for fibroblast collagenase [22–25], neutrophil collagenase [26,27],<br />

matrilysin [28], and stomelysin [29, 30] have all been determined and deposited in the Protein Data<br />

Bank [31]. The catalytic domains of MMPs, as seen in the archetypal collagenase structure (shown as a<br />

ribbon drawing [32] in Figure 1), consist of an upper 5-stranded β sheet flanked <strong>by</strong> two α helices on one<br />

side of the active site cleft and a long loop that contains the Met-turn flanked <strong>by</strong> a single α helix on the<br />

other side of the cleft.<br />

The active-site groove as seen in the solvent-accessible surface [33] is an obvious structural feature<br />

(Figure 2). The top wall of the cleft (as seen in Figure 2) is formed <strong>by</strong> the top strand of the β sheet and<br />

the loop that contains the calcium binding site. The lower wall of the cleft is formed from the residues<br />

on either side of the Met-turn. These residues can be considered as an interrupted strand which, together<br />

with the substrate, complete the twisted β sheet of the catalytic domain. The bottom of the cleft is<br />

formed <strong>by</strong> the second helix, which<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_173.html (1 of 2) [4/5/2004 5:02:46 PM]


Document<br />

Figure 2<br />

The accessible surface of HFC with a modeled substrate from human collagen<br />

showing the binding sites.<br />

contains the HExxH motif, the catalytic zinc, and the S1' pocket. Substrates bind in an extended<br />

conformation that approximates an antiparallel strand. The cleft, however, is not large enough to<br />

accommodate a triple helix collagen molecule.<br />

Page 174<br />

A structure for the full-length active porcine synovial collagenase [34] has been determined. The<br />

structure of the catalytic domain of this full-length enzyme is equivalent to the structures of the isolated<br />

catalytic domains of HFC, HNC, and matrilysin. The flexible linker domain between the catalytic and<br />

hemopexin domains is disordered and the orientation of the hemopexin domain in the structure offers no<br />

real clue as to the mode of action for the full-length collagenases. Furthermore, the matrilysin structure<br />

of the full-length active enzyme has almost identical secondary structural features (a Ca overlap of<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_174.html (1 of 2) [4/5/2004 5:02:54 PM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_175.html (1 of 2) [4/5/2004 5:03:13 PM]<br />

Page 175


Document<br />

Figure 3<br />

The sequence alignment of MMPs with the catalytic domain region highlighted.<br />

The residues that line the subtrate pockets are marked: S3 (3), S2 (2), S1 (1), S1<br />

(*), S2' (@), and S3' (#). The highlighted catalytic domain alignment was dominated <strong>by</strong><br />

the structural alignment of the determined structures of HFC, HNC, and matrilysin.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_175.html (2 of 2) [4/5/2004 5:03:13 PM]


Document<br />

Page 176<br />

0.43 Å) as the catalytic-domain structure of HFC. This demonstrates that the absence or presence of the<br />

hemopexin domain does not affect the overall structure of the catalytic domain.<br />

The sequence homology of the catalytic domains of the collagenases is 62%. This can be extended to the<br />

other members of the MMP family as seen in Figure 3. An understanding of the structural features of the<br />

target enzyme is essential for structure-<strong>based</strong> drug design. In this example we will be looking at<br />

inhibiting MMPs <strong>by</strong> binding to the active site. The numbering system used throughout this chapter will<br />

be in regard to the HFC sequence used in 1hfc.pdb.<br />

III. Surface Features<br />

The surface features of matrilysin, fibroblast collagenase, and neutrophil collagenase are all similar<br />

(Figure 4). The active-site groove can be plainly seen on the surface: two main pockets punctuated <strong>by</strong><br />

the active-site zinc. Modeling<br />

Figure 4<br />

The accessible surfaces of HFC (a), HNC (b), and matrilysin<br />

(c) are shown with a bound P'-side hydroxamate inhibitor.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_176.html (1 of 2) [4/5/2004 5:03:26 PM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_177.html (1 of 2) [4/5/2004 5:03:36 PM]<br />

Page 177


Document<br />

Page 178<br />

studies show that there is not sufficient room in the S1'-S3' cleft to accommodate the native coiled triple<br />

collagen bundle.<br />

The active site consists of a series of subsites on either side of the catalytic zinc. These subsites are<br />

numbered starting at the catalytic zinc and preceding from N to C as S1', S2', etc., corresponding to the<br />

residues (P1', P2', etc.) of the substrate that is bound (Figure 2). Likewise the subsites are numbered S1,<br />

S2, etc. outward from the other side of the catalytic zinc.<br />

While the interaction of the substrate (and the inhibitor) with the catalytic zinc is the most important<br />

interaction, the remainder of the substrate (inhibitor) also forms hydrogen bonds with residues from the<br />

top strand of the β sheet and the loop region posterior to the Met-turn. These interactions with the<br />

substrate in the binding pockets of the MMPs are the prime targets for engineering specific MMP<br />

inhibitors. An in-depth understanding of the differences of the properties of these pockets in the different<br />

MMPs and the interactions of specific residues within these pockets is essential for structure-<strong>based</strong><br />

design of inhibitors.<br />

IV. Main-Chain Substrate Interactions<br />

Most of the hydrogen bonds between the substrate and the MMP occur with the top strand of the β sheet.<br />

The P3 residue does not make any direct hydrogen bonds with the MMP. The P2 residue makes two<br />

hydrogen bonds with residue 184, which is a conserved alanine residue in all the aligned MMP<br />

sequences. Residue 183 is a conserved histidine, which is bound to the structural zinc, further stabilizing<br />

the conformation of the top strand. The carbonyl oxygen of P1 is liganded to the catalytic zinc. The lefthand<br />

side of the substrate is thus held in place <strong>by</strong> only two hydrogen bonds with the enzyme and one<br />

interaction with the catalytic zinc. Although the P3 residue does not make any hydrogenbond<br />

contributions to substrate binding, it is essential for catalytic activity [43].<br />

The right-hand side of the substrate is held much tighter. The P1' residue's carbonyl oxygen makes a<br />

hydrogen bond with the amide nitrogen of residue 181, which is a conserved leucine residue. The amide<br />

nitrogen of the P1' residue is hydrogen bonded to the conserved alanine-182 carbonyl oxygen. The P2'<br />

substrate residue is held in place <strong>by</strong> hydrogen bonds to proline 238 and tyrosine 240, two more<br />

conserved residues. The amide nitrogen of P3' makes a hydrogen bond with the carbonyl oxygen of<br />

residue 179, the only nonconserved residue, making a main-chain interaction with the substrate.<br />

The use of conserved residues to maintain the main-chain interactions along the substrate backbone<br />

makes differentiation of the MMPs through these interactions difficult. Instead, differences in the<br />

regions where the side chains of the substrate interact can be used to drive the discovery of specific<br />

MMP inhibitors. The hydrogen-bonding pattern also indicates that right-hand side (P'-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_178.html [4/5/2004 5:03:38 PM]


Document<br />

Page 179<br />

side) inhibitors will bind with greater affinity. Indeed, most of the structures of MMPs were determined<br />

with right-hand side inhibitors, and most of the pharmaceuticals currently in development are also righthand<br />

side binders. A closer look at the binding pockets themselves also demonstrates the reasons for the<br />

preference of researchers for the right-hand side.<br />

V. Substrate-Binding Pockets<br />

The nonprimed or left-hand side of the cleft consists of a large shallow depression. The S1 pocket<br />

consists of a shallow ridge that complements the glycine residue of the collagen strands. Most of the<br />

interactions with the glycine residue are brought about due to its interaction with the catalytic zinc.<br />

Asparagine 180 approaches the P1 residue in HFC. Crystallographic evidence indicates an interaction of<br />

the thiophene ring of batimastat via electrostatic interactions of the p orbitals and the catalytic zinc and<br />

the possibility of a water-mediated hydrogen bond to the carbonyl O of residue 184 [35]. Larger<br />

substituents can be accommodated in regions adjacent to the P1 pocket, possibly in the large pocket<br />

above the S1 site (see Figure 2). Increased potency was noted for several compounds with cyclic imido<br />

P1 substituents that could bind here [2].<br />

The S2 pocket is a large shallow depression offering no real binding cavities. One side of the pocket is<br />

made up from the conserved histidine at position 228 and the main chain from residue 227. The bottom<br />

of the pocket is formed <strong>by</strong> histidine 222. Both of these histidines are liganded to the catalytic zinc. The<br />

other side consists mostly of the residue 186 side chain with some hydrogen-bonding contacts possible<br />

from the tip of the glutamine side chain in the case of HFC and HNC.<br />

The S3 pocket offers a shallow cavity to bind the conserved proline. The proline residue of the substrate<br />

would lie between the side chains of His 183, Phe 185, and Ser172 [36]. Residues 183 and 185 are<br />

conserved among the MMPs with the minor exception of a tyrosine replacing phenylalanine 164 in<br />

stromelysin. Residue 172 shows some variability among the MMPs existing as a serine in HNC and<br />

HFC and a tyrosine in the remainder of the aligned MMPs.<br />

The primed or right-hand side of the active site exists as a narrow canyon with a large well at the<br />

beginning. The S1' pocket is a narrow, deep cavity providing an ideal binding site for inhibitor design.<br />

The S1' pocket is the most significant feature of the surface, extending as a tunnel completely through<br />

the enzyme in the case of neutrophil collagenase and stromelysin. This feature makes the S1' pocket an<br />

ideal candidate for use in designing an inhibitor with specificity for HNC, HFC, or HFS.<br />

The volumes of the S1' pockets vary greatly. Matrilysin has the smallest S1' pocket at 111 Å 3. The<br />

fibroblast collagenase pocket is not much larger at<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_179.html [4/5/2004 5:03:39 PM]


Document<br />

Figure 5<br />

(a) A cut away view of the S1' pocket of HFC showing the termination of<br />

the pocket <strong>by</strong> Arg214. Matrilysin also has a truncated pocket. (b) A cut away<br />

view of the S1' pocket of HNC showing the clear path through to the other<br />

side of the enzyme. The gelatinases are also likely to have large extended S1'<br />

pockets.<br />

123 Å 3. The pocket of the neutrophil collagenase that travels through the enzyme has a volume of 305<br />

Å 3. Figure 5 demonstrates the differences in the relative sizes of the S1' pockets of HFC and HNC.<br />

Stromelysin has a pocket that should be of similar size as that of neutrophil collagenase [21]. The<br />

gelatinases and collagenase-3, <strong>based</strong> on sequence alignment, should also posses long tunnel-like S1'<br />

pockets.<br />

Page 180<br />

The residues that line the S1' pocket are mostly hydrophobic residues (see Figure 3) and show a<br />

remarkable overall similarity. The specificity of a number of inhibitors of MMPs can be linked to<br />

differences in the S1' pockets. The S1' pocket of HFC is terminated <strong>by</strong> arginine 214, while matrilysin<br />

has a tyrosine residue that accomplishes the same thing. The remainder of the aligned MMPs have<br />

leucine residues at that position. In addition the conformation of the leucine residue is swung back,<br />

forming the tunnel. There are three additional residues that are significant in their differences within the<br />

S1' pocket: residues 239 and a two-residue insert, relative to HFC, after residue 242. These residues<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_180.html (1 of 2) [4/5/2004 5:03:48 PM]


Document<br />

Page 181<br />

form the lower end of the tunnel. The major effect of the different residues found here is on the diameter<br />

of the exit hole in the S1' tunnel. However, there are some residues that could be targeted for hydrogenbond<br />

formation.<br />

The S2' binding cavity is a narrow cleft that can easily accommodate a peptide backbone, but with no<br />

room for a side chain. The interaction of the P2' side chain is made with the exterior surface of the<br />

enzyme. The S2' site is exposed to solvent and presents two possible interaction sites for bound<br />

inhibitors that are related <strong>by</strong> a rotation about χ1. These sites consist of residues 179–180 on one side and<br />

residues 238–240 on the other.<br />

The S3' binding cavity opens up out of the canyon and consists almost entirely of surface interactions.<br />

The side chain of P3' interacts with residues 210 and 240, which are mostly conserved tyrosine residues.<br />

Additional interactions could be formed with some of the additional residues found in the insertions<br />

after residue 242.<br />

VI. <strong>Structure</strong>-Based <strong>Design</strong><br />

<strong>Structure</strong>-<strong>based</strong> drug design is an iterative process that starts with a lead compound, a structural model<br />

of the target, and a structure-activity relationship<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_181.html (1 of 2) [4/5/2004 5:03:56 PM]


Document<br />

Page 182<br />

(SAR) model. The lead compound can come from compound screening, previously discovered<br />

inhibitors, or it can be <strong>based</strong> on a known substrate. The model can be obtained from x-ray<br />

crystallography, high-field nuclear magnetic resonance spectroscopy, or from homology-<strong>based</strong> model<br />

building. Inhibitor structures are developed and docked into the model of the binding site of interest,<br />

typically the active site of the enzyme. The interactions of the inhibitor-enzyme complex are evaluated<br />

and ranked. The most promising compounds are then synthesized and tested. Based on the results of the<br />

testing, additional enzyme-inhibitor structures are determined, the SAR model is updated, and the<br />

process beings again.<br />

As a model case of structure-<strong>based</strong> drug design for MMPs we will look at the design of a right-handed<br />

inhibitor <strong>based</strong> on the x-ray structures of HFC and HNC.<br />

VII. Zinc-Binding Group<br />

The design of active-site inhibitors <strong>based</strong> on the natural substrate of the collagenases has produced a<br />

variety of zinc-binding groups to anchor the inhibitor to the catalytic zinc. These group include<br />

hydroxamates, thiols, phosphorous acid derivatives (phosphinate, phosphonate, phosphoramidate), and<br />

carboxylates. The selection of a suitable zinc-binding group has been studied in depth [37–40]. The most<br />

potent zinc-binding group found for the collagenases to date is the hydroxamate.<br />

The structural comparison of hydroxamate, carboxylate, and sulfodiimine in matrilysin provided<br />

information on the contribution of the zinc ligand to the overall potency of the inhibitor [19]. The<br />

potency of the zinc-bind group can be directly related to the number of bonds in which it is involved for<br />

this instance. The hydroxamate is the perfect bidentate ligand to the zinc with both oxygens being within<br />

2.2 Å of the zinc. The hydroxamate group also is involved in hydrogen bonds with Glu219 and the<br />

carbonyl oxygen of Ala182. The carboxylate group is also a bidentate ligand to zinc, however the<br />

oxygens are not equidistant from the zinc. The carboxylate forms only one additional hydrogen bond<br />

with Glu219 of the enzyme. The sulfodiimine bound to matrilysin is a monodentate zinc ligand and the<br />

weakest of the zinc-binding groups. The comparison of several inhibitors with both carboxylate and<br />

hydroxamate zinc-binding groups demonstrates this property in fibroblast and neutrophil collagenases as<br />

well. While the potency of inhibitors with different zinc-binding groups maps directly to the number of<br />

bonds formed <strong>by</strong> the zinc-binding group, some of the increase in potency of the hydroxamate group over<br />

charged groups most likely is a result of the decreased energetic cost of the desolvation of the neutral<br />

hydroxamate.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_182.html [4/5/2004 5:03:58 PM]


Document<br />

VIII. S1' Interactions<br />

Matrix metalloproteinase structural studies of the P'-side inhibitors to date show a common set of<br />

inhibitor-enzyme interactions. This can be attributed primarily to the strong directional zinc-binding<br />

forces. Further stabilizing forces from the backbone hydrogen-bonding patterns common to a β sheet<br />

allow for minor adjustments due to the zinc interactions to be made while maintaining a common<br />

pharmacophore.<br />

Page 183<br />

The fairly rigid constraints of binding <strong>based</strong> on the known hydroxamate inhibitors allows the use of<br />

computer-aided modeling to play a useful role in the design of specific MMP inhibitors. Exploration of<br />

the S1' pocket was carried out <strong>by</strong> docking the P1' group within the cavity followed <strong>by</strong> rounds of energy<br />

minimization. In order to maintain the integrity of the MMP structure several limitations were used. All<br />

MMP atoms that are greater than 8 Å from the docked inhibitor were frozen. The Cα atoms of all<br />

residues within 8 Å of the inhibitor were initially constrained to their original position <strong>by</strong> a 20 kcal/mol<br />

Å 2 force constant that was gradually relaxed to 1 kcal/mol Å 2. Strong initial constraints were also placed<br />

on the conserved hydrogen bonds and zinc-ligand interactions.<br />

This method has several advantages and disadvantages over the common static treatment of target<br />

structures. The advantages are that it more closely approximates the actual dynamic state of protein<br />

structure and is not computationally prohibitive. The disadvantages include the increased computational<br />

cost over a static enzyme target and the fact that gross structural rearrangements can still not be<br />

accounted for.<br />

The structural similarity of the active site of the MMP family allows structure-<strong>based</strong> drug design to<br />

effectively be used for those enyzmes whose structures have not been determined yet. Examination of<br />

the S1' cavities of HFC and HNC clearly indicates a path for designing inhibitors that bind preferentially<br />

(Figure 5). The cavity of HFC is mostly filled <strong>by</strong> the leucine side chain of the preferred substrate, while<br />

the S1' pocket of HNC [26] and HFS [30] remains unfilled. The sequence similarities of the gelatinases<br />

with HNC indicate that they too can accommodate a much larger P1' group.<br />

A series of compounds was designed to explore filling the long S1' tunnel of gelatinase B [41, 42]. The<br />

optimum length for binding to gelatinase B was found. There was an increase in affinity for the phenolic<br />

ethers versus the benzylic ethers of the same overall length for binding to gelatinase B, but there was no<br />

preference seen in binding to HFS. Surprisingly the phenolic ethers also showed potent binding to HFC.<br />

The size of the bottom of the S1' pocket and the differences in the preferred torsion angle for the bond to<br />

the aromatic ring both play a role in this differential binding.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_183.html [4/5/2004 5:03:59 PM]


Document<br />

IX. S2' Interactions<br />

Page 184<br />

The interactions at the S2' site display an interesting structure-activity relationship. While the<br />

interactions do not include any hydrogen bonds, favorable stacking interactions do play a significant role<br />

in binding. A glycine residue at P2' results in a loss of three orders of magnitude in potency for<br />

otherwise identical inhibitors [41]. There is a preference for an aromatic ring at this position in natural<br />

peptide substrates [43]. Structural considerations also allow the placement of a t-butyl group here. The<br />

potency of a t-butyl glycine P2' is less than that of a phenylalanine (Table 1), but the expected gain in<br />

bioavailability brought about <strong>by</strong> shielding the amide bond from solvation effects should compensate for<br />

the loss in potency.<br />

X. Conclusions<br />

High resolution x-ray crystallographic structure determination is an essential step in structure-<strong>based</strong> drug<br />

design. The need for high resolution structural data<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_184.html [4/5/2004 5:04:31 PM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_185.html (1 of 2) [4/5/2004 5:04:42 PM]<br />

Page 185


Document<br />

Figure 6<br />

(a) The pocket of HFC with a leucine residue in the S1' pocket.<br />

(b) The volume of the S1' pocket of HFC can change when there are<br />

favorable interactions. The binding of the (CH 2) 4OPh can cause<br />

Arg214 to move, there<strong>by</strong> making room for the extended side chain.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_185.html (2 of 2) [4/5/2004 5:04:42 PM]


Document<br />

to develop an appropriate SAR is demonstrated in the case of the inhibitors shown in Table 1. The<br />

unusual potency of a benzylic ether for HFC was unexpected and would not have been predicted with<br />

standard docking and minimization studies (Figure 6).<br />

Page 186<br />

The differences in the potency of the various 4-substituted analogs of inhibitor 10 against HFC suggest ππ<br />

interactions are the driving force for the displacement of arginine 214. The electron-withdrawing Cl<br />

substitution decreases the affinity for HFC while increasing the affinity for HFS. The leading 4-pentyl<br />

group can not effectively interact with arginine 214 in HFC; therefore, the rearrangement does not<br />

occur. The open channel that is present in HFS and gelatinase B presents no such impediment to binding<br />

and the affinity is essentially equal to the unsubstituted form.<br />

Not all structure-<strong>based</strong> design experiments are successful. Attempts to displace the arginine residue that<br />

caps the S1' pocket of HFC <strong>by</strong> forming a salt link with a P1' carboxylate or hydroxyl moiety were<br />

unsuccessful [42]. However, these failed attempts offer some redeeming features in the refinement of<br />

parameters that can be used to evaluate the energetic potentials for displacing buried water molecules as<br />

well as the inherent desolvation energies for polar compounds.<br />

The outlook for structure-<strong>based</strong> drug design is good. The advancement in both x-ray area detectors and<br />

computer hardware will make the determination of a series of compounds bound to a target enzyme for<br />

use in SAR development commonplace in drug-discovery efforts. The continued explosion of structural<br />

studies will lead to an increased understanding of the dynamics of protein interactions, which will, in<br />

turn, lead to better docking algorithms. The combination of structural information and greater<br />

computational power will also make more accurate predictions of protein—ligand interactions possible.<br />

References<br />

1. Birkedal-Hansen H, Moore WGI, Bodden MK, Windsor LJ, Birkedal-Hansen B, DeCarlo A, Engler<br />

JA. Matrix metalloproteinases: a review. Crit. Rev. Oral. Biol. Med. 1993; 4:197–250.<br />

2. Beckett RP, Davidson AH, Drummond AH, Huxley P, Whittaker M. Recent advances in matrix<br />

metalloproteinase inhibitor research. DDT 1996; 1:16–26.<br />

3. Birkdell-Hansen H. Proteolytic remodeling of extracellular matrix. Cur. Op. Cell Biology 1995;<br />

7:728–735.<br />

4. Stocker W, Grams F, Baumann U, Gomis-Ruth F-X, McKay DB, Bode W. The metzincins<br />

topological and sequential relations between the astacins, adamalysins, serralysins, and matrixins<br />

(collagenases) define a superfamily of zinc peptidases. Protein Sci. 1995; 4:823–840.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_186.html [4/5/2004 5:04:45 PM]


Document<br />

5. Willenbrock F, Murphy G, Phillips IR, Brocklehurst K. The second zinc atom in the matrix<br />

metalloproteinase catalytic domain is absent in the full-length enzymes: a possible for the C-terminal<br />

domain. FEBS Lett. 1995; 358:189–192.<br />

Page 187<br />

6. Lowry CL, McGeehan G, Levine H. Metal ion stabilization of the conformation of a recombinant 19kDa<br />

catalytic fragment of human fibroblast collagenase. Proteins 1992; 12:42–48.<br />

7. Hodgson J. Remodeling MMPIs. Bio Technology 1995; 13:554–557.<br />

8. Brown PD. Matrix metalloproteinase inhibitors: a novel class of anticancer agents. Adv. Enzyme<br />

Regul. 1995; 35:293–301.<br />

9. Watson SA, Morris TM, Robinson G, Crimmin MJ, Brown PD, Hardcastle JD. Inhibition of organ<br />

invasion <strong>by</strong> matrix metalloproteinase inhibitor batimastat (BB-94) in two human colon carcinoma<br />

metastasis models. Cancer Res. 1995; 16:3629–3633.<br />

10. Sledge GW Jr, Qulali M, Goulet R. Bone EA, Fife R. Effect of matrix metalloproteinase inhibitor<br />

batimastat on breast cancer regrowth and metastasis in athymic mice. J. Natl. Cancer Inst. 1995;<br />

87:1546–1150.<br />

11. Mitchel PG, Magna HA, Reeves LM, Lopresti-Morrow LL, Yocum SA, Rosner PJ, Geoghegam KF,<br />

Hambor JE. Cloning, expression, and type II collagenolytic activity of matrix metalloproteinase-13 from<br />

human osteoarthritic cartilage. J. Clin. Invest. 1996; 3:761–768.<br />

12. Freije JMP, Diez-Ita I, Balbin M, Sanchez LM, Blaco R, Toliva J, Lopez-Otin C. Molecular cloning<br />

and expression of collagenase-3, a novel human matrix metalloproteinase produced <strong>by</strong> breast<br />

carcinomas. J. Biol. Chem. 1994; 269:16766–16773.<br />

13. Rosenberg GA. Matrix metalloproteinases in brain injury. J. Neurotrauma 1995; 12:833–842.<br />

14. Nomura H, Fujimoto N, Seiki M, Mai M, Okada Y. Enhanced production of matrix<br />

metalloporteinase and activation of matrix metalloproteinase 2 (gelatinase A) in human gastric<br />

carcinomas. Int. J. Cancer 1996; 69:9–16.<br />

15. Itoh F, Yamamoto H, Hinoda Y, Imai K. Enhanced secretion and activation of matrilysin during<br />

malignant conversion of human colorectal epithelium and its relationship with invasive potential of<br />

colon cancer cells. Cancer 1996; 77:1717–1721.<br />

16. Deb S, Gottschall PE. Increased production of matrix metalloproteinases in enriched astrocyte and<br />

mixed hippocampal cultures treated with beta-amyloid peptides. J. Neurochem. 1996; 66:1641–1647.<br />

17. Crabbe T, Zucker S, Cockett MI, Willenbrock F, Tickle S, O'Connell JP, Scothern JM, Murphy G,<br />

Docherty AJP. Mutation of the active site glutamic acid of human gelatinase A: effects on latency,<br />

catalysis and the binding of tissue inhibitor of metalloproteinase-1. Biochemistry 1994; 33:6684–6690.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_187.html (1 of 2) [4/5/2004 5:04:47 PM]


Document<br />

18. Murphy G, Allan JA, Willenbrock, F, Cockett MI, Docherty AJP. The role of the C-terminal domain<br />

in collagenase and stromelysin specificity. J. Biol. Chem. 1992; 267:9612–9618.<br />

19. Murphy G, Willenbrock F, Ward RV, Cockett MI, Eaton D, Docherty AJP. The C-terminal domain<br />

of 72 kDa gelatinase A is not required for catalysis, but is essential for membrane activation and<br />

modulates interactions with tissue inhibitors of metalloproteinase. J. Biochem. 1992; 328:637–641.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_187.html (2 of 2) [4/5/2004 5:04:47 PM]


Document<br />

20. Murphy G, Docherty AJP. Assessment of the role of fibronectin-like domain of gelatinase A <strong>by</strong><br />

analysis of a deletion mutant. J. Biol. Chem. 1994; 269:6632–6636.<br />

Page 188<br />

21. Shipley JM, Doyle GA, Fliszar CJ, Ye QZ, Johnson LL, Shapiro SD, Welgus HG, Senior RM. The<br />

structural basis for the elastolytic activity of the 92-kDa and 72-kDa gelatinases. Role of the fibronectin<br />

type II-like repeats. J. Biol. Chem. 1996; 271:4335–4341.<br />

22. Spurlino, J, Smallwood AM, Carlton DC, Banks TM, Vavra KJ, Johnson JS, Cook EW, Falvo J,<br />

Wahl RC, Pulvino TA, Wendoloski JJ, Smith, DL. 1.56 Å structure of mature truncated human<br />

fibroblast collagenase. Proteins 1994; 19:98–109.<br />

23. Lovejoy B, Cleas<strong>by</strong> A, Hassell AM, Longley K, Luther MA, Weigl D, McGeehan G, McElroy AB,<br />

Drewry D, Lambert MH, Jordan SR. <strong>Structure</strong> of the catalytic domain of fibroblast collagenase<br />

complexed with an inhibitor. Science 1994; 263:375–377.<br />

24. Lovejoy B, Hassell AM, Luther MA, Weigl D, Jordan SR. Crystal structures of recombinant 19-kDa<br />

human fibroblast collagenase complexed to itself. Biochemistry 1994; 33:8207–8217.<br />

25. Borkakoti N, Winkler FK, Williams DH, D'Arcy A, Broadhurst MJ, Brown PA, Johnson WH,<br />

Murray EJ. <strong>Structure</strong> of the catalytic domain of human fibroblast collagenase complexed with an<br />

inhibitor. Struct. Biol. 1994; 1:106–110.<br />

26. Stams T, Spurlino JC, Smith DL, Wahl RC, Ho TF, Qoronfleh MW, Banks TM, Rubin B. <strong>Structure</strong><br />

of human neutrophil collagenase reveals large S1' specificity pocket. Struct. Biol. 1994; 1:119–123.<br />

27. Bode W, Reinemer P, Huber R, Kleine T, Schnierer S, Tschesche H. The X-ray crystal structure of<br />

the catalytic domain of human neutrophil collagenase inhibited <strong>by</strong> a substrate analogue reveals the<br />

essentials for catalysis and specificity. EMBO J. 1994; 13:1263–1269.<br />

28. Browner MF, Smith WW, Castelhano AL. Matrilysin-inhibitor complexes: common themes among<br />

metalloproteinases. Biochemistry 1995; 34:6602–6610.<br />

29. Wetmore DR, Hardman KD. Roles of the propeptide and metal ions in the folding and stability of<br />

the catalytic domain of stromelysin (matrix metalloproteinase 3). Biochemistry 1996; 35:6549–6558.<br />

30. Dhanaraj V, Ye Q–Z, Johnson LL, Hupe DJ, Ortwine DF, Dunbar JB, Rubin JR, Pavvlovsky A,<br />

Humblet C and Blundell TL. X-ray structure of a hydroxamate inhibitor complex of stromelysin<br />

catalytic domain and its comparison with members of the zinc metalloproteinase superfamily. <strong>Structure</strong><br />

1996; 4:375–386.<br />

31. Bernstein FC, Koetzle TF, Williams GJB, Meyer EF, Brice MD, Rodgers JR, Kennard O,<br />

Shimanouchi T, Tasumi M. The Protein Data Bank: a computer-<strong>based</strong> archival file for macromolecular<br />

structures. J. Mol. Biol. 1977; 112:535–542.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_188.html (1 of 2) [4/5/2004 5:04:49 PM]


Document<br />

32. Carson M. Ribbon models of macromolecules. J. Mol. Graphics 1987; 5:103–106.<br />

33. Connolly ML. The molecular surface package J. Mol. Graphics 1993; 11:139–141.<br />

34. Li J, O'Hare MC, Skarzynski T, Lloyd LF, Curry VA, Clark IM, Bigg HF, Hazleman BL, Cawston<br />

TE, Blow DM. X-ray structure of a hydroxamate inhibitor complex of stromelysin catalytic domain and<br />

its comparison with members of the zinc metalloproteinase superfamily. <strong>Structure</strong> 1996; 4:375–386.<br />

35. Grams F, Crimmin M, Hinnes L, Huxley P, Pieper M, Tschesche H, Bode W. <strong>Structure</strong><br />

determination and analysis of human neutrophil collagenase complexed with a hydroxamate inhibitor.<br />

Biochemistry 1995; 34:14012–14020.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_188.html (2 of 2) [4/5/2004 5:04:49 PM]


Document<br />

Page 189<br />

36. Bode W, Reinemer P, Huber R, Kleine T, Schnierer S, Tschesche H. The X-ray crystal structure of<br />

the catalytic domain of human neutrophil collagenase inhibited <strong>by</strong> a substrate analogue reveals the<br />

essentials for catalysis and specificity. EMBO J. 1994; 13:1263–1269.<br />

37. Schwartz MA, Van Wart HE. In: Ellis GP, Luscombe DK, eds. Progress in Medicinal Chemistry,<br />

Vol. 29. London: Elsevier Publishers, 1992: Chapter 8.<br />

38. Johnson WH, Roberts NA, Borkakoti N. J. Enzyme Inhibition 1987; 2:1–22.<br />

39. Wahl RC, Dunlop RP, Morgan BA. In: Bristol JA, ed. Annual Reports in Medicinal Chemistry. New<br />

York: Academic Press, 1990: Chapter 19.<br />

40. Henderson B, Docherty AJP, Beeley NRA. <strong>Drug</strong>s of the Future 1990; 15:495–408.<br />

41. Wahl RC, Pulvino TA, Mathiowetz AM, Ghose AK, Johnson JS, Delecki D, Cook ER, Gainer JA,<br />

Gowravaram MR, Tomczuk BE. Hydroxamate inhibitors of human gelatinase B (92kDa). Biorg. and<br />

Med. Chem. Lett. 1995; 5:349–352.<br />

42. Gowravaram MR, Tomzcuk BE, Johnson JS, Delecki D, Cook ER, Ghose AK, Mathiowetz AM,<br />

Spurlino JC, Rubin B, Smith DL, Pulvino T, Wahl RC. Inhibition of matrix metalloproteinases <strong>by</strong><br />

hydroxamates containing heteroatom-<strong>based</strong> modifications of the P1' group. J. Med. Chem. 1995;<br />

38:2570–2581.<br />

43. Netzel-Arnett S, Fields G, Birkdal-Hansen H, Avan Wart HE. Sequence specificities of human<br />

fibroblast and neutrophil collagenases. J. Biol. Chem. 1991; 206:6747–7855.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_189.html [4/5/2004 5:04:50 PM]


http://legacy.netlibrary.com/reader/message.asp?message=811&BookID=12640&FileName=Page_190.html<br />

The requested page could not be found.<br />

Return to previous page<br />

http://legacy.netlibrary.com/reader/message.asp?message=811&BookID=12640&FileName=Page_190.html [4/5/2004 5:04:52 PM]


Document<br />

7<br />

<strong>Structure</strong>—Function Relationships in Hydroxysteroid Dehydrogenases<br />

Igor Tsigelny and Michael E. Baker<br />

University of California, San Diego, La Jolla, California<br />

I. Introduction<br />

Page 191<br />

Steroid hormones regulate a multitude of physiological processes in humans. Androgens and estrogens<br />

regulate sexual development and reproduction; glucocorticoids are important in the response to stress;<br />

vitamin D is important in bone growth; progestins are important for a viable fetus during pregnancy;<br />

mineralocorticoids regulate sodium and potassium balance to maintain normal blood pressure.<br />

Moreover, the growth of some breast and prostate tumors depends on steroids. With this multitude of<br />

medically important steroid-dependent actions, much research has gone into understanding their mode<br />

of action, with most of the effort concerned with the receptors that mediate the actions of steriods.<br />

A. High Blood Pressure and 11β-Hydroxysteroid Dehydrogenase<br />

It is only in the last decade that the role of hydroxysteroid dehydrogenases (Figure 1) in regulating the<br />

actions of steroids has been appreciated [1–6]. This mechanism for regulating steroid hormone action<br />

was uncovered in several laboratories studying various aspects of high blood pressure. One source was<br />

the study in the 1970s that identified the syndrome, Apparent Mineralocorticoid Excess (AME), a<br />

genetic disease that results in high blood pressure in children [7–10]. Also important is the work from<br />

laboratories investigating paradoxes in the mechanism of action of aldosterone in the kidney [1–4,9,10].<br />

These studies identified 11β-hydroxysteroid dehydrogenase (11β-HSD) as a key enzyme.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_191.html [4/5/2004 5:04:54 PM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_192.html (1 of 2) [4/5/2004 5:04:59 PM]<br />

Page 192


Document<br />

Figure 1<br />

Reactions catalyzed <strong>by</strong> 11β-hydroxysteroid and 17β-hydroxysteroid<br />

dehydrogenases. (a) 11 β-hydroxysteroid dehydrogenase type 1, an NADPH-dependent<br />

enzyme, catalyzes the conversion of the inactive steroid, cortisone, to cortisol, which is<br />

the biologically active glucocorticoid. 11β-hydroxysteroid dehydrogenase type 2, an<br />

NAD + -dependent enzyme, catalyzes the reverse direction. (b) 17β-hydroxysteroid<br />

dehy-drogenase type 1, an NADPH-dependent enzyme, catalyzes the reduction of<br />

estrone to estradiol. Type 2, an NAD + -dependent enzyme, catalyzes the oxidation of<br />

estradiol to estrone. Type 3, an NADPH-dependent enzyme, catalyzes the reduction of<br />

androstene dione to testosterone. Type 4, an NAD + -dependent enzyme, catalyzes the<br />

oxidation of estradiol to estrone, and androstenediol to dehydroepiandrosterone.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_192.html (2 of 2) [4/5/2004 5:04:59 PM]


Document<br />

Page 193<br />

The enzyme 11β-HSD interconverts the active glucocorticoid cortisol and cortisone, an inactive<br />

metabolite (Figure 1a). By the oxidation of cortisol to cortisone, 11β-HSD prevents glucocorticoids<br />

from deleterious actions in certain cell types. For example, excess glucocorticoids in Leydig cells inhibit<br />

testosterone synthesis [3,11]. Expression of 11β-HSD in Leydig cells prevents this effect of<br />

glucocorticoids. In this way, 11β-HSD is important in androgen action. The enzyme 11β-HSD is also<br />

important in aldosterone action in the distal tubule of the kidney. Glucocorticoids have high affinity for<br />

the mineralocorticoid receptor [12] and can stimulate the mineralocorticoid response—uptake of sodium<br />

from urine—one effect of which is to increase blood pressure. Local expression of 11β-HSD in the<br />

distal tubule prevents this effect of glucocorticoids. The steroid aldosterone, which is not metabolized <strong>by</strong><br />

11β-HSD, can bind to the mineralocorticoid receptor and regulate sodium and potassium balance. Thus,<br />

11β-HSD has an important role in regulating the biological actions of both glucocorticoids and<br />

mineralocorticoids.<br />

As would be expected, interference with 11β-HSD activity due to a mutation [13,14] or <strong>by</strong> an inhibitor<br />

such as licorice (Figure 2) [1–3,15] has a variety of physiological effects including high blood pressure<br />

due to mineralocorticoid actions of glucocorticoids in the kidney's distal tubule. Thus, studies to unravel<br />

genetic hypertension in children and the actions of aldosterone in the kidney yielded the general insight<br />

that, at specific times, altered expression of 11β-HSD in specific tissues is an important mechanism for<br />

regulating glucocorticoid, mineralocorticoid, and androgen action.<br />

A similar mechanism has been found for 17β-hydroxysteroid dehydrogenase (17β-HSD), the enzyme<br />

that regulates the concentrations of estradiol and testosterone in human [5,16,17] (Figure 1b). Genetics<br />

diseases associated with mutations in this enzyme lead to developmental abnormalities [18]. Enzymes<br />

that regulate the concentrations of retinoids [19] and prostaglandins [20] may also have a similar role<br />

[6].<br />

B. Multiple Divergent 11β-Hydroxysteroid and 17β-Hydroxysteroid Dehydrogenases<br />

The cloning and sequencing of 11β-HSD [21–25] and 17β-HSD [16–18,26] revealed two 11β-HSDs and<br />

four 17β-HSDs with very different sequences<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_193.html [4/5/2004 5:05:01 PM]


Document<br />

Figure 2<br />

<strong>Structure</strong> of licorice and carbenoxolone. Glycyrrhizic acid, a constituent of<br />

licorice extract, is found in the root of Glycyrrhiza glabra. The glycosidic group at C3 is<br />

cleaved <strong>by</strong> bacteria in the small intestine to form glycyrrhetinic acid, the compound that<br />

inhibits 11β-hydroxysteroid dehydrogenase. Carbenoxolone, a water soluble synthetic<br />

analog of glycyrrhetinic acid, is widely used to regulate 11β-HSD in vitro and in vivo.<br />

Page 194<br />

(Figure 3, Table 1). This was surprising, as one would expect the two 11β-HSDs to be similar because<br />

they recognize the same substrates. Instead, the two 11β-HSDs have less than 20% sequence identity,<br />

after including gaps in the alignment (Table 1). The same degree of sequence divergence is found in the<br />

four 17β-HSDs [6]. This sequence divergence is reflected in differences in their catalytic properties. For<br />

example 11β-hydroxysteroid dehydrogenase-type 1 (11β-HSD-1) is an NADPH-dependent enzyme that<br />

converts cortisone to cortisol, and 11β-hydroxysteroid dehydrogenase-type 2 is an NAD +-dependent<br />

enzyme that oxidizes cortisol to cortisone. The enzyme 17βHSD-1 is an NADPH-dependent enzyme<br />

that converts estrone to estradiol, and 17βHSD-2 is an NAD +-dependent enzyme that oxidizes estradiol<br />

to estrone and testosterone to androstenedione.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_194.html (1 of 2) [4/5/2004 5:05:06 PM]


Document<br />

Figure 3<br />

Alignment of 11β-and 17β-hydroxysteroid dehydrogenases. As seen in this<br />

Figure and Table 1, the sequences of the two 11β-HSDs and four 17β-HSDs are very<br />

divergent. Boxes denote sites where either 5 or 6 residues are conserved, which are<br />

likely to be functionally important.<br />

Page 195<br />

There is considerable interest in understanding the structural bases for these differences because this<br />

information would be very useful in designing steroids and other compounds to selectively regulate the<br />

activity of one or more steroid dehydrogenases as a means of treating hormone-responsive diseases.<br />

There is precedent for this kind of treatment in the use of licorice extract from the root of the plant<br />

Glycyrrhiza glabra [15,27] to treat Addison's disease,<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_195.html (1 of 2) [4/5/2004 5:05:15 PM]


Document<br />

Figure 3<br />

(Continued)<br />

Page 196<br />

which is characterized <strong>by</strong> insufficient levels of cortisol. Licorice inhibits 11β-HSD-2, which raises the<br />

circulating levels of cortisol and provides some relief from the symptoms of Addison's disease. This use<br />

of licorice is an example of a plant-derived compound having important uses in mammalian steroid<br />

hormone physiology and indicates another reason why elucidation of the structure of steroid<br />

dehydrogenases is of medical interest. Plants contain a wide variety of compounds, many of which have<br />

been purified and had their structures determined; however, we don't know which of these compounds<br />

inhibit steroid<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_196.html (1 of 2) [4/5/2004 5:05:23 PM]


Document<br />

Table 1 Percent Identity Between Hydroxysteroid Dehydrogenase Sequences Shown in Figure 3<br />

Page 197<br />

11β-HSD-2 17β-HSD-2 17β-HSD-3 17β-HSD-1 11β-HSD-1 17β-HSD-4<br />

11β-HSD-2 0.00 46.1 20.3 28.6 20.6 18.1<br />

17β-HSD-2 0.0 21.1 21.2 17.7 18.9<br />

17β-HSD-3 0.0 19.1 19.6 17.5<br />

17β-HSD-1 0.0 21.1 20.4<br />

11β-HSD-1 0.0 17.1<br />

17β-HSD-4 0.0<br />

dehydrogenases. Knowledge of structure-activity relationships for the binding site on steroid<br />

dehydrogenases will be helpful in identifying novel compounds from plants and other sources that could<br />

be useful in regulating steroid dehydrogenases.<br />

At this time, we are just beginning to work on this ambitious goal. Structural information is limited. The<br />

3-D structure of 17β-HSD type 1 has been determined [28], but without the steroid or cofactor in the<br />

binding site. Fortunately, 11β-HSD and 17β-HSD belong to a large family of enzymes that are called<br />

short-chain alcohol dehydrogenases [29–31] or sec-alcohol dehydrogenases [32]. The structures of<br />

dehydrogenase homologs in bacteria, plants, and animals have been determined [33–37] and we used<br />

them as templates for modeling 11β-HSD and 17β-HSD [38,39]. There also is information about the<br />

effects of mutations on catalytic activity in 11β-HSD-1 [40] and 17β-HSD-1 [41] and in homologs,<br />

especially for Drosophila alcohol dehydrogenase (ADH) [42–48]. Together they enable us to begin to<br />

understand the relationship between structure and function in hydroxysteroid dehydrogenases, as we<br />

discuss in this chapter.<br />

II. Methods<br />

A. Molecular Modeling<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_197.html (1 of 2) [4/5/2004 5:05:25 PM]


Document<br />

Important for the validity of the models that we constructed is the evidence from models of other<br />

proteins indicating that two proteins can have as little as 20 to 25% sequence identity and still have very<br />

similar 3D structures, especially in α helices and β stands [49–52]. Variation is found in the loops and<br />

coiled structures. A relevant example for this chapter is the comparison of the tertiary structure of rat<br />

dihydropteridine reductase [33] and Streptomyces hydrogenans 20β-hydroxysteroid dehydrogenase [34].<br />

As noted <strong>by</strong> Varughese et al. [33], despite less than 20% sequence identity between dihydropteridine<br />

reductase and<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_197.html (2 of 2) [4/5/2004 5:05:25 PM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_198.html (1 of 2) [4/5/2004 5:05:32 PM]<br />

Page 198


Document<br />

Figure 4<br />

Amino acids important in cofactor and catalysis in human 11b-hydroxysteroid<br />

dehydrogenase types 1 and 2. (a) 11b-HSD type 1. Preference of 11b-HSD type 1<br />

for NADPH resides in lysine-44 and arginine-66, which have positively charged side<br />

chains that stabilize the binding of the 2'-phosphate on NADPH. These residues also<br />

counteract the repulsive interaction between glutamic acid 69 and the phosphate group.<br />

(b) 11b-HSD type 2. Preference of 11b-HSD type 2 for NAD+ is due to favorable bonds<br />

with aspartic acid-91, serine-92, and threonine-112. Moreover there is a coulombic<br />

repulsion between aspartic acid-91 and NADP+, which destabilizes binding of NADP+<br />

11b-HSD type 2 lacks a near<strong>by</strong> amino acid with a positively charged side chain that<br />

could diminish the repulsive interaction between NADP+ and aspartic acid-91. Also<br />

shown are threonine residues that could hydrogen bond<br />

to nicotinamide's carboxamidemoiety.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_198.html (2 of 2) [4/5/2004 5:05:32 PM]


Document<br />

Page 199<br />

20β-hydroxysteroid dehydrogenase, the root mean square deviation for the two tertiary structures is 2 Å<br />

over 160 C α carbon atoms.<br />

We aligned human 11β-HSD-1 [21] and 11β-HSD-2 [22] with S. hydrogenans 20β-hydroxysteroid<br />

dehydrogenase and Escherichia coli 7α-hydroxysteroid dehydrogenase [37] for 3D modeling. Human<br />

11β-HSD has extra segments at the amino terminus and carboxyl terminus. Previously reported<br />

alignments [30,31,53] were used to find the core structure consisting of about 225 residues that are<br />

structurally similar to the template. The first 190 residues of the 255 residues are reasonably well<br />

conserved among the hydroxysteroid dehydrogenases. Alignment of the C-terminal 65 residues is less<br />

certain as this part contains gaps and insertions. Fortunately, the core 190 residues contains the catalytic<br />

site and the cofactor binding site. We also superimposed the two 11β-HSD structures on mouse carbonyl<br />

reductase [37]. The 11β-HSD 3D structures superimpose nicely on α helices E and F and other helices<br />

and strands that are important in binding of cofactor and substrate. Then, we extracted NADPH from<br />

carbonyl reductase and NAD + from 7α-hydroxysteroid dehydrogenase and inserted the cofactors into<br />

the structures of the two 11β-HSDs.<br />

The α helix F in 17β-HSD-1 [16], 17β-HSD-2 [17], 17β-HSD-3 [18], and porcine 17β-HSD-4 [26] was<br />

constructed <strong>by</strong> modeling on α helix F in 20β-hydroxysteroid dehydrogenase. Comparisons with other<br />

3D structures [33–37] have demonstrated that this α helix is highly conserved. The modeled dimers<br />

were not minimized as a dimer complex to avoid the artifactual adjustment of α helix F side chains.<br />

The Homology program (Biosym Technologies, Inc., 1995) was used to model a 255-residue segment of<br />

11β-HSD and α helix F of 17β-HSD on the S. hydrogenans 20β-hydroxysteroid dehydrogenase<br />

template. To produce the final model (Figures 4 and 5) this program finds an optimal configuration of<br />

the residues when arranged in the template structure <strong>by</strong> minimizing unfavorable interactions between<br />

amino acid side chains. The side chains of each monomer were then minimized (1,000 iterations of the<br />

conjugate gradient) using the Discover program (Biosym Technologies Inc., 1995).<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_199.html [4/5/2004 5:05:34 PM]


Document<br />

III. Results and Discussion<br />

Figure 5<br />

<strong>Structure</strong> of α helix F interface of human 11β-hydroxysteroid dehydrogenase types 1<br />

and 2. The α helix F part of the dimer interface on 11β-HSD-1 and -2 is shown along<br />

with side chains of the highly conserved tyrosine and lysine residues and other residues<br />

that are oriented into the cavity that binds substrate and nucleotide cofactor.<br />

A. NADPH Binding Site on 11β-Hydroxysteroid Dehydrogenase Types 1 and 2<br />

Page 200<br />

Several lines of evidence—sequence analysis, mutagenesis studies, and the solved 3D structure of<br />

homologs of 11β-HSD—indicate that the nucleotide binding site in these enzymes has many similarities<br />

to that in other classes of dehydrogenases. For many dehydrogenases, the nucleotide binding domain<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_200.html (1 of 2) [4/5/2004 5:05:58 PM]


Document<br />

Page 201<br />

consists of a β strand α helix, β strand in a fold that provides a hydrophobic pocket for the adenosine<br />

monophosphate (AMP) part of the nucleotide cofactor [51,54,55]. In short-chain alcohol<br />

dehydrogenases, this βαβ fold is at the amino terminus. The turn between the first β strand and the α<br />

helix is a glycine-rich segment of the form Gly-X-X-X-Gly-X-Gly. This glycine-rich segment forms a<br />

hydrophobic pocket that allows close association of the AMP part of the cofactor.<br />

However, this glycine-rich segment has other functions in short-chain alcohol dehydrogenases. Tanaka<br />

et al. [37] and our 3D modeling [60] indicate that this glycine rich segment has an important role in<br />

cofactor specificity and binding of the nicotinamide moiety to 11β-HSD.<br />

B. 11β-HSD-1 Preference for NADPH<br />

Figure 4 shows our 3D model of human 11β-HSD types 1 and 2. These models identify residues<br />

important in preference of 11β-HSD-1 for NADPH and 11β-HSD-2 for NADH. In 11β-HSD type 1,<br />

lysine-44 and arginine-66 have favorable coulombic interactions with the 2'-phosphate on NADP + that<br />

stabilize binding (Figure 4a). Moreover, their positively charged side chains compensate for the negative<br />

interaction between glutamic acid-69 and the 2'-phosphate group. Tanaka et al. [37] found a similar<br />

function for lysine-14 and arginine-39 in the preference of mouse carbonyl reductase for NADPH.<br />

C. 11β-HSD-2 Preference for NAD+<br />

The 3D structure of 11β-HSD-2 shows that NAD + has stabilizing interactions between the ribose<br />

hydroxyl and aspartic acid-91, serine-92, and threonine-112. Replacement of NAD + with NADP +<br />

reveals a coulombic repulsion between the 2'-phosphate group and aspartic acid-91. However, 11β-HSD<br />

type 2 lacks a near<strong>by</strong> amino acid with a positively charged side chain that could compensate for the<br />

negative charge on aspartic acid-91. This explains the preference of 11β-HSD-2 for NAD +.<br />

D. Amino Acids Important in Binding the Nicotinamide Ring and Carboxamide Moiety<br />

Both 11β-HSD types 1 and 2 contain residues in the C-terminal half that interact with the nicotinamide<br />

ring and carboxamide moiety to limit rotations about the N-glycosidic bond. These intersections are<br />

important in positioning the cofactor for proS hydride transfer at C4.<br />

In 11β-HSD-1, cysteine-213 stabilizes the nicotinamide ring; threonine-220 and threonine-222 stabilize<br />

the carboxamide moiety. In 11β-HSD-2, there are more interactions: proline-262, phenylalanine-265,<br />

threonine-267, serine-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_201.html [4/5/2004 5:06:00 PM]


Document<br />

Page 202<br />

269, and valine-270 are close to either the nicotinamide ring or the carboxamide moiety. In addition, the<br />

face of the side chain of phenylalanine-94 is below the nicotinamide ring and its carboxamide group.<br />

This latter interaction is unusual because phenylalanine-94 is between the two canonical glycine<br />

residues in the βαβ fold, which is usually thought of as interacting mainly with the AMP part of the<br />

cofactor. In 11β-HSD-2, there is an interesting configuration of amino acids with aromatic side chains<br />

that are below the nicotinamide ring and which provide a hydrophobic cushion for NAD +.<br />

E. Catalytic Site<br />

Comparison of 11β-HSD-1 with homologs identifies tyrosine-183 and lysine-187 as being highly<br />

conserved residues. Mutagenesis of these residues [40] and the homologous tyrosine and lysine in 17β-<br />

HSD-1 and Drosophila alcohol dehydrogenase [44,45] shows that these residues are important for<br />

catalytic function. The 3D model of 11β-HSD-1 presented in Figure 4 shows that tyrosine-183 is 3.6 Å<br />

from the nicotinamide C4, where hydride transfer occurs. Similarly, in 11β-HSD-2, tyrosine-232 is 4 Å<br />

from C4 on NAD +. Their positions support the notion that tyrosine is the catalytically active residue.<br />

However, a problem with this model is that the pKa of tyrosine is about 10, which would make this<br />

residue a poor nucleophile at neutral pH. To resolve this problem for the homologous tyrosine in<br />

Drosophila alcohol dehydrogenase, Chen et al. [44] proposed that the pKa of tyrosine is lowered <strong>by</strong> a<br />

near<strong>by</strong> positively charged lysine. The 3D structure of the two 11β-HSDs shows that lysine-187 and<br />

lysine-236 are close to the proposed catalytically active tyrosine residues, which supports the hypothesis<br />

of Chen et al. [44].<br />

F. Dimer Interface and the Catalytic Site<br />

Most short-chain alcohol dehydrogenases are active as either dimers or tetramers. Analysis of rat<br />

dihydropteridine reductase <strong>by</strong> Varughese et al. [33] indicates that the dimer interface consists of α helix<br />

E and α helix F from each monomer arranged in a four α helix bundle, a structure in which the<br />

hydrophobic surfaces on the helices form a core that yields very stable structure in a wide variety of<br />

proteins [56–59]. A four-helix bundle also appears to stabilize S. hydrogenans 20β-hydroxysteroid<br />

dehydrogenase, a tetrameric enzyme [34]. The α helix F contains the conserved tyrosine and lysine<br />

residue, which adds a constraint to changes in the sequence of this helix. It has at least two functions:<br />

stabilizing the dimer and orienting tyrosine and lysine and other residues for optimal interaction with<br />

substrate and nucleotide cofactor.<br />

The role of a specific site on the outer hydrophobic surface of α helix F in dimerization was suggested<br />

recently when a Drosophila ADH mutant that does<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_202.html [4/5/2004 5:06:02 PM]


Document<br />

Page 203<br />

not form stable dimers was sequenced [48]. This ADH mutant has alanine-159 replaced with threonine.<br />

A 3D model of Drosophila ADH shows alanine-159 on the opposite surface of α helix F from tyrosine-<br />

153 and lysine-157 [48]. Alanine-159 along with alanine-158 form a hydrophobic anchor that stabilizes<br />

the dimer interface. These two residues of ADH and the homologous residues in other short-chain<br />

alcohol dehydrogenases have been overlooked in sequence analyses because they are not absolutely<br />

conserved. In fact, at least five amino acids are found in these positions among the different sec-alcohol<br />

dehydrogenases.<br />

G. Dimer Interface in 11β-and 17β-Hydroxysteroid Dehydrogenases<br />

Because α helix F at the dimer interface also contains the catalytic tyrosine and the near<strong>by</strong> lysine<br />

residue, any structural analysis of the catalytic site must also consider the structure of this part of the<br />

dimer interface. For this reason, we modeled α helix F on 11β-HSD-1 and -2 and 17β-HSD-1, -2, -3,<br />

and -4 to gain an insight into stabilizing interactions and how they may affect the catalytic site.<br />

H. Human 11β-HSD-1<br />

Figure 5 shows the modeled structure for the α helix F interface in human 11β-HSD-1, in which<br />

phenylalanine-188 and alanine-189 form an anchor. Alanine-189 is 3.5 Å and 4.7 Å from alanine-189<br />

and alanine-185, respectively, on the other subunit. The phenylalanine-188 side chain is 3.2 Å from<br />

glycine-192. There is a hydrogen bond between serine-185 and serine-196, which are 3.2 Å apart.<br />

Alanine-185 is 4.7 Å from phenylalanine-193. There also is a hydrophobic interaction between<br />

phenylalanine-193 and alanine-181, which are 3.9 Å apart.<br />

This web of interaction between side chains on the outer surface of α helix F on each subunit influences<br />

residues that have side chains oriented to the interior where catalysis occurs. Alanine-185, which is<br />

stabilized <strong>by</strong> interactions with phenylalanine-193, and serine-184, which interacts with serine-196, are<br />

between the conserved tyrosine-183 and lysine-187. Phenylalanine-193 is adjacent to phenylalanine-<br />

194, which is positioned into the catalytic site.<br />

I. Human 11β-HSD-2<br />

Figure 5 shows the α helix F interface of human 11β-HSD-2. Alanine-237 is about 3 Å from leucine-<br />

241; alanine-238 is about 3.7 Å from both alanine-238 and threonine-234. Threonine-234 has a<br />

stabilizing hydrophobic interaction<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_203.html [4/5/2004 5:06:03 PM]


Document<br />

Figure 6<br />

<strong>Structure</strong> of α helix F interface of mammalian 17β-hydroxysteroid dehydrogenases.<br />

The α helix F part of the dimer interface on 17β-hydroxysteroid dehydrogenases is<br />

shown along with side chains of the highly conserved tyrosine and lysine residues<br />

and three other residues that are oriented into the cavity that binds substrate and nucleotide<br />

cofactor. (a) Modeled structure of human 17β-hydroxysteroid dehydrogenase type 1.<br />

(b) Modeled structure of human 17β-hydroxysteroid dehydrogenase type 2. (c) Modeled<br />

structure of human 17β-hydroxysteroid dehydrogenase type 3. (d) Modeled structure of<br />

porcine 17β-hydroxysteroid dehydrogenase type 4.<br />

with leucine-242, which is 4.2 Å distant. Threonine-245 is 4.2 Å from the C α carbon of glycine-230.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_204.html (1 of 2) [4/5/2004 5:06:14 PM]<br />

Page 204


Document<br />

The type 1 and type 2 enzymes preferentially catalyze the reduction of the 11-keto group and the<br />

oxidation of the 11-hydroxyl group, respectively, on glucocorticoids. The chemistry of the side chains<br />

on methionine-243 in the type-2<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_204.html (2 of 2) [4/5/2004 5:06:14 PM]


Document<br />

Page 205<br />

enzyme and of phenylalanine-194 on 11β-hydroxysteriod dehydrogenase type 1 is quite different; it may<br />

be important in the different catalytic properties of these two enzymes.<br />

J. Human 17β-HSD-1<br />

Figure 6a shows the modeled α helix F interface in human 17β-hydroxysteroid dehydrogenase type 1 in<br />

which phenylalanine-160 and alanine-161 form an anchor. Both residues have important stabilizing<br />

interactions across the dimer interface. Alanine-161 is 4.1 Å from alanine-161 on the other subunit.<br />

Alanine-161 has a hydrophobic interaction with alanine-157, which is in the segment between the<br />

conserved tyrosine-155 and lysine-159. There is a hydrophobic<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_205.html [4/5/2004 5:06:24 PM]


Document<br />

Page 206<br />

interaction between alanine-157 and leucine-165, which are about 3.8 Å apart. Phenylalanine-160 is 4 Å<br />

from glycine-164. There also is a hydrogen bond between cysteine-156 and serine-168, which are 3.2 Å<br />

apart. This is an interesting structural property of residues in the segment between the conserved<br />

tyrosine and lysine residues: this segment is important in stabilizing dimers. This pattern is repeated in<br />

the other 11β- and 17β-hydroxysteroid dehydrogenases suggesting conservation of this stabilizing<br />

structure, although the residues are not as well conserved as the tyrosine and lysine.<br />

K. Human 17β-HSD-2<br />

Figure 6b shows the modeled α helix F interface in human 17β-hydroxysteroid dehydrogenase type 2.<br />

Alanine-237 is 3 Å from the hydrophobic part of the side chain of methionine-241 on the other subunit.<br />

Methionine-241 is 3.2 Å from serine-234. Alanine-230 is 3.7 Å from phenylalanine-242 and 4.5 Å from<br />

valine-245. Alanine-238, the other anchoring residue, is 4.1 Å from alanine-238 on the other subunit.<br />

L. Human 17β-HSD-3<br />

Figure 6c shows the modeled α helix F interface in human 17β-hydroxysteroid dehydrogenase-type 3.<br />

Alanine-203 is 3.1 Å from alanine-207. Phenylalanine-204 is 3.2 Å from the other phenylalanine-204<br />

and alanine-200. These are the only stabilizing interactions that we find in our analysis. Human 17βhydroxys-teroid<br />

dehydrogenase type 3 has the weakest interactions across the α helix F interface among<br />

the four types of 17β-hydroxysteroid dehydrogenases. The conformation of this part of 17βhydroxysteroid<br />

dehydrogenase type 3 could change upon binding of substrate, leading to other<br />

stabilizing interactions. And, of course other parts of the protein may have intersubunit interactions that<br />

stabilize the dimer. Alternatively, the hydrophobic surface of α helix F may interact with another protein<br />

or a membrane surface, a potentially important regulatory mechanism that we discuss later in this paper.<br />

M. Pig 17β-HSD-4<br />

Figure 6d shows the modeled α helix F interface in pig 17β-hydroxysteroid dehydrogenase type 4.<br />

Leucine-169 is 2.9 Å from glycine-173. Leucine-174 is 4.3 Å from alanine-162 and 3.3 Å from alanine-<br />

166. There also is a hydrogen bond between serine-165 and serine-175, which are 3 Å apart.<br />

N. Prospects for the Application of <strong>Structure</strong>-Function Analysis of Steroid Dehydrogenases in<br />

Hormone Therapy<br />

In the last two years there has been impressive progress in understanding the structure of steroid<br />

dehydrogenases that are important in regulating blood pres<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_206.html [4/5/2004 5:06:25 PM]


Document<br />

Page 207<br />

sure and the actions of reproductive hormones. This progress has come from several directions. First, the<br />

cloning and sequencing the dehydrogenases that regulate the actions of aldosterone, cortisol, estradiol,<br />

and testosterone. Second, determination of the 3D structure of 17β-HSD-1 and several homologs.<br />

Analyses of their 3D structures confirm a general principle that structural similarity is much higher than<br />

sequence similarity. This supports proposed molecular models of medically important steroid<br />

dehydrogenases using the alignment of their sequences onto the templates of 3D structural homologs.<br />

Models of 11β-HSD-1 and -2 are beginning to reveal important properties about these enzymes. We<br />

now have a good picture of the structural basis for specificity for NADPH and NADH in 11β-HSD-1<br />

and -2. With this information, we can now turn our attention to modeling cortisol in these two enzymes.<br />

This information will open up the possibility for developing analogs to regulate the actions of these two<br />

enzymes for use in regulating blood pressure and other physiological processes. The development of<br />

carbenoxolone, a water-soluble synthetic analog of glycyrrhetinic acid, shows that chemists can create<br />

compounds that have high affinity for 11β-HSD. The next task is to synthesize compounds that are<br />

specific for 11β-HSD-1 or 11β-HSD-2. Molecular modeling can contribute important information to<br />

solving this kind of problem.<br />

Similarly important information will come from the 3D models of 17β-HSD-1, -2, -3 and -4. These<br />

models will be useful in developing compounds to regulate estrogen and androgen action, which have<br />

important application in reproductive medicine and in treating estrogen-dependent breast tumors and<br />

androgen-dependent prostatic tumors. Many compounds in plants have estrogenic and androgenic<br />

activity; some of these compounds are likely to work via inhibition of one of the 17β-HSD enzymes<br />

[27]. Analogous to the development of carbenoxolone to regulate 11β-HSD, we can seek synthetic<br />

compounds that regulate specific types of 17β-HSD, which may be useful in reproductive medicine and<br />

in treating cancers.<br />

Considering the explosive pace of biomedical research and the new developments in computers for<br />

sophisticated structural analyses, the next few years promise to yield important advances in design of<br />

new hormone therapies <strong>based</strong> on the knowledge of the structure of steroid dehydrogenases.<br />

Acknowledgments<br />

We thank Drs. Tanaka, Nonaka, Nakanishi, Deyashiki, Hara, and Mitsui for providing us with the x-ray<br />

crystallographic coordinates of carbonyl reductase and 7α-hydroxysteroid dehydrogenase. The support<br />

of the Supercomputer Center of the University of California, San Diego is gratefully acknowledged.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_207.html [4/5/2004 5:06:27 PM]


Document<br />

References<br />

1. Funder JW, Pearce PT, Smith R, Smith AI. Mineralocorticoid action: target tissue specificity is<br />

enzyme, not receptor, mediated. Science 242; 1988:583–586.<br />

2. Edwards CRW, Stewart PM, Burt D, Brett L, McIntyre MA, Sutanto WS, De Kloet ER, Monder C.<br />

Localization of 11β-hydroxysteroid dehydrogenase-tissue specific protector for the mineralocorticoid<br />

receptor. Lancet 2; 1988:986–989.<br />

3. Monder C. Corticosteroids, receptors, and the organ-specific functions of 11β-hydroxysteroid<br />

dehydrogenase. FASEB J5; 1991:3047–3054.<br />

Page 208<br />

4. White PC. Disorders of aldosterone biosynthesis and action. New Eng J Med 331; 1994:250–258.<br />

5. Andersson S. 17β-hydroxysteroid dehydrogenase: isozymes and mutations. J Endocrinol 146;<br />

1995:197–200.<br />

6. Baker ME. Unusual evolution of 11β- and 17β-hydroxysteroid and retinol dehydrogenases. Bioessays<br />

18; 1996:63–70.<br />

7. New MI, Levine LS, Biglieri EG, Pareira J, Ulick S. Evidence for an unidentified steroid in a child<br />

with apparent mineralocorticoid hypertension. J Clin Endocrinol Metab 44; 1977:924–933.<br />

8. Ulick S, Levine LS, Gunczler P, Zanconato G, Ramirez LC, Rauh W, Rosler A, Bradlow HL, New<br />

MI. A syndrome of apparent mineralocorticoid excess associated with defects in the peripheral<br />

metabolism of cortisol. J Clin Endocrinol Metab 49; 1979:757–764.<br />

9. Funder JW, Pearce PT, Myles K, Roy LP. Apparent mineralocorticoid excess,<br />

pseudohypoaldosteronism, and urinary electrolyte excretion: toward a redefinition of mineralocorticoid<br />

action. FASEB J 4; 1990:3234–3238.<br />

10. Stewart PM, Edwards CRW. The cortisol-cortisone shuttle and hypertension. J Steroid Biochem<br />

Molec Biol 40; 1991:501–509.<br />

11. Monder C. Comparative aspects of 11β-hydroxysteroid dehydrogenase. Testicular 11βhydroxysteroid<br />

dehydrogenase: development of a model for the mediation of Leydig cell function <strong>by</strong><br />

corticosteroids. Steroids 59; 1994:69–73.<br />

12. Arriza JL, Weinberger C, Cerelli G, Glaser TM, Handelin BL, Housman DE, Evans RM. Cloning of<br />

the human mineralocorticoid receptor complementary DNA: structural and functional kinship with the<br />

glucocorticoid receptor. Science 237; 1987:268–275.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_208.html (1 of 2) [4/5/2004 5:06:29 PM]


Document<br />

13. Wilson RC, Krozowski ZS, Li K, Obeyesekere VR, Razzaghy-Azar M, Harbison MD, Wei JQ,<br />

Shackleton CHL, Funder JW, New MI. A mutation in the HSD11B2 gene in a family with apparent<br />

mineralocorticoid excess. J Clin Endocrinology Metab 80; 1995:2263–2266.<br />

14. Mune T, Rogerson FM, Nikkila H, Agarwal AK, White PC. Human hypertension caused <strong>by</strong><br />

mutations in the kidney isozyme of 11β-hydroxysteroid dehydrogenase. Nature Genetics 10;<br />

1995:394–399.<br />

15. Baker ME. Licorice and enzymes other than 11β-hydroxysteroid dehydrogenase. Steroids 59;<br />

1994:136–141.<br />

16. Peltoketo H, Isomaa V, Vihko R. Genomic organization and DNA sequences of human 17βhydroxysteroid<br />

dehydrogenase genes and flanking regions. Eur J Biochem 209; 1992:459–466.<br />

17. Wu L, Einstein M, Geissler WM, Chan HK, Elliston KO, Andersson S. Expression cloning and<br />

characterization of human 17β-hydroxysteroid dehydrogenase type 2,<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_208.html (2 of 2) [4/5/2004 5:06:29 PM]


Document<br />

a microsomal enzyme possessing 20α-hydroxysteroid dehydrogenase activity. J Biol Chem 268;<br />

1993:12964–12969.<br />

18. Geissler WM, Davis DL, Wu L, Bradshaw KD, Patel S, Mendonca BB, Elliston KO, Wilson JD,<br />

Russell DW, Andersson S. Male pseudohermaphrodites, caused <strong>by</strong> mutations to testicular 17βhydroxysteroid<br />

dehydrogenase-3. Nature Genetics 7; 1994:34–39.<br />

Page 209<br />

19. Napoli JL, Boerman MHEM, Chai X, Zhai Y, Fiorella PD. Enzymes and binding proteins affecting<br />

retinoic acid concentrations. J Ster Biochem Molec Biol 55; 1995:589–600.<br />

20. Baker ME. Evolution of enzymatic regulation of prostaglandin action: novel connections to<br />

regulation of human sex and adrenal function, antibiotic synthesis and nitrogen fixation. Prostaglandins<br />

42; 1991:391–407.<br />

21. Tannin GM, Agarwal AK, Monder C, New MI, White PC. The human gene for 11β-hydroxysteroid<br />

dehydrogenase. J Biol Chem 266; 1991:16653–16658.<br />

22. Albiston AL, Obeyesekere VR, Smith RE, Krozowski ZS. Cloning and tissue distribution of the<br />

human 11β-hydroxysteroid dehydrogenase type 2 enzyme. Mol Cell Endocrinol 105; 1994:R11–R17.<br />

23. Agarwal AK, Mune T, Monder C, White PC. NAD +-dependent isoform of 11β-hydroxysteroid<br />

dehydrogenase. Cloning and characterization of cDNA from sheep kidney. J Biol Chem 269;<br />

1994:25959–25962.<br />

24. Naray-Fejes-Toth A, Fejes-Toth G. Expression cloning of the aldosterone target cell-specific 11βhydroxysteroid<br />

dehydrogenase from rabbit collecting duct cells. Endocrinology 136; 1995:2579–2586.<br />

25. Cole TJ. Cloning of the mouse 11β-hydroxysteroid dehydrogenase type 2 gene: tissue specific<br />

expression and localization in distal convoluted tubules and collecting ducts of the kidney.<br />

Endocrinology 136; 1995:4693–4696.<br />

26. Leenders F, Adamski J, Husen B, Thole, HH, Jungblut PW. Molecular cloning and amino acid<br />

sequence of the porcine 17β-estradiol dehydrogenase. Eur J Biochem 222; 1994:221–227.<br />

27. Baker ME. Endocrine activity of plant-derived compounds: an evolutionary perspective. Proc Soc<br />

Exper Biol Med 208; 1995: 131–138.<br />

28. Ghosh D, Pletnev VZ, Zhu D-W, Wawrkak Z, Duax WL, Pangborn W, Labrie F, Lin S–X. <strong>Structure</strong><br />

of human estrogenic 17β-hydroxysteroid dehydrogenase at 2.20 Aº resolution. <strong>Structure</strong> 3;<br />

1995:503–513.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_209.html (1 of 2) [4/5/2004 5:06:31 PM]


Document<br />

29. Baker ME. Genealogy of regulation of human sex and adrenal function, prostaglandin action,<br />

snapdragon and petunia flower colors, antibiotics, and nitrogen fixation: functional diversity from two<br />

ancestral dehydrogenases. Steroids 56; 1991:354–360.<br />

30. Persson B, Krook M, Jornvall H. Characteristics of short-chain alcohol dehydrogenases and related<br />

enzymes. Eur J Biochem 200; 1991:537–543.<br />

31. Krozowski Z. 11β-hydroxysteroid dehydrogenase and the short chain alcohol dehydrogenase<br />

(SCAD) superfamily. Mol Cell Endocrinol 84; 1992:C25–C31.<br />

32. Baker ME. Protochlorophyllide reductase is homologous to human carbonyl reductase and pig 20βhydroxysteroid<br />

dehydrogenase. Biochem J 300; 1994:605–607.<br />

33. Varughese KI, Xuong NH, Kiefer PM, Matthews DA, Whiteley JM. Structural and mechanistic<br />

characteristics of dihydropteridine reductase: a member of the Tyr- (Xaa) 3-Lys-containing family of<br />

reductases and dehydrogenases. proc Natl Acad Sci USA 91; 1994:5582–5586.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_209.html (2 of 2) [4/5/2004 5:06:31 PM]


Document<br />

Page 210<br />

34. Ghosh D, Wawrzak Z, Weeks CM, Duax WL, Erman M. The refined three-dimensional structure of<br />

3α,20β-hydroxysteroid dehydrogenase and possible roles of the residues conserved in chart-chain<br />

dehydrogenases. <strong>Structure</strong> 2; 1994:629–640.<br />

35. Dessen A, Quemard A, Blanchard JS, Jacobs Jr, WR, Sacchettini JC. Crystal structure and function<br />

of the isoniazid target for Mycobacterium tuberculosis. Science 267; 1995:1638–1641.<br />

36. Rafferty JB, Simon JW, Baldock C, Artymiuk PJ, Stuitje AR, Slabas AR, Rice DW. Common<br />

themes in redox chemistry emerge from the X-ray structure of oilseed rape, Brassica napus, enoyl acyl<br />

carrier protein reductase. <strong>Structure</strong> 3; 1995:927–938.<br />

37. Tanaka N, Nonaka T, Nakanishi M, Deyashiki Y, Hara A, Mitsui Y. Crystal structure of the ternary<br />

complex of mouse lung carbonyl reductase at 1.8 Å resolution: the structural origin of coenzyme<br />

specificity in the short-chain dehydrogenase/reductase family. <strong>Structure</strong> 4; 1996:33–45.<br />

38. Tsigelny I, Baker ME. <strong>Structure</strong>s stabilizing the dimer interface on human 11β-hydroxysteroid<br />

dehydrogenase-types 1 and 2 and human 15-hydroxyprostaglandin dehydrogenase and their homologs.<br />

Biochem Biophys Res Commun 217; 1995:859–868.<br />

39. Tsigelny I, Baker ME. <strong>Structure</strong>s important in mammalian 11β-and 17β-hydroxysteroid<br />

dehydrogenases. J Ster Biochem Molec Biol 55; 1995:589–600.<br />

40. Obeid J, White PC. Tyr-179 and lys-183 are essential for enzymatic activity of 11β-hydroxysteroid<br />

dehydrogenase. Biochem Biophys Res Comm 188; 1992:222–227.<br />

41. Puranen TJ, Poutanen MH, Peltoketo HE, Vihko PT, Vihko RK. Site-directed mutagenesis of the<br />

putative active site of human 17β-hydroxysteroid dehydrogenase type 1. Biochem J 304; 1994:289–293.<br />

42. Chen Z, Lu L, Shirley M, Lee WR, Chang SH. Site-directed mutagenesis of glycine-14 and two<br />

“critical” cysteinyl residues in Drosophila alcohol dehydrogenase. Biochemistry 29; 1990:1112–1118.<br />

43. Chen Z, Lin Z-G, Lee WR, Chang SH. Role of aspartic acid-38 in the cofactor specificity of<br />

Drosophila alcohol dehydrogenase. Eur J Biochem 202; 1991:263–267.<br />

44. Chen Z, Jiang JC, Lin Z-G, Lee WR, Baker ME, Chang SH. Site-specific mutagenesis of Drosophila<br />

alcohol dehydrogenase: evidence for involvement of tyrosine-152 and lysine-156 in catalysis.<br />

Biochemistry 32; 1993:3342–3346.<br />

45. Cols N, Marfany G, Atrian S, Gonzalez-Duarte R. Effect of site-directed mutagenesis on conserved<br />

positions of Drosophila alcohol dehydrogenase. FEBS Lett 319; 1993:90–94.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_210.html (1 of 2) [4/5/2004 5:06:33 PM]


Document<br />

46. Ribas dePoplana L, Fothergill-Gilmore LA. The active site architecture of a short chain<br />

dehydrogenase defined <strong>by</strong> site-directed mutagenesis and structure modeling. Biochemistry 33;<br />

1994:7047–7055.<br />

47. Chen Z, Tsigelny I, Lee WR, Baker ME, Chang SH. Adding a positive charge at residue 46 of<br />

Drosophila alcohol dehydrogenase increases cofactor specificity for NADP +. FEBS Lett 356;<br />

1994:81–85.<br />

48. Chenevert S, Fossett N, Lee WR, Tsigelny I, Baker ME, Chang SH. Amino acids important in<br />

enzyme activity and dimer stability for Drosophila alcohol dehydrogenase. Biochem J 308;<br />

1995:419–423.<br />

49. Chothia C, Lesk AM. The relation between the divergence of sequence and structure in proteins.<br />

EMBO J 5; 1986:823–826.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_210.html (2 of 2) [4/5/2004 5:06:33 PM]


Document<br />

50. Greer J. Comparative modeling of homologous proteins. Methods Enzymol 202; 1991:239–252.<br />

51. Branden C, Tooze J. Introduction to protein structure. New York: Garland Publishing, 1991.<br />

52. Ring CS, Cohen FE. Modeling protein structures: construction and their applications. FASEB J 7;<br />

1993:783–790.<br />

53. Baker ME. Sequence analysis of steroid-and prostaglandin- metabolizing enzymes: application to<br />

understanding catalysis. Steroids 59; 1994:248–258.<br />

Page 211<br />

54. Wierenga RK, De Maeyer MC, Hol WGJ. Interaction of pyrophosphate moieties with α-helixes in<br />

dinucleotide binding proteins. Biochemistry 24; 1985:1346–1357.<br />

55. Wierenga RK, Terpstra PP, Hol WGJ. Prediction of the occurrence of the ADP-binding βαβ-fold in<br />

proteins, using an amino acid sequence fingerprint. J Mol Biol 187; 1986:101–107.<br />

56. Weber P, Salemme FR. Structural and functional diversity in four-α-helical proteins. Nature 287;<br />

1980:82–84.<br />

57. Chou K-C, Maggiora GM, Nemethy G, Scheraga HA. Energetics of the structure of the four-α-helix<br />

bundle in proteins. Proc Natl Acad Sci USA 85; 1988:4295–4299.<br />

58. Presnell SR, Cohen FE. The topological distribution of four-α-helical proteins. Proc Natl Acad Sci<br />

USA. 86; 1989:6592–6596.<br />

59. Harris NL, Presnell SR, Cohen FE. Four helix bundle diversity in globular proteins. J Mol Biol 236;<br />

1994:1356–1368.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_211.html [4/5/2004 5:06:34 PM]


http://legacy.netlibrary.com/reader/message.asp?message=811&BookID=12640&FileName=Page_212.html<br />

The requested page could not be found.<br />

Return to previous page<br />

http://legacy.netlibrary.com/reader/message.asp?message=811&BookID=12640&FileName=Page_212.html [4/5/2004 5:06:37 PM]


Document<br />

8<br />

<strong>Design</strong> of ATP Competitive Specific Inhibitors of Protein Kinases Using<br />

Template Modeling<br />

Janusz M. Sowadski, * Charles A. Ellis, * Rolf Karlsson *<br />

University of California, San Diego, La Jolla, California<br />

Madhusudan<br />

Scripps Research Institute, La Jolla, California<br />

I. Protein Kinases and Diseases<br />

Page 213<br />

The protein kinase family encompasses more than three hundred members of critically important<br />

enzymes, each one with a specific role or function within the cell. These enzymes, ATPphosphotransferases,<br />

recognize target proteins and through the phosphorylation of specific sites either<br />

activate or deactivate a particular pathway of signal transduction. Many of these signaling pathways are<br />

associated with cell surface receptors, which are located in the membranes that surround cells. The<br />

difference between the families of protein kinases is that they have different targets and generally fall<br />

into two major classes:<br />

The serine/threonine protein kinases transfer a phosphate from ATP to a serine or threonine residue in<br />

the target protein. This class of enzymes are generally associated with cytoplasmic signaling events.<br />

The tyrosine protein kinases transfer phosphate from ATP to tyrosine residues in the target protein and<br />

are generally associated with receptors that become activated after binding a growth factor or other<br />

ligand.<br />

Protein kinases are significant targets for therapeutic drug development and have been implicated as the<br />

disease causing components of numerous tumor<br />

* Current affiliation: Tufts University, Boston, Massachusetts.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_213.html [4/5/2004 5:06:38 PM]


Document<br />

Page 214<br />

viruses. Specifically, it is the deregulation of the activity of protein kinases that leads to disease <strong>by</strong><br />

tumor viruses. The importance of this deregulation can be dramatically illustrated <strong>by</strong> the large number of<br />

viral oncogenes (or cancer causing genes) that encode structurally modified protein kinases. These<br />

deregulated enzymes are able to <strong>by</strong>pass the normal tightly regulated processes of growth control, leading<br />

to acute malignant transformation. These oncogenes are one of the first examples of the identification of<br />

disease-causing genes. Many of these viral genes have subsequently been implicated in human diseases.<br />

Malignant tissues also share the common characteristic of an acquired independence from controls. The<br />

receptor—for example, PDGF and EGFR—can be stimulated <strong>by</strong> a ligand coming either from the cell<br />

itself (autocrine) or from near<strong>by</strong> tissues (paracrine). Regardless of the mechanism leading to receptor<br />

activity, the resulting kinase activity results in a cascade of signals that turn on cellular proliferation<br />

programs. Therefore, selective inhibition of receptor tyrosine kinase will block tyrosine kinase driven<br />

cell proliferation resulting in antitumor activity. In addition to cancer, a growing number of<br />

nonmalignant proliferative diseases, (e.g., psoriasis, atherosclerosis, restenosis, fibrosis, etc.) or<br />

inflammatory responses (e.g., septic shock, asthma, osteo and rheumatoid arthritis, etc.) involve<br />

dysfunctional signaling pathways. Successful development of drugs that target this class of enzymes will<br />

depend on the discovery of selective inhibitors designed for the appropriate protein kinase within the<br />

family.<br />

In the past several years there has been an explosion of structural studies within the protein kinase<br />

family [1–8]. These studies, initiated <strong>by</strong> the crystal structure of Protein Kinase A [9–12] (CAPK) have<br />

shown that all members of the protein kinase family fold into a uniform three-dimensional catalytic core.<br />

Yet this uniform three-dimensional fold exhibits both different surface charges and at least two major<br />

conformations.<br />

II. Protein Kinase Template<br />

The stereo view of the ribbon diagram of cAPK is presented in Figure 1a. The overall topology of the<br />

core extending from strand 1 through helix h, Figure 1b, is identical (except helix B) with the eight other<br />

structures of protein kinases determined to this point. Furthermore, Figure 2 presents an overall<br />

structural comparison of the catalytic cores of the five kinases, cAPK, CDK2, CDK2-CYCLIN, IR, and<br />

MAP. The N-terminal helix A, which is present only in the cAPK crystal structure, is anchored <strong>by</strong><br />

myristic acid in the mammalian bovine heart of cAPK. Myristic acid inserts itself into the hydrophobic<br />

pocket of the lower lobe of the enzyme, which results in the structural ordering of helixA and Ser10[13],<br />

one of the four autophosphorylation sites.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_214.html [4/5/2004 5:06:40 PM]


Document<br />

Figure 1<br />

Diagram of cAPK fold. (a) Stereo MOLSCRIPT diagram. (b) The key loops<br />

are as follows: phosphate anchor located between strand 1 and strand 2,<br />

catalytic loop located between strands 6 and 7, DFG motif located between<br />

strands 8 and 9, activation loop including P+1 site between strand 9 and<br />

helix F. Phosphorylation site Thr197 is indicated <strong>by</strong> a large circle,<br />

inhibitor PKI (5–24) is colored in dark, and the P site in the peptide<br />

is shown in the dark circle. This figure has been generated using<br />

MOLSCRIPT [27].<br />

Page 215<br />

Following the connectivity diagram, helix A is connected to β-strand 1, then to the phosphate anchor<br />

encompassing signature motif Gly50XGly 52XXGly55. The β-strand 2 is followed <strong>by</strong> β-strand 3<br />

carrying invariant Lys72. Three antiparallel beta strands create the unique fold of the nucleotide binding<br />

site of the protein kinase. The β-strands 3 is followed <strong>by</strong> helix B, which is present only in cAPK, helix<br />

C, and β-strands 4 and 5. Helix C shows the largest displacements among many different protein kinase<br />

structures and consists of invariant Glu91, which forms a salt bridge with Lys72. This salt bridge is<br />

absent in the inactive CDK-2 structure [1] but present in the crystal structure of the complex of CDK-2<br />

and its activator-cyclin [7]. Displacement of helix C is perhaps most pronounced in the case of the<br />

insulin receptor tyrosine kinase structure (IRK) [3]. In PKA, Phe185 resides in the hydrophobic pocket<br />

formed <strong>by</strong> the hydrophobic residues of helix C (upper lobe) and Tyr164 (lower lobe). In the IRK crystal<br />

structure, the hydrophobic residues of helix C, which provide the pocket for invariant Phe185 (of DFG<br />

motif), no longer interact with<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_215.html [4/5/2004 5:06:47 PM]


Document<br />

Figure 1<br />

(Continued)<br />

this residue. The DFG motif in IRK occupies the ATP site, which blocks the access of ATP.<br />

Page 216<br />

The division between the upper and lower lobes of the enzyme is well defined <strong>by</strong> the two major<br />

conformations of the upper lobe observed in the crystal structures of cAPK. One conformation has been<br />

observed in the orthorhombic crystals of recombinant cAPK [9,10,14] and another in the cubic<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_216.html (1 of 2) [4/5/2004 5:06:54 PM]


Document<br />

Figure 2<br />

The diagrams of the Cα trace of the five kinases, cAPK, CDK2, CDK2-CYCLIN,<br />

IR, MAP. The thin line represents crystallographically determined homologous<br />

regions of the five kinases (R. Karlsson and J.M. Sowadski, personal<br />

communication).<br />

Page 217<br />

crystals of bovine heart mammalian cAPK [15,16]. A comparison between the two structures has shown<br />

that there is rotation of the upper lobe <strong>by</strong> 15 degrees and a translation of 1.9 Å in the mammalian<br />

structure, which results in the opening of the nucleotide binding cleft.<br />

The motion of this lobe, which includes His87, one of the ligands to Thr197 [15], indicates that the<br />

phosphorylation of this site will be important for conformational diversity of the upper lobe. This is<br />

confirmed <strong>by</strong> the varying degrees of displacement of the upper lobe of all structures of the inactive<br />

unphosphorylated protein kinases (see review [17]). The lower lobe of the enzyme starts with helix D,<br />

which is followed <strong>by</strong> helix E and β-strands 6 and 7. The catalytic loop connecting both strands consists<br />

of a critical set of residues with Tyr164 and Arg165 at the beginning of the loop. The Tyr164 residue<br />

forms a hydrogen bond with invariant Asp220. The Arg165 residue, which is present in a great majority<br />

of protein kinases, provides two hydrogen bonds to the oxygens of the phosphate of Thr197. Invariant<br />

Asp166 (the catalytic base) and Asn171 (the ligand to one of the metal sites) are also located within this<br />

loop.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_217.html (1 of 2) [4/5/2004 5:07:00 PM]


Document<br />

Page 218<br />

The catalytic loop is the region of divergence between Ser/Thr and Tyr kinases. In cAPK and all Ser/Thr<br />

Kinases, Lys168 interacts with the γ phosphate of ATP during catalysis [12]. The role of Lys is replaced<br />

<strong>by</strong> Arg [9] and the insulin receptor tyrosine kinase structure [3] shows Arg1136 in a similar position as<br />

Lys168 in the active site of cAPK.<br />

The catalytic loop and β-strand 7 are followed <strong>by</strong> β-strand 8 and a short DFG conserved motif. This<br />

conserved motif consists of invariant Asp184, a ligand to the metal site, and invariant Phe185. The DFG<br />

motif is followed <strong>by</strong> β-strand 9, an activation loop that includes Thr197. The activation loop differs<br />

considerably among unphosphorylated kinases, CDK-2, ERK-2, and insulin receptor kinase.<br />

Furthermore, two crystal structures, twitchin protein kinase [2] and phosphorylase kinase [5], both lack a<br />

phosphorylation site in this region. In the first structure, twitchin protein kinase, Thr197 and Arg165, are<br />

replaced <strong>by</strong> hydrophobic residues, Val6098 and Leu6062 respectively [2]. This was predicted using<br />

modeling [18] and subsequently confirmed <strong>by</strong> the three-dimensional structure. In the second structure,<br />

phosphorylase kinase, the Thr197 is replaced <strong>by</strong> Glu182, which interacts with Arg148 [5]. Hence, in<br />

both structures the regulatory function of the phosphorylation site is replaced <strong>by</strong> a stable scaffold<br />

secured <strong>by</strong> either hydrophobic or electrostatic interactions. Since it has been shown that this loop<br />

provides a stable template for PKI(5–24) binding in cAPK, the status of phosphorylation of the<br />

activation loop critically affects the substrate binding. This is demonstrated in c-Src, a homolog of the<br />

Rous Sarcoma virus oncogene <strong>by</strong> mutation of Arg385, which is predicted to interact with the phosphate<br />

of Tyr416, and results in loss of activity toward the exogenous substrate [19].<br />

The activation loop is followed <strong>by</strong> a P+1 loop which accommodates the P+1 site of the substrate. The<br />

P+1 loop is followed <strong>by</strong> invariant Glu208, which forms a salt bridge with invariant Arg280. This<br />

conserved pair plays a structural role and as the structure of CK-1 [16] shows, it can be replaced <strong>by</strong><br />

other charged residues that maintain the same fold of the lower lobe. The P+1 loop and Glu 208 are<br />

followed <strong>by</strong> helix F consisting of invariant Asp220, followed <strong>by</strong> helices G, H, and I. The helix J and the<br />

C-terminal tail of cAPK, which are absent in other protein kinases, undergo a large motion during the<br />

cleft opening.<br />

The opening of the cleft results in loss of hydrogen bonds provided <strong>by</strong> the γ phosphate of ATP and the<br />

peptide that would bridge the lower and upper lobe of the enzyme. The motion of the upper domain<br />

increases the accessibility of the ATP binding site and one can envision that in the “open” conformation<br />

ATP binds. Yet, in the “closed” conformation ATP and its γ phosphate are positioned for a nucleophilic<br />

attack on the substrate. The motion of this lobe—which includes His87, one of the ligands to Thr197<br />

[15]—indicates that the phosphorylation of this site will be important for conformational diversity of the<br />

upper lobe. This is confirmed <strong>by</strong> the varying degrees of displacement of the upper lobe<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_218.html [4/5/2004 5:07:03 PM]


Document<br />

of all structures of the inactive unphosphorylated protein kinases (see review [17]).<br />

Page 219<br />

The various displacements of the conserved upper domain of the catalytic cores of various kinases<br />

documented <strong>by</strong> crystallographic work suggest that this is the important underlying mechanism of<br />

catalysis. Analysis of crystal contacts of various kinases is however required to define the extent of<br />

displacement due to the lattice forces. In the case of cAPK, the displacement as observed for mammalian<br />

cAPK in the cubic crystal form is due to the intermolecular interaction in the lattice [20]. Analysis of the<br />

two crystal structures of the cell cycle-controlling kinases clearly shows two binding modes of ATP. In<br />

the inactive state without cyclin, ATP binding of its triphosphate moieties is different from that in the<br />

active form with cyclin bound. The major difference is the re-arrangement upon cyclin binding of the<br />

conserved Lys33-Glu51 pair, which is responsible for the binding of the α and β phosphates of ATP.<br />

III. Crystallographic Analysis of Substrate Specificities of Individual Kinases<br />

The most important contribution of subsequent crystallographic studies has been the confirmation of the<br />

structural homology extending through the members of this family of enzymes. The crystal structures of<br />

CDK-2 [1], ERK-2 [5], twitchin [2], insulin receptor kinase [3], phosphorylase kinase, CK-1 [6], along<br />

with structure of calcium/calmodulin-dependent protein kinase I [8] provide solid proof for the structural<br />

conservation of the catalytic core in the family. This is further confirmed <strong>by</strong> the recent structure of the<br />

active complex of CDK2/cyclin, which shows that Lys33-Glu51 pair is at a distance of 3.0 Å [7] as<br />

predicted in the model of CDK-2 <strong>based</strong> on the cAPK structure [21]. The structure of the complex has<br />

also confirmed the binding of cyclin to helix C and to the upper lobe, demonstrating the mechanism of<br />

activation <strong>by</strong> cyclin that results in bridging the invariant residues into the common network of distances<br />

required <strong>by</strong> structural homology of the protein kinase catalytic core.<br />

The crystallographic analysis of the structural homology of protein kinases can now be carried out using<br />

structures of various kinases to find a common search model to be used in molecular replacement<br />

methods (J.M. Sowadski and R. Karlsson private communication). The structures of various kinases<br />

have been used as search models to solve the structure of the cAPK using cAPK diffraction data. The<br />

best search model consists of fragments of the catalytic core excluding the activation loop, inserts, and<br />

upper lobe due to rotational motion observed in each structure. A structure solution has been found for<br />

several protein kinases using this selected model as shown in Figure 2. One of the most critical aspects<br />

of this analysis is the presence of the structurally<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_219.html [4/5/2004 5:07:06 PM]


Document<br />

Page 220<br />

conserved substrate-binding cleft as observed in the crystal structure of cAPK:PKI complex (Figure<br />

la,b). This finding allows the charges within this cleft to be predicted using the amino acid sequence of<br />

any given kinase.<br />

In the analysis of the structural data of other protein kinases, it is noted that only cAPK has been<br />

crystallized with its specific peptide inhibitor. Nevertheless, three other structures of protein kinases<br />

compared with the structure of the cAPK-PKI complex provide substantial evidence for the conservation<br />

of the substrate binding cleft. The substrate binding cleft of the phosphorylase kinase structure has been<br />

analyzed in detail and it is clear that all amino acids of the known specific substrate can be built into the<br />

PKI model and all required corresponding charges can be found in the cleft of the phosphorylase kinase<br />

structure. In the CK-1 structure determined without a peptide, the requirement of the peptide specificity<br />

resides on the P-3 site, which has to be phosphorylated. An analysis of the surface charges of the cleft of<br />

the CK-1 structure reveals the exact correspondence of the residues required to interact with a<br />

phosphorylated substrate at this site.<br />

Finally, the tripeptide of the pseudosubstrate site of IRK consisting of the Asp-Tyr-Tyr motif has<br />

corresponding charges in the structure of the enzyme's substrate cleft and confirms the data obtained<br />

from the degenerated peptide library for the unique sequence motifs of nine tyrosine kinases [22]. It is<br />

becoming increasingly clear that the wealth of structural data of protein kinases with cAPK as a<br />

prototype provides evidence for two important features concerning substrate binding. First, the substrate<br />

binding cleft is structurally conserved and second, the surface charges of this cleft and hydrophobic<br />

cavities on the surface are very diverse and correspond to the specificity requirement of the substrate for<br />

individual protein kinases (see Figure la). It is now possible to use the structural conservation of the<br />

substrate binding cleft to predict the charges and hydrophobic residues of the cleft to define substrate<br />

specificities for individual kinases.<br />

IV. Crystallographic Analysis of the ATP Binding Site Reveals Distinct Differences Utilized for the<br />

Further <strong>Design</strong> of Specific Inhibitors<br />

The diagram elucidating detailed interactions of ATP with the enzyme is presented in Figure 3a,b. Six<br />

out of nine invariant residues of the catalytic core of protein kinase are involved in ATP binding and<br />

catalysis. The key residues that hold the β and γ phosphates in position are the phosphate anchor, the<br />

metal sites, and Lys168. The amides of the residues—Phe54, Gly55, and Ser53—are essential for the<br />

position of the γ and β phosphates. The metal site coordinated <strong>by</strong> invariant Asp184 is also sequestered<br />

<strong>by</strong> the β and γ phosphates and the metal<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_220.html [4/5/2004 5:07:08 PM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_221.html (1 of 2) [4/5/2004 5:07:18 PM]<br />

Page 221


Document<br />

Figure 3<br />

(a) Ternary complex of MnATP with the inhibitor peptide PKI (5–24).<br />

Glu121 and Val123 of the conserved linker region of the protein kinase catalytic core<br />

form the bidentate hydrogen bond with the 6-amino group and N1 nitrogen of the purine<br />

base. Thr183, non-conserved, forms a hydrogen bond with the N7 position of purine.<br />

2'-OH of ribose interacts with the side chain of Glu127 of helix D and P-3 Arg of the<br />

specific inhibitor while the 3'-OH interacts with Glu170 of the catalytic loop. (b) The<br />

local environment of serine nucleophile at P site (left site) and local environment of<br />

phosphorylated P site serine (right side). The side chain of the catalytic base is at<br />

hydrogen bond distance from-OH of the Ser nucleophile which defines the conserved<br />

substrate binding P site.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_221.html (2 of 2) [4/5/2004 5:07:18 PM]


Document<br />

Page 222<br />

site coordinated <strong>by</strong> invariant Asn171 is also sequestered <strong>by</strong> the α and γ phosphates. The residue Lys168<br />

is at hydrogen-bond distance from the γ phosphate. The postulated in-line mechanism of<br />

phosphotransfer in cAPK [23] can be examined through an analysis of the MnATP and PKI (5–24), Ser<br />

substrate peptide and ADP complexes [24]. A comparison between cAPK and IRK structures indicates<br />

that Ser versus Tyr specificity is obtained <strong>by</strong> displacement of the substrate binding site in such a way<br />

that the hydroxyl of nucleophiles of both Ser and Tyr fall in the same point in the active site facing<br />

corresponding catalytic bases Asp1132 and Asp166 for IRK and cAPK respectively.<br />

The purine base of ATP is anchored to the enzyme <strong>by</strong> three hydrogen bonds, two of them involve the 6amino<br />

group and N1 nitrogen, which interact with the backbone atoms of Glu121 and Val123. the 6amino<br />

group and N1 nitrogen of the purine base form the hydrogen bonds also in the structure of CK-1<br />

and in the structure of phosphorylase kinase. Yet, in the structure of inactive CDK-2, the purine base<br />

forms only one hydrogen bond via its N6 position. While N7 nitrogen interacts directly in cAPK with<br />

the side chain of Thr183 in CK-1, this nitrogen interacts directly with Glu55 and Tyr59 via two<br />

hydrogen-bonded water molecules and in phosposhorylase kinase via one water molecule. Hence, the<br />

N7 nitrogen and its interaction with the enzyme is a region for potential modification of ATP<br />

competitive inhibitors. Ribose is held <strong>by</strong> both enzyme (Glu127 and Glu170) and inhibitor (P-3 Arg).<br />

While the side chain of Glu127 interacts with 2'-OH of ribose in cAPK, in CK-1 the 2'-OH interacts via<br />

two water molecules with Ser91 and Asp94. In ERK-2, 2'-OH interacts with Asp109. This region has<br />

been utilized to design ATP-<strong>based</strong> specific inhibitors <strong>by</strong> modifications of the ATP-competitive<br />

nonspecific inhibitor staurosporine [25], see Figure 4.<br />

In the model cAPK with bound staurosporine inhibitor, the lactam amide group of the inhibitor functions<br />

as a bidentate hydrogen bond donor-acceptor<br />

Figure 4<br />

Chemical structures of staurosporine inhibitor, left, and<br />

CGP52411.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_222.html [4/5/2004 5:07:22 PM]


Document<br />

Figure 5<br />

(a) Staurosporine molecule docked in the ATP binding site of PKA.<br />

hydrogen bonds anchoring staurosporine molecule in the active site of PKA<br />

consists of the 6-amino group and N1 nitrogen and carbonyl of Glu121 and<br />

the amide hydrogen atom of Val123. This bidentate hydrogen bond formation<br />

has been observed in all complexes of protein kinases and ATP solved so far.<br />

(b) Inhibitor of CGP 52411 and ATP docked on the active site of PKA.<br />

Residues of Glu127 and Glu170 are also shown and these are not<br />

conserved in the EGFR kinase.<br />

Page 223<br />

(Figure 5a). This key observation is supported <strong>by</strong> chemical data of lactam amide derivatives, which<br />

provide a plausible model of staurosporine inhibition. This is in the protonated boat-type conformation<br />

found to fit in the ATP binding cleft with minimal steric hindrance. In this model, the 4-amino group<br />

forms hydrogen bonds with the backbone carbonyl of Glu170 and the carboxylate group of Glu127 of<br />

cAPK. A model of the EGFR kinase shows that Glu127 and 170 are replaced <strong>by</strong> Cys and Arg,<br />

respectively (Figure 5b). Replacement of Glu127 <strong>by</strong> Cys is critical according to the model and explains<br />

the several-fold decrease of potency of staurosporine inhibitor toward the EGFR kinase (IC 50=630 nM<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_223.html (1 of 2) [4/5/2004 5:07:30 PM]


Document<br />

Figure 5<br />

(Continued)<br />

Page 224<br />

versus IC 50=15 nM for cAPK). Yet enhancement of specificity utilizing the inhibitor CGP52411 (Figure<br />

4), whose selectivity originates in the occupancy <strong>by</strong> one of the anilino moieties of the inhibitor in the<br />

region of the enzyme cleft that normally binds the ribose ring of ATP, is considerable. The inhibitor<br />

CGP52411 inhibits the EGFR tyrosine kinase with an IC 50 value of 300 nM while it is less active <strong>by</strong> at<br />

least two orders of magnitude on a panel of protein kinases including cAPK, phosphorylase kinase,<br />

casein kinase, protein kinase C (most isoforms), and v-abl, c-lyn, c-fgr tyrosine kinases.<br />

Hence, the three-dimensional model of EGFR tyrosine kinase rationalizes the specificity of the<br />

CGP52411 inhibitor. This model suggests that analysis of the putative regions of the ribose binding of<br />

ATP in other kinases through template modeling would provide the required chemical modification of<br />

pharmacophores to enhance their selectivity. In modeling, the use of the bidendate hydrogen bonding<br />

provided <strong>by</strong> the linker region (Figure 3a) of the conserved<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_224.html [4/5/2004 5:07:34 PM]


Document<br />

protein kinase core is essential. Yet to predict the region of the structure of the inhibitor that will form<br />

these strong Watson-Crick hydrogen bonds remains difficult.<br />

Page 225<br />

The use of the 6-amino group and N1 nitrogen of the purine base to model the pharmacophore, as seen<br />

with staurosporine, is not possible in two other adenine-type inhibitors. Both isopentenyl adenine, a<br />

nonspecific inhibitor of protein kinases, and olomoucin, a more specific inhibitor of Ser/Thr protein<br />

kinases, are modified only at the 6-amino group position. Thus, bidentate hydrogen-bond formation as<br />

seen in the ATP purine base is not possible. Furthermore, there are inhibitors that do not contain the<br />

chemical structure of adenine, for example des-chloro-flavopyridol, a potent inhibitor of cdc-2 cell cycle<br />

kinase.<br />

In crystallographic analysis of the binding of three inhibitors, olomoucin (OLO), isopentenyl adenine<br />

(ISO) and des-chloro-flavopyridol (DFP) to inactive CDK-2 cell cycle protein kinase, Kim and coworkers<br />

[26] have provided additional insight into the binding of adenine-and nonadenine-<strong>based</strong><br />

inhibitors. Inhibitors with purine rings (OLO and ISO) bind in relatively the same area of the binding<br />

cleft as the adenine ring of ATP. Relative orientation of each purine ring with respect to the protein is<br />

different for all three ligands. This is most likely due to the fact that the 6-amino group of adenine in<br />

ATP is replaced <strong>by</strong> an isopentenylamino group in ISO and <strong>by</strong> the bulky benzylamino group in OLO. In<br />

the case of the third inhibitor, which is not an adenine derivative, the benzopyran ring occupies<br />

approximately the same region as the purine ring of ATP. The two ring systems overlap in the same<br />

plane but benzopyran is rotated about 60 degrees relative to the adenine of ATP. In this orientation, two<br />

strong bidendate hydrogen bond are formed with the oxygens in the 4th and 5th positions of the<br />

inhibitor. Furthermore, these bonds are the same ones formed <strong>by</strong> the 6-amino group and N1 nitrogen of<br />

the adenine ring.<br />

Crystallographic analysis has shown that both the substrate and ATP-binding clefts are structurally<br />

conserved yet differ in the surface charges between individual protein kinases. The structural template of<br />

the protein kinase family as discovered in the structure solution of cAPK predicts these differences.<br />

Template modeling provides a rational basis for the design of specific inhibitors for protein kinases<br />

<strong>based</strong> on the ATP binding site [25]. This is the first significant step in the design of specific inhibitors<br />

targeted at the ATP site.<br />

Recent work has now shown the conformational diversity of inhibitors binding in the interdomain ATPbinding<br />

cleft [26]. Although the residues of the protein kinase catalytic core that form the bidentate<br />

donor-acceptor bond with inhibitors are identical throughout different structures, the residues of the<br />

inhibitors vary greatly. All inhibitors use this common bidentate bond yet the specificity lies in several<br />

other bonds formed between the inhibitor and specific regions of the individual protein kinases.<br />

Furthermore, it is difficult to model<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_225.html [4/5/2004 5:07:35 PM]


Document<br />

Page 226<br />

the bidentate bond-forming residues of the inhibitor and it is through protein crystallography that these<br />

are determined. Template modeling can then be used to determine the specific surface charges of the<br />

modeled kinase that can be exploited <strong>by</strong> the inhibitor for its specificity. Modeling and focused<br />

combinatorial chemistry are the tools to achieve the goal of inhibitor specificity.<br />

Acknowledgments<br />

This work was supported <strong>by</strong> CTR grant 4237. We thank Drs. Furet, P. Traxler, and N. Lydon of Ciba for<br />

their drawings presented in Figure 5. We also thank Dr. Nikola P. Pavletich for the coordinates of the<br />

CDK2-CYCLIN complex used in Figure 2.<br />

References<br />

1. De Bondt HL, Rosenblatt J, Jancarik J, Jones HD, Morgan DO, Kim SH. Crystal structure of cyclindependent<br />

kinase2. Nature 1993; 363:595–602.<br />

2. Hu S-H, Parker MW, Lei JY, Wilce MCJ, Benian GM, Kemp BE. Insights into autoregulation from<br />

the crystal structure of twitchin kinase. Nature 1994; 369:581–584.<br />

3. Hubbard SR, Wei L, Ellis L, Hendrickson WA. Crystal structure of the tyrosine kinase domain of the<br />

human insulin receptor. Nature 1994; 372:746–754.<br />

4. Zhang F, Strand A, Robbins D, Cobb MH, Goldsmith EJ. Atomic structure of the MAP kinase ERK2<br />

at 2.3 Å resolution. Nature 1994; 367:704–710.<br />

5. Owen DJ, Noble MEM, Garman EF, Papageorgiou AC, Johnson LN. Two structures of the catalytic<br />

domain of phosphorylase kinase; an active protein kinase complexed with substrate analogue and<br />

product. <strong>Structure</strong> 1995; 3:467–482.<br />

6. Xu R-M, Carmel G, Sweet RM, Kuret J, Cheng X. Crystal structure of casein kinase-1, a phosphatedirected<br />

protein kinase. The EMBO Journal 1995; 14:1015–1023.<br />

7. Jeffrey PD, Russo AA, Polyak K, Gibbs E, Hurwitz J, Massague J, Pavletich NP. Mechanism of CDK<br />

activation revealed <strong>by</strong> the structure of a cyclinA-CDK2 complex. Nature 1995; 376:313–320.<br />

8. Goldberg J, Nairn AC, Kuryian J. Structural basis for the autoinhibition of calcium/calmodulindependent<br />

protein kinase I. Cell 1996; 84:875–887.<br />

9. Knighton DR, Zheng J-H, Ten Eyck LF, Xuong N-H, Taylor SS, Sowadski JM. <strong>Structure</strong> of a peptide<br />

inhibitor bound to the catalytic subunit of cyclic adenosine monophosphate-dependent protein kinase.<br />

Science 1991; 253:414–420.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_226.html (1 of 2) [4/5/2004 5:07:38 PM]


Document<br />

10. Knighton DR, Zheng J-H, Ten Eyck LF, Ashford VA, Xuong N-H, Taylor SS, Sowadski JM. Crystal<br />

structure of the catalytic subunit of cyclic adenosine monophosphate dependent protein kinase. Science<br />

1991; 253:407–414.<br />

11. Zheng J-H, Trafny EA, Kninghton DR, Xuong N-H, Taylor SS, Ten Eyck LF, Sowadski JM. 2.2Å<br />

refined crystal structure of the catalytic subunit of cAMP-dependent protein kinase complexed with<br />

MnATP and a peptide inhibitor. Acta Cryst 1993; D49:362–365.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_226.html (2 of 2) [4/5/2004 5:07:38 PM]


Document<br />

Page 227<br />

12. Zheng J-H, Knighton DR, Ten Eyck LF, Karlsson R, Xuong N-H, Taylor SS, Sowadski JM. Crystal<br />

structure of the catalytic subunit of cAMP-dependent protein kinase complexed with MgATP and<br />

peptide inhibitor. Biochemistry 1993; 32:2154–2161.<br />

13. Sowadski JM, Ellis C, Madhusudan. Detergent binding to unmyristylated protein kinase<br />

A—structural implications for the role of myristate. Journal of Bioenergetics and Biomembranes 1996;<br />

28:7–12.<br />

14. Knighton DR, Bell S, Xuong N-H, Ten Eyck LF, Taylor SS, Sowadski JM. 2.0Å refined crystal<br />

structure of catalytic subunit of cAMP-dependent protein kinase complexed with a peptide inhibitor and<br />

detergent. Acta Cryst 1993; D49:357–361.<br />

15. Karlsson RF, Zheng J-H, Xuong N-H, Taylor SS, Sowadski JM. Crystal structure of the mammalian<br />

catalytic subunit of cAMP-dependent protein kinase and an inhibitor peptide displays an open<br />

conformation. Acta Cryst 1993; D49:381–388.<br />

16. Zheng J-H, Knighton DR, Xuong N-H, Taylor SS, Sowadski JM, Ten Eyck LF. Crystal structures of<br />

the myristylated catalytic subunit of cAMP-dependent protein kinase reveal open and closed<br />

conformations. Proteins Science 1993; 2:1559–1573.<br />

17. Taylor SS, Radzio-Andzelm E. Three protein kinase structures define a common motif. <strong>Structure</strong><br />

1994; 2:345–355.<br />

18. Knighton DR, Pearson RB, Sowadski JM, Means AR, Ten Eyck LF, Taylor SS, Kemp BE.<br />

Structural basis of the intrasteric regulation of myosin light chain kinase. Science 1992; 258:130–135.<br />

19. Senften M, Schenker G, Sowadski JM, Ballmer-Hofer K. Catalytic activity and transformation<br />

potential of v-Src require arginine 385 in the substrate binding pocket. Oncogene 1995; 10:199–203.<br />

20. Karlsson RF, Madhusudan Taylor SS, Sowadski JM. Intermolecular contacts in various crystal<br />

forms related to the open and closed conformational states of the catalytic subunit of cAMP-dependent<br />

protein kinase. Acta Cryst 1994; D50:657–662.<br />

21. Marcote MJ, Knighton DR, Basi G, Sowadski JM, Brambilla P, Draetta G, Taylor SS. A threedimensional<br />

model of the cdc2 protein kinase: identification of cyclin and suc1 binding regions.<br />

Molecular and Cellular Biology 1993; 13:5122–5133.<br />

22. Songyang, Z et al., Catalytic specificity of protein tyrosine kinase is critical for selective signalling.<br />

Nature 1991; 373:536–539.<br />

23. Ho M-F, Bramson HN, Hansen DE, Knowles JR, Kaiser ET. Stereochemical course of the phospho<br />

group transfer catalyzed <strong>by</strong> cAMP-dependent protein kinase. J Am Chem Soc 1988; 110:2680–2681.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_227.html (1 of 2) [4/5/2004 5:07:40 PM]


Document<br />

24. Madhusudan Xuong N-H, Ten Eyck LF, Taylor SS, Sowadski JM. cAMP-dependent protein kinase:<br />

Crystallographic insights into substrate recognition and phosphotransfer. Protein Science 1994;<br />

3:176–187.<br />

25. Furet P, Caravattti G, Priestle J, Sowadski J, Trinks U, Traxler P. Modeling study of protein kinase<br />

inhibitors: Binding mode of staurosporine-origin of the selectivity of CGP 52 411. J Comp Aid Mol<br />

<strong>Design</strong> 1995; 9:465–472.<br />

26. Azevedo WF Jr, Mueller-Diechmann H-J, Schulze-Gahmen U, Worland PJ, Sausville E, Kim S-H.<br />

Proc Natl Acad Sci 1996; In press.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_227.html (2 of 2) [4/5/2004 5:07:40 PM]<br />

27. Kraulis PJ,<br />

MOLSCRIPT—A<br />

program to<br />

produce both<br />

detailed and<br />

schematic plots of<br />

protein structures.<br />

Journal of<br />

Applied<br />

Crystallography<br />

1991;<br />

24:946–950.


http://legacy.netlibrary.com/reader/message.asp?message=811&BookID=12640&FileName=Page_228.html<br />

The requested page could not be found.<br />

Return to previous page<br />

http://legacy.netlibrary.com/reader/message.asp?message=811&BookID=12640&FileName=Page_228.html [4/5/2004 5:07:46 PM]


Document<br />

9<br />

Structural Studies of Aldose Reductase Inhibition<br />

David K. Wilson and Florante A. Quiocho<br />

Baylor College of Medicine, Houston, Texas<br />

J. Mark Petrash<br />

Washington University School of Medicine, St. Louis, Missouri<br />

I. Introduction<br />

Page 229<br />

Aldose reductase (ALR2; EC 1.1.1.21) is an ~36 kDa enzyme that catalyzes the reduction of a wide<br />

range of carbonyl-containing compounds to their corresponding alcohols. It is a member of an extensive<br />

aldo-keto oxidoreductase enzyme family, a collection of structurally similar proteins expressed in both<br />

animals and plants. Most members of the enzyme family possess similarities in molecular mass, pH<br />

optimum, coenzyme dependence, and demonstrate overlapping specificity for many substrates and<br />

inhibitors.<br />

While no essential physiological function has been established for ALR2, extensive experimental<br />

evidence suggests that it plays an important role in the development of diabetic complications affecting<br />

the visual, nervous, and renal systems [1]. The linkage between ALR2 and pathogenesis of diabetic<br />

complications lies in the polyol pathway of glucose metabolism (Figure 1). In hyperglycemic tissues<br />

such as in diabetes mellitus, the capacity of hexokinase to shunt glucose to glycolysis and other major<br />

pathways of glucose metabolism is exceeded. Consequently, enhanced flux of glucose through the<br />

polyol pathway occurs. The enzyme ALR2 catalyzes the first step in this pathway, producing sorbitol, an<br />

active osmolyte. The polyol pathway is completed <strong>by</strong> the NAD +-dependent oxidation of sorbitol to<br />

fructose, mediated <strong>by</strong> sorbitol dehydrogenase.<br />

Extensive evidence exists to suggest a linkage between the pathogenesis of diabetic complications and<br />

enhanced glucose metabolism via the polyol pathway. The polyol pathway functions in all tissues<br />

susceptible to clinically<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_229.html [4/5/2004 5:07:49 PM]


Document<br />

Figure 1<br />

Schematic of the polyol pathway showing the NADPH-dependent reduction of open<br />

chain D-glucose to sorbitol, which is catalyzed <strong>by</strong> ALR2. This step is followed <strong>by</strong> the<br />

NAD + -dependent oxidation of sorbitol <strong>by</strong> sorbitol dehydrogenase to yield D-fructose.<br />

Page 230<br />

significant diabetic complications. Transgenic animals overexpressing ALR2 in target tissues of diabetic<br />

complications are more prone to development of experimentally induced diabetic complications [2,3].<br />

The most extensive body of evidence linking ALR2 to the pathogenesis of diabetic complications comes<br />

from numerous successes in the treatment of experimental animals with a variety of ALR2 inhibitors<br />

(ARI) [4]. Many of these studies demonstrated that ARIs substantially delay or in some cases prevent<br />

the onset of complications.<br />

Clinical trials of ARIs have yielded encouraging results in alleviating painful symptoms of diabetic<br />

complications. However, unacceptable side effects related to toxicity or inadequate pharmocokinetic<br />

profiles have rendered most of the drug candidates undesirable. Nevertheless, several ARIs are<br />

commercially available in some countries and more appear to be in the pipeline. The therapeutic<br />

rationale for treatment of human diabetics with ARIs to delay or prevent onset of diabetic complications<br />

is compelling. Animal models with experimentally induced hyperglycemia develop complications that<br />

are morphologically and functionally similar to that seen in the human diabetic patient. Many<br />

structurally<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_230.html (1 of 2) [4/5/2004 5:07:56 PM]


Document<br />

diverse ARIs have been shown to substantially delay or completely prevent the onset of such<br />

complications in experimental animal models. While some studies indicate that ALR2 may play a<br />

functional role in osmotic homeostasis in the kidney, evidence from animal studies suggests that it is<br />

metabolically dispensible.<br />

Page 231<br />

Long-term complications exact a terrible toll of morbidity and mortality on patients with diabetes<br />

mellitus. For example, patients with diabetes have about a 25-fold increased risk for becoming blind<br />

over that of the general population. Diabetic retinopathy is one of the most common causes of visual loss<br />

and accounts for about 12% of new cases of blindness each year in the United States alone [5].<br />

II. <strong>Drug</strong> <strong>Design</strong> Prior to Structural Data<br />

Many inhibitors have been developed over the past two decades without the advantage of a structural<br />

understanding of the enzyme [4,6,7]. Significant improvement has been made since the discovery of the<br />

first such orally active compound to show in vivo activity, alrestatin (Figure 2), which had an IC 50 in the<br />

low micromolar range [8]. Many high-affinity inhibitors with IC 50s in the low nanomolar range are now<br />

under study.<br />

Recent drug-design efforts have yielded compounds usually with one of two chemical motifs:<br />

carboxylates or spirohydantoins. A number of these compounds such as tolrestat [9], ponalrestat [10],<br />

epalrestat [11], sorbinil [12], and zopolrestat [13] have progressed to the point of clinical trials.<br />

Unfortunately, clinical ineffectiveness and/or unacceptable side effects have limited the usefulness of<br />

most of those that had been shown to be effective in vitro. The latter problems may be associated with a<br />

lack of specificity since many aldose reductase inhibitors inhibit both ALR2 and aldehyde reductase<br />

[14]. For this reason, ALR2 as well as the other members of the aldo-keto reductase family have been<br />

the subject of crystallographic studies with the hope of determining a structural basis for inhibitor<br />

specificity and ultimately to provide a basis for enhancing binding affinity.<br />

III. Structural Studies of Aldose Reductase<br />

The first crystal structures available for ALR2 were those of the porcine form complexed with the<br />

NADPH analog 2'-monophosphoadenosine-5'-diphosphoribose [15] and the human enzyme complexed<br />

with the NADPH cofactor [16]. Further studies have been conducted on mutants of the human enzyme<br />

[17] and ternary complexes of the human enzyme with an inhibitor [18]. All of these<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_231.html [4/5/2004 5:08:00 PM]


Document<br />

Figure 2<br />

A number of ALR2 inhibitors that have entered clinical trails.<br />

structures show the protein to fold into a (β/α) 8 barrel (Figure 3). This fold has emerged as the most<br />

common enzyme motif [19] although most of the proteins adopting this structure share no sequence<br />

homology. The ALR2 enzyme is, however, the first NAD(P)H binding protein to adopt this fold. It<br />

contains an extra β hairpin preceding the first β strand, which caps the N-terminal end of the barrel. It<br />

also has two helices that are not part of the regular barrel. One precedes α7 and the other follows α8.<br />

A. Cofactor Binding<br />

The NADPH cofactor is bound in an extended conformation across the C-terminal end of the β barrel.<br />

The catalytically active nicotinamide moiety is located<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_232.html (1 of 2) [4/5/2004 5:08:09 PM]<br />

Page 232


Document<br />

Figure 3<br />

C α trace of the ALR2 holoenzyme looking down the (β/α) 8 barrel. The NADPH<br />

cofactor is seen bound across the carboxy-terminal end of the β-barrel with the active<br />

nicotinamide moiety in the center. Figure produced using the MOLSCRIPT<br />

program [48].<br />

Page 233<br />

at the center of the barrel while the adenosine extends away to bind between α7 and α8. A belt<br />

composed of residues 213 to 227 folds over the pyrophosphate of the NADPH to sequester a large part<br />

of the cofactor from the solvent. It is fastened to the other side of the NADPH binding site via Asp216<br />

on the loop that forms bifurcated salt links with Lys21 and Lys262. The dominant interactions holding<br />

the coenzyme in place are directional hydrogen bonds and salt links from positively charged side chains<br />

to the phosphates. The interaction between the 2' phosphate on the NADPH and the side chains from<br />

Lys262 and Arg268 account for the enzyme's preference for NADPH over NADH. Earlier biochemical<br />

studies had shown that it is the 4-pro-R hydride that is transferred from the nicotinamide to the substrate<br />

[20]. This is ensured <strong>by</strong> a hydrogen-bonding network using side chains from Ser159 and Asn160 and the<br />

main chain of Gln183 to orient the amide. It is also determined <strong>by</strong> the stacking interactions with Tyr209,<br />

which is adjacent to the 4-pro-S side of the nicotinamide.<br />

B. Mechanism<br />

The catalytic site was unambiguously identified using the location of the nicotinamide moiety of the<br />

NADPH cofactor in the holoenzyme structure. The region surrounding the catalytic site is a 12-Å deep<br />

groove that measures<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_233.html (1 of 2) [4/5/2004 5:08:15 PM]


Document<br />

Page 234<br />

approximately 7 <strong>by</strong> 13 Å and is lined primarily with hydrophobic side chains. This is entirely consistent<br />

with earlier experiments showing that the enzyme has a marked preference for lipophilic substrates<br />

versus polar substrates such as sugars [21].<br />

The catalytic site also suggested a model for the enzyme's chemical mechanism, which is substantially<br />

similar to most other NAD(P)H-dependent oxido reductases. Upon binding of the reduced cofactor, the<br />

enzyme is able to form a ternary complex with the substrate. The pro-R hydride from the C-4 of the<br />

nicotinamide is transferred to the carbonyl carbon of the substrate, which in turn causes the carbonyl<br />

oxygen to abstract a proton from a general acid, which is presumably located on the protein, to form the<br />

alcoholic product. Three proton-donating side chains are located within 6 Å of the C-4 atom in the<br />

NADPH cofactor that could potentially fulfill this role: Tyr48, His110, and Cys298. Since it is not<br />

conserved in other members of the aldo-keto reductase family that exhibit enzymatic activity (Figure 8)<br />

Cys298 was unlikely as a candidate proton donor. The histidine is surrounded <strong>by</strong> several hydrophobic<br />

residues including Val47, Trp79, and Trp111, which would serve to lower the pK a of the side chain,<br />

making it less effective as a proton donor at physiological pHs. The tryrosine, which ordinarily has a pK a<br />

of approximately 11 engages in an interaction with the charged ammonium group of Lys77, which in<br />

turn charge-pairs with Asp43. This network serves to depress the pK a of the phenolic oxygen, increasing<br />

the exchangability of the proton.<br />

Subsequent activity studies involving site-directed mutants support this model [17,22]. The Tyr48<br />

rarrow.gif Phe mutation shows a complete lack of activity while the Asp43 rarrow.gif Asn, Lys77<br />

rarrow.gif Met, His110 rarrow.gif Asn, and Cys298 rarrow.gif Ser showed losses in catalytic<br />

efficiency of approximately 100-, 1000-, 10 6-, and 10-fold respectively when compared with the wildtype<br />

enzyme. These results correlate well with the functions predicted for each residue with the<br />

exception of the histidine.<br />

The structure of the ALR2 holoenzyme showed that the catalytic site was situated atop the nicotinamide<br />

moiety of the NADPH cofactor. The substrate binding site, which would determine the enzyme's<br />

specificity and also presumably bind inhibitors, appeared to be composed of a deep cleft (Figures 3 and<br />

4). It extended away from the catalytic site towards the loop composed of residues between β4 and α4<br />

and the last 20 residues of the carboxy-terminal meander. This hypothesis was supported <strong>by</strong> the<br />

appearance of poorly resolved density that occupied this region, which suggested the presence of an<br />

endogenously bound substrate or inhibitor in the structure of the holoenzyme [16]. Subsequent studies<br />

indicate that this electron density may be a citrate molecule, one of the components included in the<br />

crystallization mixture. Activity studies indicate that citrate is indeed one of the many inhibitors of the<br />

enzyme with a K i in the millimolar range [23].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_234.html [4/5/2004 5:08:18 PM]


Document<br />

Figure 4<br />

Surface representation of the ALR2 holoenzyme in an orientation similar to Figure 3.<br />

The nicotinamide moiety that defines the active site of the enzyme is seen in the center.<br />

The groove that extends down from it is highly hydrophobic and was initially assumed<br />

to be the inhibitor binding site. Figure prepared using the GRASP program [49].<br />

IV. Aldose Reductase Complexed with Inhibitor<br />

Page 235<br />

While a large number of high-potency inhibitors for ALR2 have been developed [4], a structural<br />

understanding of the exact molecular features that foster this affinity have been only vaguely<br />

understood. Several general chemical motifs such as hydrophobic ring systems, a spirohydrantoin group<br />

or carboxylate group are seen repeatedly when examining a list of known inhibitors (Figure 2) but little<br />

was known about the specific role for each in inhibitor binding.<br />

The structure of the ALR2/NADPH/zopolrestat ternary complex [18] has provided some answers about<br />

the mode of binding of zopolrestat (Figure 2), a high-affinity, carboxylate-containing compound<br />

developed <strong>by</strong> Pfizer, Inc. [13]. While the overall structure was preserved, the inhibitor binding induced a<br />

conformational change of the enzyme. This change, which involved the movement of several loops in<br />

the active site of the molecule, was large enough to cause a change in crystal packing relative to the<br />

holoenzyme. As a consequence of the shifting of the active-site loops, a cavity is created inside the<br />

protein in which the benzothiazole ring is seated and the groove that was implicated in substrate and<br />

inhibitor binding <strong>by</strong> the holoenzyme structure vanishes. This illustrates the unpredictability of<br />

conformational changes within a protein in response to substrate or inhibitor binding. It also implies that<br />

modeling compounds<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_235.html [4/5/2004 5:08:24 PM]


Document<br />

Page 236<br />

in the active site of an enzyme or drug-design algorithms such as inhibitor docking may be inappropriate<br />

in some cases if they do not adequately model considerable plasticity in the binding site.<br />

A closer look at the binding of zopolrestat shows that it is dominated <strong>by</strong> extensive hydrophobic contacts<br />

between the protein and the inhibitor. These include side chains from Trp20, Tyr48, Trp79, Trp111,<br />

Phe115, Phe122, Trp219, Ala299, Leu300, Tyr309, and Pro310 (Figure 7). This is not surprising given<br />

the apolar nature of the enzyme's active site as determined <strong>by</strong> the holoenzyme structure. What is<br />

surprising is that the inhibitor created the part of its own binding site that the benzothiazole rings occupy<br />

<strong>by</strong> “burrowing” into the hydrophobic core of the protein to carve out a region with very good steric<br />

complementarity to this moiety. It does this rather than binding in the solvent-exposed hydrophobic<br />

binding groove that is seen in the holoenzyme structure.<br />

The remaining interactions involving hydrogen bonds and salt links also appear to be very important in<br />

inhibitor binding. With the exception of one of the fluorine atoms, all atoms that are able to engage in<br />

hydrogen bonding do so. The carboxylate, which is seen in so many aldose reductase inhibitors, is saltlinked<br />

to His110, which is located very near the catalytic site (Figure 7). Presumably, the carboxylate in<br />

the other inhibitors plays the same role and could be used as an anchor when modeling these into the<br />

active site.<br />

Inhibition studies involving ALR2 have indicated noncompetitive inhibition for virtually all compounds<br />

examined to date when the forward (reduction) reaction is monitored. This mode of inhibition is often<br />

interpreted as meaning that the inhibitor binds to a site on the enzyme that is independent of the catalytic<br />

site. Kinetic and competition studies have both led to this conclusion in the case of ALR2 [24,25]. The<br />

crystal structure of the enzyme complexed with both the NADPH cofactor and zopolrestat, however,<br />

clearly shows the inhibitor occupying the region directly above the nicotinamide of the NADPH and,<br />

therefore, the active site (Figures 5, 6, and 7).<br />

Most previous inhibition studies reported noncompetitive and/or uncompetitive inhibition patterns when<br />

aldose reductase inhibitors were examined in the forward direction, i.e. inhibition of NADPH-dependent<br />

aldehyde reduction. With the finding that the overall rate-limiting step in the direction of aldehyde<br />

reduction is at the level of structural isomerization following alcohol product release [26,27], it is not<br />

surprising that lack of competitive inhibition would be observed in such standard double reciprocal<br />

plots. To further complicate matters, many aldose reductase inhibitors were not recognized in previous<br />

studies as tight-binding inhibitors and were inappropriately evaluated using Michaelis-Menten kinetics.<br />

Thus, noncompetitive or uncompetitive inhibition patterns were previously reported for inhibitors that<br />

were subsequently shown to bind directly at the active site. Recent structure-function and kinetic studies<br />

have revealed important details concerning the structural basis for the catalytic<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_236.html [4/5/2004 5:08:26 PM]


Document<br />

Figure 5<br />

Schematic representation of the ALR2/NADPH/zopolrestat ternary complex. The<br />

NADPH is bound across the enzyme from the center to the right while the zopolrestat<br />

binds atop the nicotinamide and extends to the lower left. Conformational changes are<br />

seen in the C-terminal loop below the zopolrestat in the picture and the loop to the right<br />

of the inhibitor between β4 and α4.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_237.html (1 of 2) [4/5/2004 5:08:37 PM]<br />

Page 237


Document<br />

Figure 6<br />

A surface representation of the ternary complex as seen in Figure 5. Note that the<br />

inhbitor creates part of its binding site <strong>by</strong> “burrowing” into the protein rather than<br />

binding entirely in the groove seen Figure 4.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_237.html (2 of 2) [4/5/2004 5:08:37 PM]


Document<br />

Figure 7<br />

Stereo of zopolrestat binding to the active site of ALR2. The salt link made <strong>by</strong> the carboxylate of the inhibitor and<br />

hydrogen bonds are depicted with dashed lines. The remainder of the interactions are apolar with the residues shown.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_238.html [4/5/2004 5:08:48 PM]<br />

Page 238


Document<br />

and inhibition mechanisms and provided a clarification for the mechanism <strong>by</strong> which inhibitors of both<br />

the carboxylate [28] and spirohydantoin [29] classes bind at the active site of ALR2.<br />

V. Aldo-Keto Reductase Family<br />

A. Effects of Sequence on <strong>Drug</strong> <strong>Design</strong><br />

Page 239<br />

The problem of structure-<strong>based</strong> drug design for ALR2 and the drug-design effort in general is<br />

compounded <strong>by</strong> the fact that this enzymes is a member of a large family of aldo-keto reductases with<br />

overlapping substrate specificity. In humans at least three such enzymes have been found: ALR2 [30],<br />

aldehyde reductase (ALR1) [30], and chlordecone reductase [31]. Other members of the family that have<br />

been isolated in other species include rat 3α-hydroxysteroid dehydrogenase [32], murine fibroblast<br />

growth factor induced protein [33], bovine prostagladin F synthase [34], murine vas deferens protein<br />

[35], frog ρ-crystallin [36], the P100/11E gene product in Leishmania major [37], and Corynebactium<br />

diketogluconate reductase [38]. This large number suggests that there may be more such enzymes to be<br />

found in humans. The similarities between the proteins with respect to both the sequence and substrate<br />

specificity implies that the nature of the substrate binding sites are similar across the family. This has<br />

indeed been the case in all the structures determined from this family to date (see below).<br />

While detailed binding studies of various inhibitors with all the different enzymes have not been<br />

conducted, it is likely that drugs intended for ALR2 are likely to “cross react” with many of the other<br />

enzymes within the family. One such case that has recently been studied both crystallographically and<br />

biochemically is the murine FR-1 protein described below.<br />

The binding sites of all of these enzymes are characterized <strong>by</strong> their large size and hydrophobicity<br />

suggesting that ideal substrates may be steroids or molecules of a similar size and nature. Sequence<br />

comparisons of all the proteins, including those whose structure has not yet been determined, show that<br />

there is a large amount of similarity involving residues implicated in substrate binding (see Figure 8).<br />

One region that diverges somewhat is the 15-amino acid segment at the carboxy terminus of the protein.<br />

This segment is likely to be responsible for what little differences in substrate specificity exhibited <strong>by</strong><br />

the enzymes. It is the same segment that is seen adopting a different conformation upon zopolrestat<br />

binding to ALR2. It may then be possible that it is not only the chemical nature of this loop that—in<br />

making positive and negative interactions with the substrate/inhibitor—modulates specificity, but also<br />

the flexibility conferred <strong>by</strong> the amino acid sequence. Such a difference is seen when contrasting the<br />

structures of ALR2 and FR-1 bound to zopolrestat.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_239.html [4/5/2004 5:09:05 PM]


Document<br />

B. <strong>Structure</strong>s<br />

Figure 8<br />

Sequence alignment of several members of the aldo-keto reductase family. Abbreviations used<br />

are HALR2, human aldose reductase (30); HALR1, human aldehyde reductase [30]; 3α-HSD,<br />

3α-hydroxysteroid dehydrogenase [32]; FR-1, murine FR-1 [33]; BPGFS, bovine<br />

prostaglandin F synthase [34]; CCDR12, human chlordecone reductase [31]; CDGR,<br />

Corynebacterium diketogluconate reductase [38]; MVDP, murine vas deferens protein [35],<br />

JFRC, Japanese frog ρ crystallin [36].<br />

Page 240<br />

<strong>Structure</strong>s of several other members of the aldo-keto reductase family have also been determined. These<br />

include aldehyde reductase [39,40], FR-1 [41] and 3α-hydroxysteroid dehydrogenase [42]. Since each of<br />

these proteins retain a large amount of sequence identity and homology with human ALR2, it is not<br />

surprising to note that the overall tertiary structures are very similar. Root-meansquare C α deviations<br />

between the human ALR2 holoenzyme structure and the rest of the family are in the range of 1–2Å<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_240.html (1 of 2) [4/5/2004 5:09:22 PM]


Document<br />

Page 241<br />

Closer examination of the active site shows that the residues involved in catalysis (most notably the<br />

residues analogous to Asp43, and Tyr48, and Lys77 in ALR2) are structurally conserved among each of<br />

these proteins (Figure 8), suggesting that the mechanism is also conserved throughout the family. Many<br />

of the residues found in the binding site (defined as those making contact with the zopolrestat in the<br />

ternary complex) are also largely conserved with the exception of a number of residues in the carboxy<br />

terminus of the protein. This is indeed where the most structure variation appears to be concentrated<br />

among the proteins. These residues compose a loop that is the same loop that shifts upon binding of<br />

zopolrestat in ALR2. Inhibitors with improved specificity will very likely take advantage of the subtle<br />

structural differences that are introduced <strong>by</strong> the variation in sequences in this area.<br />

VI. Future <strong>Design</strong> of Aldose Reductase Inhibitors<br />

The availability of structural data for ALR2 in its holoenzyme and different ternary forms is likely to<br />

lead to improvements in the affinity of future generations of inhibitors. As the architecture and plasticity<br />

of the binding site are better understood, increasingly potent inhibitors may be designed to occupy it.<br />

Although these structures provide a positive target for drug design, there are a number of negative<br />

targets. Increased in vivo potency is likely to be derived from the specific inhibition of ALR2 that would<br />

entail the avoidance of other members of the aldo-keto reductase family. Determination of the structures<br />

of other members of the family may increase the specificity of compounds <strong>by</strong> providing structures of<br />

targets to avoid. While the incorporation of the negative targets in the drug-design process relies on the<br />

determination of other structures and is likely to be complicated, conventional computational techniques<br />

may be applied to the problem of the positive target. Two such methods that may hold promise are<br />

docking [43] and computational thermodynamic perturbation [44].<br />

A. Inhibitor Docking to the Enzyme<br />

Our initial efforts to exploit the ALR2 holoenzyme structure for drug design utilized the program DOCK<br />

[43]. This program is capable of finding depressions on the surface of the enzyme that could serve as<br />

binding sites for substrates or inhibitors. Once the correct area is defined, the program rotates structures<br />

of candidate compounds within this space and scores each compound <strong>based</strong> upon its steric<br />

complementarity with the binding site. However, the program does not include the potential polar<br />

interactions between the inhibitor and protein when scoring. The search was further constrained <strong>by</strong> the<br />

inability to include conformational variations both in the test compounds as well as the protein, due to<br />

computational limitations.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_241.html [4/5/2004 5:09:24 PM]


Document<br />

This method was used to screen a significant portion of the contents of the Cambridge Structural<br />

Database [45] against the ALR2 holoenzyme binding site (D.K. Wilson, J. M. Petrash, and F. A.<br />

Quiocho, unpublished data). Among the approximately 30 highest scoring compounds were several<br />

aromatic aldoximes that had inhibition constants in the micromolar range. These were similar to<br />

aldoximes such as benzaldoxime, which has been previously observed to have similar inhibition<br />

constants [46].<br />

Page 242<br />

A disappointing result was that this search did not “rediscover” any of the known high-affinity ALR2<br />

inhibitors that were contained in the search library. Before the determination of the ternary complex of<br />

the enzyme with zopolrestat, this was interpreted as meaning that these compounds bound to the enzyme<br />

in a conformation somewhat different than the one adopted in the crystal structure used for the search.<br />

The structure of the ternary complex showed this to be a wrong assumption; it was the protein that<br />

changed conformation upon inhibitor binding, creating a pocket that did not exist in the holoenzyme<br />

structure. When bound to the protein, zopolrestat is actually quite similar in conformation to its small<br />

molecule x-ray structure. It is therefore very possible that different ALR2 inhibitors and substrates may<br />

cause the enzyme to flex in different ways, creating binding sites that may be different in size and<br />

chemical nature.<br />

B. Computational Thermodynamic Perturbation<br />

Computational thermodynamic perturbation is a powerful, albeit computationally expensive, group of<br />

techniques that are designed to estimate relative binding affinities of two closely related drugs, given the<br />

structure of at least one of them complexed with the target protein [44]. This approach has the potential<br />

to assay candidate compounds in the computer for improvements in inhibitor binding, there<strong>by</strong> removing<br />

the necessity to sythesize and assay these compounds in the lab.<br />

For a number of reasons, ALR2 promises to be a good system for the application of this technique and<br />

the experimental verification of the results. The structure is very well determined in complex with<br />

zopolrestat, a high-affinity inhibitor. A number of zopolrestat derivatives with various functional groups<br />

decorating the compound have been sythesized and partially characterized with respect to ALR2<br />

inhibition [13,47]. These compounds could serve as a sort of basis set of controls for the theoretical<br />

calculations. If parameters used in these calculations can be selected such that the computationally<br />

derived binding energies agree even qualitatively with the experimentally determined binding energies,<br />

serious consideration should be given to new compounds that are predicted to bind with enhanced<br />

affinity. Since ALR2 is crystallizable with zopolrestat bound, there is every reason to believe that<br />

crystals of the enzyme complexed with similar compounds will be obtainable. Such structures could<br />

provide the basis for further rounds of drug improvement.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_242.html [4/5/2004 5:09:26 PM]


Document<br />

VII. Conclusions<br />

Page 243<br />

The observation of a protein undergoing a conformational change when binding to an inhibitor, as seen<br />

with ALR2 and zopolrestat, illustrates a common problem associated with structure-<strong>based</strong> drug design.<br />

It is tempting to view proteins as static structures since their crystal structures are static. Attempts to<br />

design drugs to fit the apparent active site of an enzyme may fail when the plasticity of the protein is not<br />

taken into account. While the disorder associated with amino acid side chains can be modeled with a<br />

moderate computational effort, larger conformational changes—such as the loop movement seen in<br />

ALR2—are virtually impossible to predict. Until this becomes possible, x-ray crystal structures of<br />

complexes will continue to be indispensible.<br />

Finally, it can be easy to forget that a compound's affinity for the protein is not the only consideration<br />

when designing inhibitors of enzymes from a structural point of view. The structures of aldose reductase<br />

and the FR-1 protein complexed with the drug zopolrestat, a compound with a very high affinity for<br />

ALR2, can serve as a reminder of how specificity can also be a very important factor. This is<br />

particularly true when a protein is a member of a family of proteins that share sequence homology and<br />

are apt to have overlapping specificities. <strong>Structure</strong> may then play a key role in the determination of<br />

features that are unique to the target protein and therefore prime considerations when designing<br />

inhibitors.<br />

Acknowledgments<br />

We thank T. Reynolds who assisted with the production of the figures. This work was supported <strong>by</strong> a<br />

grant from Research to Prevent Blindness, Inc. and grants EY05856, EY02687, and DK20579 to J. Mark<br />

Petrash. Florante A. Quiocho is an investigator of the Howard Hughes Medical Institute.<br />

References<br />

1. Kinoshita JH, Nishimura C. The involvement of aldose reductase in diabetic complications. Diabetes-<br />

Metebolism Rev 1988; 4:323–337.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_243.html (1 of 2) [4/5/2004 5:09:27 PM]


Document<br />

2. Lee AYW, Chung<br />

SK, Chung SSM.<br />

Demonstration that<br />

polyol accumulation<br />

is responsible for<br />

diabetic cataract <strong>by</strong><br />

the use of transgenic<br />

mice expressing the<br />

aldose reductase<br />

gene in the lens.<br />

Proc Natl Acad Sci<br />

USA<br />

1995;92:2780–2784.<br />

3. Yamaoka T, Nishimura C, Yamashita K, Itakura M, Yamada T, Fujimoto J, Kokai Y. Acute onset of<br />

diabetic pathological changes in transgenic mice with human aldose reductase cDNA. Diabetologia<br />

1995;38:255–261.<br />

4. Sarges R, Oates P. Aldose reductase inhibitors: Recent developments. Prog <strong>Drug</strong> Res 1993;<br />

40:99–161.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_243.html (2 of 2) [4/5/2004 5:09:27 PM]


Document<br />

5. National Advisory Eye Council (1994). In: Vision Research: A National Plan 1994–1998. National<br />

Institutes of Health Publication No. 93–3186.<br />

6. Kador PF. The role of aldose reductase in the development of diabetic complications. Med Res Rev<br />

1988; 8:325–352.<br />

Page 244<br />

7. Dvornik D. Aldose Reductase Inhibition: An Approach to the Prevention of Diabetic Complications.<br />

New York: Biomedical Information Corporation 1987.<br />

8. Dvornik D, Simard-Duquesne N, Krami M, Sestanj K, Gabbay KH, Kinoshita JN, Varma SD, Merola<br />

LO. Polyol accumulation in galatosemic and diabetic rats: control <strong>by</strong> an aldose reductase inhibitor.<br />

Science 1973; 182:1146–1148.<br />

9. Sestanj K, Bellini F, Fung S, Abraham N, Treasurywala A, Humber L, Simard-Duquesne N, Dvornik<br />

D. N-[[5-(trifluoromethyl)-6-methoxy-1-naphthalenyl]thioxomethyl]-N-methylglycine (Tolrestat), a<br />

potent, orally active aldose reductase inhibitor. J Med Chem 1984;27:255–256.<br />

10. Ward WHJ, Sennitt CM, Ross H, Dingle A, Timms D, Mirrlees DJ, Tuffin DP. Ponalrestat: a potent<br />

and specific inhibitor of aldose reductase. Biochem Pharmacol 1990; 39:337–346.<br />

11. Terashima H, Hama K, Yamamoto R, Tsuboshima M, Kikkawa R, Hatanaka 1, Shigeta Y. Effects of<br />

a new aldose reductase inhibitor on various tissues in vitro. J Pharmacol Exp. Ther 1984;229:226–230.<br />

12. Peterson MJ, Sarges R, Aldinger CD. CP-45634: a novel aldose reductase inhibitor that inhibits<br />

polyol pathway activity in diabetic and galactosemic rats. Metabolism 1979; 28(supp 1):456–461.<br />

13. Mylari BL, Larson ER, Beyer TA, Zembrowski WJ, Aldinger CE, Dee MF, Siegel TW, Singleton<br />

DH. Novel, potent aldose reductase inhibitors: 3,4-dihydro-4-oxo-3-[[5-(trifluoromethyl)-2benzothiazolyl]methyl]-1-phthalazine-acetic<br />

acid (zopolrestat) and congeners. J Med Chem 1991;<br />

34:108–122.<br />

14. Srivastava SK, Petrash JM, Sadana AJ, Partridge CA. Susceptibility of aldose and aldehyde<br />

reductases to aldose reductase inhibitors. Curr Eye Res 1982; 2:407–410.<br />

15. Rondeau JM, Tete-Favier F, Podjarny A, Reymann JM, Barth P, Biellmann JF, Moras D. Novel<br />

NADPH-binding domain revealed <strong>by</strong> the crystal structure of aldose reductase. Nature 1992;<br />

355:469–472.<br />

16. Wilson DK, Bohren KM, Gabbay KH, Quiocho FA. An unlikely sugar substrate site in the 1.65 Å<br />

structure of the human aldose reductase holoenzyme implicated in diabetic complications. Science 1992;<br />

257:81–84.<br />

17. Borhani DW, Harter TM, Petrash JM. The crystal structure of the aldose reductase-NADPH binary<br />

complex. J Biol Chem 1992; 267:24841–24847.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_244.html (1 of 2) [4/5/2004 5:09:29 PM]


Document<br />

18. Wilson DK, Tarle I, Petrash JM, Quiocho FA. Refined 1.8 Å structure of human aldose reductase<br />

complexed with the potent inhibitor zopolrestat. Proc Natl Acad Sci USA 1993; 90:9847–9851.<br />

19. Branden CI. The TIM barrel—the most frequently occurring folding motif in proteins. Curr Opin<br />

Struct Biol 1991; 1:978–983.<br />

20. Feldman HB, Szczepanik PA, Havre P, Corrall RJM, Yu LC, Rodman HM, Rosner BA, Klein PD,<br />

Landau, BR. Stereospecificity of the hydrogen transfer catalyzd <strong>by</strong> human placental aldose reductase.<br />

Biochim Biophys Acta 1997; 480:14–20.<br />

21. Wermuth B, Buergisser HB, Bohren KM, von Wartburg JP. Purification and characterization of<br />

human brain aldose reductase. Eur J Biochem 1982; 127:279–284.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_244.html (2 of 2) [4/5/2004 5:09:29 PM]


Document<br />

Page 245<br />

22. Tarle I, Borhani DW, Wilson DK, Quiocho FA, Petrash JM. Probing the active site of human aldose<br />

reductase: site directed mutagenesis of Asp-43, Lys-77 and His-110. J Biol Chem 1993;<br />

268:25687–25693.<br />

23. Harrison DH, Bohren KM, Ringe D, Petsko GA, Gabbay KH. An anion binding site in human aldose<br />

reductase: mechanistic implications for the binding of citrate, cacodylate, and glucose 6-phosphate.<br />

Biochemistry 1994; 33:2011–2020.<br />

24. Kador PF, Sharpless NE. Pharmacophor requirements of the aldose reductase inhibitor site. Mol<br />

Pharmacol 1983; 24:521–531.<br />

25. Kador PF, Goosey JD, Sharpless NE, Kolish J, Miller DD. Stereospecific inhibition of aldose<br />

reductase. Eur J Med Chem 1981; 16:293–298.<br />

26. Kubiseski TJ, Hyndman DJ, Morjana NA, Flynn TG. Studies on pig muscle aldose reductase.<br />

Kinetic mechanism and evidence for a slow conformational change upon coenzyme binding. J Biol<br />

Chem 1992; 267:6510–6517.<br />

27. Grimshaw CE, Shahbaz M, Putney CG. Mechanistic basis for nonlinear kinetics of aldehyde<br />

reduction catalyzed <strong>by</strong> aldose reductase. Biochemistry 1990; 29:9947–9955.<br />

28. Grimshaw CE, Bohren KM, Lai CJ, Gabbay KH. Human aldose reductase: pK of tyrosine 48 reveals<br />

the preferred ionization state for catalysis and inhibition. Biochemistry 1995; 34:14374–14384.<br />

29. Liu SQ, Bhatnagar A, Srivastava SK. Does sorbinil bind to the substrate binding site of aldose<br />

reductase? Biochem Pharmacol 1992; 44:2427–2429.<br />

30. Bohren KM, Bullock B, Wermuth B, Gabbay KH. The aldo-keto reductase superfamily: cDNAs and<br />

deduced amino acid sequences of human aldehyde and aldose reductases. J Biol Chem 1989;<br />

264:9547–9551.<br />

31. Winters CJ, Molowa DT, Guzelian PS. Isolation and characterization of cloned cDNAs encoding<br />

human liver chlordecone reductase. Biochemistry 1990; 29:1080–1087.<br />

32. Pawlowski JE, Huizinga M, Penning TM. Cloning and sequencing of the cDNA for rat liver 3αhydroxysteroid/dihydrodiol<br />

dehydrogenase. J Biol Chem 1991; 266:8820–8825.<br />

33. Donohue PJ, Alberts GF, Hampton BS, Winkles JA. A delayed-early gene activated <strong>by</strong> fibroblast<br />

growth factor-1 encodes a protein related to aldose reductase. J Biol Chem 1994; 269:8604–8609.<br />

34. Watanabe K, Fujii Y, Nakayama K, Ohkubo H, Kuramitsu S, Kagamiyama H, Nakanishi S,<br />

Hayaishi O. Structural similarity of bovine lung prostaglandin F synthase to lens ε crystallin of the<br />

European common frog. Proc Natl Acad Sci U S A 1988; 85:11–15.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_245.html (1 of 2) [4/5/2004 5:10:06 PM]


Document<br />

35. Pailhoux EA, Martinez A, Veyssiere GM, Jean CG. Androgen-dependent protein from mouse vas<br />

deferens: cDNA cloning and protein homology with the aldo-keto reductase superfamily. J Biol Chem<br />

1990; 265:19932–19936.<br />

36. Fujii Y, Watanabe K, Hayashi H, Urade Y, Kuramitsu S, Kagamiyama H, Hayashi O. Purification<br />

and characterization of p-crystallin from Japanese common bullfrog lens. J Biol Chem 1990;<br />

265:9914–9923.<br />

37. Samaras N, Spithill TW. The developmentally regulated P100/11E gene of Leishmania major shows<br />

homology to a superfamily of reductase genes. J Biol Chem 1989; 264:4251–4254.<br />

38. Anderson S, Marks CM, Lazarus R, Miller J, Stafford K, Seymour J, Light D, Rastetter W, Estell D.<br />

Production of 2-keto-L-gulonate, an intermediate in L-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_245.html (2 of 2) [4/5/2004 5:10:06 PM]


Document<br />

ascorbate synthesis, <strong>by</strong> a genetically modified Erwinia herbicola. Science 1985; 230:144–149.<br />

Page 246<br />

39. El-Kabbani O, Green NC, Lin G, Carson M, Narayanam SVL, Moore K, Flynn TG, DeLucas LJ.<br />

<strong>Structure</strong>s of human and porcine aldehyde reductase: an enzyme implicated in diabetic complications.<br />

Acta Crystallogr D 1994; 50:859–868.<br />

40. El-Kabbani O, Judge K, Ginell SL, Myles Daa, DeLucas LJ, Flynn TG. <strong>Structure</strong> of porcine<br />

aldehyde reductase holoenzyme. Nat Struct Biol 1995; 2:687–692.<br />

41. Wilson DK, Nakano T, Petrash JM, Quiocho FA. 1.7 Å structure of FR-1, a fibroblast growth factorinduced<br />

member of the aldo-keto reductase family complexed with coenzyme and inhibitor.<br />

Biochemistry 1995; 34:14323–14330.<br />

42. Hoog SS, Pawlowski JE, Alzari PM, Penning TM, Lewis M. Three-dimensional structure of rat liver<br />

3α-hydroxysteroid/dihydrodiol dehydrogenase: a member of the aldo-keto reductase superfamily. Proc<br />

Natl Acad Sci USA 1994; 91:2517–2521.<br />

43 Schoichet B, Bodian D, Kuntz I. Molecular docking using shape descriptors. J Comp Chem 1992;<br />

13:380–397.<br />

44 Straatsma TP, McCammon JA. Computational alchemy. Annu Rev Phys Chem 1992 43:407–435.<br />

45 Allen FG, Bellar SA, Brice MD, Cartwright BA, Doubleday A, Higgs H, Hummelink T, Hummelink-<br />

Peters BG, Kennard O, Motherwell WDS, Rodgeres JR, Watson DG. The Cambridge Crystallographic<br />

Data Centre: Computer <strong>based</strong> search retrieval, analysis and display of information. Acta Crystallogr<br />

B35:2331–2339.<br />

46 Shen C, Sigman DS. New inhibitors of aldose reductase: anti-oximes of aromatic aldehydes. Arch<br />

Biochem Biophys 1991; 286:596–603.<br />

47 Mylari BL, Beyer TA, Scott PJ, Aldinger CE, Dee MF, Siegel TW, Zembrowski WJ. Potent, orally<br />

active aldose reductase inhbitors related to zopolrestat: surrogates for benzothiazole side chain. J Med<br />

Chem 1992; 35:457–465.<br />

48 Kraulis PJ. MOLSCRIPT: a program to produce both detailed and schematic plots of protein<br />

structures. J Appl Crystallog 1991; 24:946–950.<br />

49 Nicholls A, Sharp KA, Honig B. Protein folding and association: insights from the interfacial and<br />

thermodynamic properties of hydrocarbons. Proteins 1991; 11:281–296.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_246.html [4/5/2004 5:10:07 PM]


Document<br />

10<br />

<strong>Structure</strong>-Based <strong>Design</strong> of Thrombin Inhibitors<br />

Patricia C. Weber and Michael Czarniecki<br />

Schering-Plough Research Institute, Kenilworth, New Jersey<br />

I. Roles of Thrombin in Hemostasis and the Therapeutic Utility of Thrombin Inhibitors<br />

Page 247<br />

Thrombin is a serine protease that plays critical roles in both blood clot formation and anticoagulation.<br />

In the penultimate step of the coagulation cascade, thrombin cleaves soluble fibrinogen to form<br />

insoluble fibrin. Thrombin also activates other coagulation factors including Factor XIII, the enzyme<br />

responsible for crosslinking fibrin to further stabilize the thrombus. Additional clot-promoting functions<br />

include stimulation of platelet aggregation <strong>by</strong> cleavage of the thrombin receptor. In contrast to its roles<br />

in clot formation, thrombin participates in anticoagulant functions. For example, thrombin-mediated<br />

activation of protein C, a protease involved in anticoagulation, is enhanced when thrombin is complexed<br />

with thrombomodulin, and in this complex, thrombin can neither cleave fibrinogen nor activate platelets.<br />

The interrelationship among thrombin's many roles in hemostasis is complex and presents several<br />

mechanisms for inhibition of thrombus formation. For recent reviews see References 1 through 5.<br />

Most drug discovery efforts focus on thrombin inhibition as a means to prevent the serious<br />

consequences of thrombus formation in myocardial infarction and stroke. Thrombin inhibitors may also<br />

prevent clot formation in patients prone to deep vein thrombosis or repeat heart attack. In combination<br />

with thrombus dissolution therapies, thrombin inhibitors may decrease the incidence of reocclusion due,<br />

in part, to the release of active clot-bound thrombin.<br />

In this article, recent examples of small molecule inhibitors interacting at the fibrinogen primary<br />

specificity pocket and with residues of the catalytic triad<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_247.html [4/5/2004 5:10:09 PM]


Document<br />

are given. Inhibitors designed to make more extended interactions with thrombin are also presented.<br />

II. <strong>Structure</strong> of Thrombin<br />

Page 248<br />

Thrombin consists of two polypeptides, an A chain of 36 residues and a 259-residue B chain, linked <strong>by</strong> a<br />

disulfide bond. The crystallographic structure of thrombin reveals a globular protein organized about<br />

two β barrels with the overall folding pattern of the chymotrypsin serine protease family [6,7]. The<br />

catalytic triad and near<strong>by</strong> oxyanion hole are located roughly between the β barrels and adopt the<br />

geometric arrangement required for serine-protease-assisted, peptide bond cleavage (Figure 1).<br />

Thrombin's multifunctionality and regulation of activity are achieved <strong>by</strong> specialized subsites on the<br />

enzyme's surface (Figure 2). Fibrinogen cleavage, for example, involves interactions at the primary<br />

specificity pocket, the extended fibrinogen recognition exosite, and an additional specificity pocket.<br />

Subsite interactions differ for cleavage of other thrombin substrates including the thrombin receptor and<br />

protein C. Additional and overlapping subsites exist for thrombin effector molecules including heparin,<br />

antithrombin III, and heparin cofactor II [8,9].<br />

Figure 1<br />

Stereoscopic view of the crystallographic structure of thrombin complexed with<br />

N-acetyl-(D-Phe)-Pro-boroArg-OH. Helical regions are represented in the standard<br />

way and arrows indicate regions of β sheet. Solid lines show the thrombin<br />

bound conformation of N-acetyl-(D-Phe)-Pro-boroArg-OH (taken from Reference 10).<br />

Active-site residues, His57 and Ser195, are shown with a ball-and-stick representation.<br />

The authors thank Dr. C. L. Strickland for the drawing.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_248.html [4/5/2004 5:10:15 PM]


Document<br />

Figure 2<br />

Schematic representation of Subsite Utilization in Thrombin Complexes (after<br />

Reference 8). Fibrinogen interacts with three thrombin subsites (here thrombin is<br />

represented <strong>by</strong> a large oval and the interconnected subsites <strong>by</strong> an irregular<br />

three-armed shape). Physiological effectors of thrombin and thrombin inhibitors form<br />

distinct interactions at these subsites. Additional subsites, such as the<br />

heparin-binding site, exist on the thrombin surface and are not indicated here.<br />

The catalytic triad is represented <strong>by</strong> three circles at the vertices of a triangle.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_249.html (1 of 2) [4/5/2004 5:10:27 PM]<br />

Page 249


Document<br />

III. Thrombin Inhibitors Directed at the Fibrinopeptide a Binding Pocket<br />

Page 250<br />

The majority of synthetic thrombin inhibitors interact at the fibrinopeptide A binding pocket, which<br />

includes the catalytic residues Ser195 and His57, hydrogen-bonding capabilities within the oxyanion<br />

hole, peptide backbone functional groups that hydrogen bond with the peptide backbone of the substrate,<br />

and residues involved in amino acid recognition (Figure 3). Many of these binding determinants are<br />

utilized <strong>by</strong> N-acetyl-(D-Phe)-Pro-Arg-chloromethylketone (PPACK [7]) and its boronic acid analog<br />

(DUP714 [10]). The crystallographic structures of these molecules complexed with thrombin have both<br />

served as starting points for structure-<strong>based</strong> drug design and as reference structures for comparison of<br />

binding modes of other inhibitors.<br />

The use of arginine boronate esters as transition-state mimetics results in potent peptidyl thrombin<br />

inhibitors. These inhibitors, however, exhibit significant affinity for other serine proteases that have in<br />

common a specificity for substrates with basic residues at P1 (e.g. trypsin, Factor Xa, and plasmin).<br />

Earlier work demonstrated that neutral side chains of P1 boronate esters impart greater selectivity for<br />

thrombin. The boropeptide shown in Figure 4 was investigated as the prototype of neutral side chain,<br />

tripeptide thrombin inhibitors [11]. It had a K i against thrombin of 7 nM and shows selectivity relative to<br />

other trypsin-like plasma proteases. Since these inhibitors have a neutral residue at the P1 site, Deadman<br />

and coworkers [11] sought to demonstrate the mode of binding to thrombin in the absence of a salt<br />

bridge with Asp189.<br />

Figure 3<br />

Schematic diagram of binding determinants within<br />

the fibrinopeptide A binding pocket of thrombin and<br />

their utilization <strong>by</strong> N-acetyl-(D-Phe)-Pro-boroArg-OH.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_250.html [4/5/2004 5:10:30 PM]


Document<br />

Figure 4<br />

Schematic representation of the active-site orientation of a<br />

“neutral” P1 boronic acid thrombin inhibitor.<br />

Page 251<br />

Boron-11-NMR, a sensitive probe of the chemical environment around boronate esters, can distinguish<br />

between trigonal and tetrahedral forms of boron. The 11B-NMR spectrum of this inhibitor complexed<br />

with thrombin showed a single peak at -17 ppm that remained constant for 7 hours. The chemical shift<br />

suggests boron adopts a tetrahedral geometry on binding to thrombin and is consistent with the<br />

orientation of the inhibitor in the active site shown in Figure 4. While the 11B-NMR revealed an<br />

interaction within the catalytic site, it could not distinguish between bonding with Ser195 or His57.<br />

Kahn and coworkers [12] recently investigated the application of synthesized peptidomimetics as novel<br />

inhibitors of thrombin. Fibrinogen peptide A mimetic (FPAM, Figure 5) incorporates a bicyclic<br />

peptidomimetic within the turn region of fibrinogen peptide A. The bicyclic peptidomimetic confers<br />

conformational stability to the turn region as suggested <strong>by</strong> x-ray crystal structures of fibrinogen peptide<br />

complexes as well as complexes of BPTI with thrombin.<br />

X-ray crystallographic studies of FPAM complexed with thrombin (Figure 5) showed that the S1 subsite<br />

is occupied <strong>by</strong> the arginine guanidinium [12]. The Val group of FPAM makes extensive hydrophobic<br />

contacts within the S2 apolar binding site. The Gly at P3 interacts with thrombin via a β-sheet-type<br />

hydrogen bond with the carbonyl group of Gly216 and appears to be important in the positioning of the<br />

bicyclic ring corresponding to the β bend. This bicyclic ring, although not aromatic, forms an edge-toface<br />

contact with Trp215. One of the phenyl rings shows hydrophobic contact with lle174, while the<br />

other shows no significant interactions with thrombin. By comparison to other inhibitors complexed to<br />

thrombin, FPAM appears to have a new binding mode that differs from that of substrate, or hirudin, or<br />

argatroban.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_251.html (1 of 2) [4/5/2004 5:10:32 PM]


Document<br />

Figure 5<br />

Schematic representation of the principal intermolecular<br />

interactions of a fibrinogen peptide A mimetic within the active site<br />

of thrombin.<br />

Page 252<br />

Obst et al. [13], departing from the peptide template, designed and synthesized novel nonpeptide<br />

inhibitors of thrombin. They began with a cyclic template having attachment sites for three side chains<br />

that would be complementary to the S1, S2, and S3 sites in thrombin. Important to the design was a rigid<br />

template that would avoid hydrophobic collapse of the side chains and loss of conformational degrees of<br />

freedom upon complex formation with thrombin. Using computational approaches (Insight<br />

II/Discover/CVFF force field), possible templates were modeled within the active site of thrombin.<br />

These studies resulted in the synthesis of thirteen analogs that shared a common template.<br />

The most active molecule (K i = 90 nM, 8-fold selective versus trypsin) was studied further <strong>by</strong> x-ray<br />

crystallography (Figure 6). The positively charged benzamidine binds into the S1 pocket of thrombin<br />

forming a bidentate hydrogen bond with Asp189. The proximal carbonyl of the rigid template acts as a<br />

hydrogen-bond acceptor for the amide NH of Gly216. The methylene dioxybenzyl group at P3 interacts<br />

with thrombin in two ways. An edge-to-face interaction was observed with Trp215, and an oxygen of<br />

the methylenedioxy group acts as an acceptor for a hydrogen bond with the OH hydrogen of Tyr60A.<br />

Recent communications from Bristol-Myers Squibb [14,15] describe peptidomimetic inhibitors (Figure<br />

7) that were designed to bind thrombin with an N- to C-polypeptide chain sense opposite that of the<br />

substrate and form interactions similar to those made <strong>by</strong> the first three residues of hirudin (lle1, Thr2,<br />

Tyr3). In the x-ray crystal structure of BMS-183507 (K i = 17.2 nM) with thrombin [15], the N terminus<br />

is facing the catalytic site while the methyl ester is<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_252.html (1 of 2) [4/5/2004 5:10:35 PM]


Document<br />

Figure 6<br />

Schematic representation of the principal intermolecular<br />

interactions of a nonpeptide bicyclic inhibitor within the<br />

active site of thrombin.<br />

exposed to solvent. No specific interactions were observed with the catalytic triad. A bound water<br />

molecule hydrogen bonded to the Ser195 hydroxyl.<br />

Page 253<br />

The complex is stabilized <strong>by</strong> a network of hydrogen bonds as well as hydrophobic interactions. The<br />

Phe1-O and the Phe3-NH form hydrogen bonds with Gly216, and the Phe1-NH hydrogen bonds to the<br />

backbone carbonyl of Ser214. The Phe1 phenyl group occupies the S2 site, while Phe3 interacts within<br />

the S3 site.<br />

The retro-inhibitors contain a 4-guanidinobutanoyl group that extends into the S1 specificity site. Rather<br />

than forming two hydrogen bonds between the guanidine and Asp189 in a manner similar to PPACK,<br />

BMS-183507 forms only one, with the second hydrogen bond being directed to the carbonyl oxygen of<br />

Gly219. Binding affinity, as evidenced <strong>by</strong> loss of more than two orders of magnitude in affinity on<br />

addition of one or two methylene groups, was sensitive to chain length at this position.<br />

The allo-Thr hydroxyl oxygen accepts a hydrogen bond from the backbone NH of Gly219. This<br />

additional interaction accounts, at least in part, for the increase in affinity when compared to the<br />

inhibitor with Leu in this position. Comparison of the crystal structures of thrombin complexed with<br />

BMS-183507 and with hirudin reveals that the hirudin residue, Thr2, and the allo-Thr of BMS-183507<br />

interact differently with thrombin. The hirudin Thr2 binds at S2,<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_253.html (1 of 2) [4/5/2004 5:10:38 PM]


Document<br />

Figure 7<br />

Schematic representation comparing the principal<br />

inhibitor-to-thombin interactions of related inhibitors with either<br />

Leu (7a) or allo-Thr (7b) at P3.<br />

Page 254<br />

whereas the allo-Thr sidechain is oriented toward the protein exterior and is partially exposed to solvent.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_254.html (1 of 2) [4/5/2004 5:10:42 PM]


Document<br />

Cyclotheonamide A (CtA), a macrocyclic marine natural product derived from the Japanese sponge,<br />

Theonella sp., inhibits thrombin with an IC 50 value of<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_254.html (2 of 2) [4/5/2004 5:10:42 PM]


Document<br />

Figure 8<br />

Schematic representation of the principal intermolecular<br />

interactions of cyclotheonamide A within the active site of<br />

thrombin.<br />

Page 255<br />

100 nM and represents a novel structural class of serine protease inhibitors. An x-ray crystal structure of<br />

CtA complexed with thrombin was used to determine the molecular basis for this inhibition (Figure 8<br />

[16]). The Arg-Pro unit binds to the S1 and S2 sites in a manner similar to the Arg-Pro of PPACK. The<br />

Arg guanidinium group forms a bidentate hydrogen bond with Asp189 while the Pro establishes a βsheet<br />

interaction with the Ser214-Gly216 backbone. The α-ketoamide acts as a transition-state mimetic<br />

forming a tetrahedral hemiketal with the hydroxyl of Ser195. Within the complex, CtA adopts a<br />

relatively open conformation with the Pro orthogonal to the macrocycle and confined <strong>by</strong> a hydrophobic<br />

pocket defined <strong>by</strong> Tyr60A, Trp60D, and Leu99. Two aromatic residues are involved in stacking<br />

interactions with Tyr60A and Trp60D. Cyclotheonamide A, however, does not effectively match the S3<br />

interactions provided <strong>by</strong> the D-Phe group found in PPACK. In CtA, the formamide group is too polar to<br />

effectively complement the S3 site adjacent to Trp215. The authors note that the complex of CtA with<br />

thrombin does not appear optimal and suggest that synthetic analogs could significantly improve both<br />

potency and selectivity.<br />

Starting with the known thrombin inhibitors Argatroban and Nα(2-naphthyl-sulfonyl-glycyl)-DL-pamidinophenylalanyl-piperidine<br />

(NAPAP), a group at Roche initiated a medicinal chemistry program to<br />

develop thrombin inhibitors with reduced toxicity and an improved hemodynamic profile [17].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_255.html (1 of 2) [4/5/2004 5:10:45 PM]


Document<br />

Page 256<br />

The discovery program proceeded in four iterative phases which are shown in Table 1. Initial screening<br />

of low molecular weight organic bases led to the discovery of 1-amidinopiperidine (1–1) as a new<br />

surrogate for the guanidine and amidine functionality in Argatroban and NAPAP, respectively. A<br />

distinct advantage of 1-amidinopiperidine is its intrinsic selectivity for thrombin over<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_256.html (1 of 2) [4/5/2004 5:10:50 PM]


Document<br />

Page 257<br />

trypsin. Application of three-fold iterative strategy of design involving synthesis, x-ray crystallography,<br />

and molecular modeling, this group elaborated the 1-amidinopiperidine from structures that inhibited in<br />

the micromolar range to some inhibiting in the picomolar range. In doing so significant improvements in<br />

the selectivity of thrombin relative to trypsin were also achieved. In the case of the D-amino acid series<br />

(1–2), a “second inhibitor binding mode” that differed from that of Argatroban was identified. In this<br />

new and unexpected binding mode, the S2 pocket is unoccupied and the napthalenesulfonyl group fills<br />

the S3 site and overlaps the front of the S2 site. The benzyl group of the phenylalanine is oriented<br />

toward the protein surface and is partially exposed to solvent. The Argatroban or “inhibitor binding”<br />

mode was favored <strong>by</strong> the more potent L-amino acid series (1–3 and 1–4) where the piperidide (1–3) or Nbenzyl<br />

(1–4) binds to the S2 site and the aryl groups are found in the S3 site.<br />

IV. Bivalent Thrombin Inhibitors Directed at the Fibrinopeptide a Binding Pocket and the<br />

Fibrinogen Recognition Site<br />

A strategy to prepare highly selective thrombin inhibitors involves linkage of molecules capable of<br />

interacting at distinct subsites. This approach should result in inhibitors more specific for thrombin:<br />

while serine proteases possess common structural features related to catalysis and some serine<br />

proteases—including the coagulation enzyme Factor Xa—also exhibit primary substrate specificity for<br />

positively charged residues, only thrombin possesses recognition subsites for fibrinogen and effector<br />

molecules such as thrombomodulin. Nature has used this strategy in the evolution of hirudin, the<br />

anticoagulant protein produced <strong>by</strong> the medicinal leech. When this effective anticoagulant binds<br />

thrombin [18–20], the N-terminal domain blocks the primary specificity pocket while the C-terminal<br />

residues adopt an extended conformation and make multiple interactions within the fibrinogen<br />

recognition exosite.<br />

Guided <strong>by</strong> structural and biochemical information, small molecules capable of simultaneous interactions<br />

with both the primary specificity pocket and the fibrinogen recognition exosite were designed and<br />

synthesized. These bivalent inhibitors are composed of three regions: a group to block the primary<br />

specificity pocket, a sequence to bind the fibrinogen recognition site, and a chemical linker. The bivalent<br />

inhibitor approach was first executed with peptides [21–22]. In 1990, DiMaio et al. (3–3 [22]) used the<br />

peptide sequence from hirudin to link (d-Phe)-Pro-Arg-Pro, known to bind at the primary specificity<br />

pocket [23], with hirudin C-terminal residues, known to bind at the fibrinogen recognition site.<br />

Polyglycine linkers were also used to connect these sequences (Maraganore<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_257.html [4/5/2004 5:10:52 PM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_258.html [4/5/2004 5:11:22 PM]<br />

Page 258


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_259.html [4/5/2004 5:11:26 PM]<br />

Page 259


Document<br />

et al. [21]). Among these hirudin analogs, the tetraglycine linker appeared optimal (3–8, K i = 2.3 nM,<br />

Table 2).<br />

Page 260<br />

Most of the peptide-<strong>based</strong> bivalent inhibitors were slowly cleaved <strong>by</strong> thrombin. Incorporation of a<br />

ketomethylene pseudo peptide bond (3–4) resulted in a noncleavable bivalent inhibitor that retained high<br />

thrombin affinity [24]. Decreased proteolysis in bivalent inhibitors increasingly nonpeptide in character<br />

continues to be observed.<br />

Chemically simpler linkers were made using multiple methylene-containing glycine variants [25]. The<br />

dependence of affinity on placement of amide linkage within linkers containing the same number of<br />

atoms indicated some specific thrombin-to-linker interactions (3–13,14,15,16). This was confirmed in<br />

the crystal structure of hirutonin-6:thrombin complex (3–26 [26]) where continuous electron density was<br />

observed for the entire bivalent inhibitor including the linker region.<br />

The extended nature of the fibrinogen recognition site complicates attempts to reduce inhibitor<br />

molecular weight while maintaining affinity. Although of similar molecular weight, substitution of the<br />

sequence -Asp-Tyr-Glu-Pro-lle-Pro-Glu-Glu-Ala-cyclohexylalanine-(D-Glu) for -Asp-Phe-Glu-Glu-lle-<br />

Pro-Glu-Glu-Tyr-Leu-Gin increases affinity an order of magnitude (compare 3–17 and 3–18). Within a<br />

series of bivalent inhibitors, inclusion of sulfated tyrosine, the naturally occurring residue of hirudin,<br />

increases affinity 5 to 6 fold (3–8 compared to 3–11, and 3–1 to 3–2). Only seven residues are present in<br />

one of the smallest bivalent inhibitors (3–26).<br />

Increasingly nonpeptide substituents have been incorporated into the primary specificity pocket binding<br />

portion of the bivalent inhibitors. Higher affinity for thrombin was achieved <strong>by</strong> replacement of the (D-<br />

Phe)-Pro-Arg with either dansyl-Arg-(D-pipecolic acid) (3–17, [27]) or 4-tert-butylbenzenesulfonyl-Arg-<br />

(D-pipecolic acid) (3–18, [27]). While the arginine side chain of these and the (D-Phe)-Pro-Argcontaining<br />

inhibitors make similar interactions with the aspartic acid within the S1 specificity pocket,<br />

the dansyl-Arg-(D-pipecolic acid) inhibitors bind in a nonsubstrate mode [27]. This initial result suggests<br />

that other nonpeptide thrombin inhibitors may be successfully incorporated into bivalent inhibitors.<br />

Recently, a pyridinium methyl ketone bivalent inhibitor capable of forming a reversible covalent<br />

complex with thrombin was synthesized (3–26, [28]). Crystallographic analysis of its complex with<br />

thrombin showed the ketone carbonyl becomes tetrahedrally coordinate <strong>by</strong> bonding to the side chain of<br />

thrombin's active site residue, Ser195. Substitutions of cyclohexylalanine for phenylalanine (3–4<br />

compared to 3–5) and the cyclohexylalanine-containing fibrinogen recognition peptide for the hirudin<br />

sequence (3–17 compared to 3–18) also contribute to the increased affinity of this bivalent inhibitor.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_260.html [4/5/2004 5:11:28 PM]


Document<br />

V. Inactivated Thrombin as an Inhibitor of Clot Formation<br />

Page 261<br />

A means to selectively inhibit thrombin's role in coagulation while preserving its anticoagulant functions<br />

involves site-directed mutagenesis of thrombin itself. By introduction of a single mutation, Gibbs et al.<br />

[29] altered thrombin's relative specificity for fibrinogen and protein C. The engineered thrombin's<br />

increased activation of protein C over fibrinogen cleavage offers the possibility of inhibiting clot<br />

formation with a modified human protein, a molecule likely to exhibit few side effects.<br />

VI. The Role of Structural Information<br />

The discovery of thrombin inhibitors has benefited from available protein structural information. Models<br />

of the thrombin overall structure and its active site geometry, constructed from available structures of<br />

related serine proteases [30], aided in the design of the mechanism-<strong>based</strong> inhibitors such as PPACK [31]<br />

and its boroarginine analog [10]. The unexpected, nonsubstrate binding mode of early thrombin<br />

inhibitors such as NAPAP was revealed <strong>by</strong> x-ray crystallographic analyses [32]. Iterative structure<strong>based</strong><br />

design methods have been critical in the optimization of bivalent inhibitors and inhibitors directed<br />

at the primary specificity pocket. <strong>Structure</strong>s of inhibitor:thrombin complexes are essential for the<br />

optimization of substitutents forming interactions within the aryl-binding site of the primary specificity<br />

pocket. In some cases (e.g. Table 1), seemingly minor alterations of the inhibitor can result in dramatic<br />

changes in the inhibitor's overall interactions with thrombin [17].<br />

<strong>Drug</strong> discovery efforts have also been strongly influenced <strong>by</strong> results of structural studies of thrombin<br />

complexed with effectors and substrate peptides. For example, recently the structures of thrombin<br />

complexed with fibrinopeptide A [33] and human prothrombin fragment F1 [34] have been determined.<br />

In addition to their role in design of high-affinity inhibitors, these structures provide valuable insights<br />

for design of drugs specific for the various subsites and conformational states of thrombin.<br />

VII. Conclusion<br />

Discovery of therapeutically effective thrombin inhibitors involves issues such as affinity and<br />

selectivity, bioavailability, and formulation. In addition to these relatively common concerns, the<br />

complex in vivo mechanisms designed to<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_261.html [4/5/2004 5:11:30 PM]


Document<br />

balance its pro- and anticoagulant activities present additional challenges in the discovery of<br />

therapeutically effective thrombin inhibitors.<br />

References<br />

1. Tapparelli C, Metternich R, Ehrhardt C, Cook NS. Synthetic low-molecular weight thrombin<br />

inhibitors: molecular design and pharmacological profile. TIPS 1993; 14:366–376.<br />

2. Stubbs MT, Bode W. <strong>Structure</strong> and specificity in coagulation and its inhibition. Trends Cardiovasc<br />

Med 1995; 5:157–166.<br />

Page 262<br />

3. Stone SR. Thrombin Inhibitors: A new generation of antithrombotics. Trends Cardiovasc. Med. 1995;<br />

5:134–140.<br />

4. Harker LA. Strategies for inhibiting the effects of thrombin. Blood Coagulation and Fibrinolysis<br />

1994; 5:47–58.<br />

5. Claeson G. Synthetic peptides and peptidomimetics as substrates and inhibitors of thrombin and other<br />

proteases in the blood coagulation system. Blood Coagulation and Fibrinolysis 1994; 5:411–436.<br />

6. Bode W, Mayr I, Baumann U, Huber R, Stone SR, Hofsteenge J. The refined 1.9 Å crystal structure<br />

of human α-thrombin: interaction with D-Phe-Pro-Arg chloromethylketone and significance of the Tyr-<br />

Pro-Pro-Trp insertion segment. EMBO J. 1989; 8:3467–3475.<br />

7. Bode W, Turk D, Karshikov A. The refined 1.9-Å X-ray crystal structure of D-Phe-Pro-Arg<br />

chloromethylketone-inhibited human α-thrombin: structure analysis, overall structure, electrostatic<br />

properties, detailed active-site geometry, and structure-function relationships. Protein Science 1992;<br />

1:426–471.<br />

8. Stubbs MT, Bode W. A Player of many parts: the spotlight falls on thrombin's structure. Thrombosis<br />

Research. Vol. 69. Pergamon Press, 1993; 1–58.<br />

9. Whinna HC, Church FC. Interaction of thrombin with antithrombin, heparin cofactor II and protein C<br />

inhibitor. Journal of Protein Chemistry 1993; 12:677–688.<br />

10. Weber PC, Lee S-L, Lewandowski FA, Schadt MC, Chang C“<br />

H, Kettner C. Kinetic and crystallographic studies of thrombin with Ac-(D)Phe-Pro-boroArg-OH and its<br />

lysine, amidine, homolysine and ornithine analogs. Biochemistry 1995; 34:3750–3757.<br />

11. Deadman JJ, Elgendy S, Goodwin C, Green D, Baban J, Patel G, Skordalakes E, Chino N, Claeson<br />

G, Kakkar V, Scully M. Characterization of a class of peptide boronates with neutral P1 side chains as<br />

highly selective inhibitors of thrombin. J Med Chem 1995; 38:1511–1522.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_262.html (1 of 2) [4/5/2004 5:11:34 PM]


Document<br />

12. Wu T-P, Yee V, Tulinsky A, Chrusciel R, Nakanishi H, Shen R, Priebe C, Kahn M. The structure of<br />

a designed peptidomimetic inhibitor complex of α-thrombin. Protein Engineering 1993; 5:471–478.<br />

13. Obst U, Gramlich V, Diederich F, Weber L, Banner DW. <strong>Design</strong> of novel, nonpeptidic thrombin<br />

inhibitors and structure of a thrombin-inhibitor complex. Angew Chem Int Ed Engl 1995; 34:1739.<br />

14. Iwanowicz EJ, Lau WF, Lin J, Roberts DGM, Seiler SM. Retro-binding tripeptide thrombin active<br />

inhibitors: discovery, synthesis and molecular modeling. J Med Chem 1994; 37:2122–2124.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_262.html (2 of 2) [4/5/2004 5:11:34 PM]


Document<br />

Page 263<br />

15. Tabernero L, Chang CÁ, Ohringer SL, Lau WF, Iwanowicz EJ, Han W-C, Wang T, Seiler S,<br />

Roberts D, Sack JS. <strong>Structure</strong> of a retro-binding peptide inhibitor complexed with human α-thrombin. J<br />

Mol Biol 1995; 246:14–20.<br />

16. Maryanoff BE, Qiu Z, Padmanabhan KP, Tulinsky A, Almond HR, Andrade-Gordon P, Greco M,<br />

Kauffman J, Nicolaou KC, Liu A, Brung P, Fusetani N. Molecular basis for the inhibition of human αthrombin<br />

<strong>by</strong> the macrocyclic peptide cyclotheonamide A. Natl Academy Sci USA 1993; 90:8048–8052.<br />

17. Hilpert K, Ackermann J, Banner DW, Gast A, Gubernator K, Hadvary P, Labler L, Muller K,<br />

Schmid G, Tschopp T, Van de Waterbeemd H. <strong>Design</strong> and synthesis of potent and highly selective<br />

thrombin inhibitors. J Med Chem 1994; 37:3889–3901.<br />

18. Rydel TJ, Tulinsky A, Bode W, Ravichandran KG, Huber R, Roitsch R, Fenton JW, II. The structure<br />

of a complex of recombinant hirudin and human α-thrombin. Science 1990; 249:277–280.<br />

19. Grutter MG, Priestle JP, Rahuel J, Grossenbacher H, Bode W, Hofsteenge J, Stone SR. Crystal<br />

structure of the thrombin-hirudin complex: a novel mode of serine protease inhibitor. EMBO J 1990;<br />

9:2361–2365.<br />

20. Rydel TJ, Tulinsky A, Bode W, Huber R. Refined structure of the hirudin-thrombin complex. J Mol<br />

Biol 1991; 221:583–601.<br />

21. Maraganore JM, Bourdon P, Jablonski J, Ramachandran KL, Fenton JW, II. <strong>Design</strong> and<br />

characterization of hirulogs: a novel class of bivalent peptide inhibitors of thrombin. Biochemistry 1990;<br />

29:7095–7101.<br />

22. DiMaio J, Gibbs B, Munn D, Lefebvre J, Ni F, Konishi Y. Bifunctional thrombin inhibitors <strong>based</strong> on<br />

the sequence of hirudin. J Biol Chem 1990; 265:21698–21703.<br />

23. Kettner C, Shaw E. D-Phe-Pro-Arg CH 2C1—A selective affinity label for thrombin. Thromb Res<br />

1979; 14:969–973.<br />

24. DiMaio J, Ni F, Gibbs B, Konishi Y. A new class of potent thrombin inhibitors that incorporates a<br />

scissile pseudopeptide bond. FEBS 1991; 282:47–52.<br />

25. DiMaio J, Gibbs B, Lefebvre J, Konishi Y, Munn D, Yue SY. Synthesis of a homologous series of<br />

ketomethylene arginyl pseudodipeptides and application to low molecular weight hirudin-like thrombin<br />

inhibitors. J Med Chem 1992; 35:3331–3341.<br />

26. Zdanov A, Wu S, DiMaio Y, Konishi Y, Li Y, Wu X, Edwards B, Martin P, Cygler M. Crystal<br />

structure of the complex of human α-thrombin and nonhydrolyzable bifunctional inhibitors, hirutonin-2<br />

and hirutonin-6. PROTEINS: <strong>Structure</strong>, Function and Genetics 1993; 17:252–265.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_263.html (1 of 2) [4/5/2004 5:11:39 PM]


Document<br />

27. Tsuda Y, Cygler M, Gibbs BF, Pedyczak A, Fethiere J, Yue SY, Konishi Y. <strong>Design</strong> of potent<br />

bivalent thrombin inhibitors <strong>based</strong> on hirudin sequence: incorporation of nonsubstrate-type active site<br />

inhibitors. Biochemistry 1994; 33:14443–14451.<br />

28. Rehse PH, Steinmetzer T, Li Y, Konishi Y, Cygler M. Crystal structure of a peptidyl pyridinium<br />

methyl ketone inhibitor with thrombin. Biochemistry 1995; 34:11537–11544.<br />

29. Gibbs CS, Coutre SE, Tsiang M, Li WX, Jain AK, Dunn KE, Law VS, Tao CT, Matsumura SY,<br />

Mejza SJ, Paborsky LR, Leung LLK. Conversion of thrombin into an anticoagulant <strong>by</strong> protein<br />

engineering. Nature 1995; 378:413–416.<br />

30. Greer J. Comparative model-building of the mammalian serine proteinases. J Mol Biol 1981;<br />

153:1027–1042.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_263.html (2 of 2) [4/5/2004 5:11:39 PM]


Document<br />

31. Kettner C, Shaw E. D-PHE-PRO-ARGCH 2Cl-1 selective affinity label for thrombin. Thrombosis<br />

Research 1979; 14:969–973.<br />

Page 264<br />

32. Brandstetter H, Turk D, Hoeffken W, Grosse D, Sturzebecher J, Martin PD, Edwards BFP, Bode W.<br />

X-ray crystal structure of thrombin complexes with the benzamidine- and arginine-base inhibitors<br />

NAPAP, 4-TAPAP and MQPA: a starting point for elaborating improved antithrombotics. J Mol Biol<br />

1992; 226:1085–1099.<br />

33. Martin PD, Robertson W, Turke D, Bode W, Edwards BFP. The structure of residues 7–16 of the<br />

Aα-chain of human fibrinogen bound to bovine thrombin at 2.3 Å resolution. J Biol Chem 1992;<br />

267:7911–7920.<br />

34. Arni RK, Padmanabhan K, Padmanabhan KP, Wu TP, Tulinsky A. The structure of the non-covalent<br />

complex of prothrombin kringle 2 with PPACK-thrombin. Chem Phys Lipids; 1994; 67–68:59–66.<br />

35. Stone SR, Hofsteenge J. Kinetics of the inhibition of thrombin <strong>by</strong> hirudin. Biochemistry 1986;<br />

25:4622–4628.<br />

36. Witting JI, Bourdon P, Maraganore JM, Fenton JW II. Hirulog-1 and -B2 thrombin specificity.<br />

Biochem J 1992; 287:663–664.<br />

37. Bourdon P, Jablonski J, Chao BH, Maraganore JM. <strong>Structure</strong>-function relationships of hirulog<br />

peptide interactions with thrombin. FEBS 1991; 294:163–166.<br />

38. Szewczuk Z, Gibbs BF, Yue SY, Purisima E, Zdanvo A, Cygler M, Konishi Y. <strong>Design</strong> of a linker<br />

for trivalent thrombin inhibitors: interaction of the main chain of the linker with thrombin. Biochemistry<br />

1993; 32:3396–3404.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_264.html [4/5/2004 5:11:41 PM]


Document<br />

11<br />

<strong>Design</strong> of Antithrombotic Agents Directed at Factor Xa<br />

William C. Ripka<br />

Corvas International, Inc., San Diego, California<br />

I. Introduction<br />

Page 265<br />

Serine proteases have long been recognized as important players in a number of biochemical processes<br />

and their specific and selective inhibition provides multiple therapeutic opportunities [1]. In particular,<br />

the blood coagulation process is the result of an amplified cascade of proteolytic events in which several<br />

specific zymogens of serine proteases in blood are activated sequentially <strong>by</strong> selective cleavages to<br />

produce active enzymes [2,3]. This process, in pathological circumstances, may lead to the formation of<br />

a thrombus—an insoluble matrix of fibrin and platelets. Thrombosis is a serious medical problem in the<br />

United States and Europe as exemplified <strong>by</strong> the fact that half the people who die each year die of<br />

cardiovascular related problems. While much recent work in antithrombotic therapeutic approaches has<br />

focused on inhibition of thrombin, the central role that Factor Xa plays in the coagulation response to<br />

vascular injury also makes it an ideal pharmacological target for antithrombotic drug development. The<br />

recent report of the x-ray crystal structure of native Factor Xa [4] allows, for the first time, a wellfounded<br />

structure-<strong>based</strong> drug design approach for inhibitors. A number of reviews describing the<br />

biology [5–10] and chemistry [11,12] of Factor Xa inhibitors have appeared.<br />

II. Coagulation Cascade<br />

In the coagulation cascade (Figure 1), a highly amplified process leads to the formation of thrombin,<br />

which is the primary mediator for the conversion of fibrinogen to fibrin, as well as activation of platelets<br />

through the thrombin<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_265.html [4/5/2004 5:11:42 PM]


Document<br />

Figure 1<br />

The coagulation cascade.<br />

Page 266<br />

receptor. Thrombin generation is itself, however, the result of Factor Xa in complex with Factor Va on a<br />

phospholipid surface (the prothrombinase complex) acting on prothrombin. Vascular injury is the<br />

initiating event in the coagulation process, causing the activation of Factor Xa <strong>by</strong> the Factor VIIa/tissue<br />

factor complex. Factor Xa is, therefore, a central and crucial enzyme directly leading to the production<br />

of thrombin and its inhibition should be effective in blocking thrombogenesis. As a consequence of its<br />

key role early in the coagulation cascade process Factor Xa represents a potentially valuable therapeutic<br />

target for potent and specific inhibition.<br />

III. Proof of Principle for a Factor Xa Inhibitor<br />

In recent years the method <strong>by</strong> which certain hematophageous organisms maintain blood flow during<br />

feeding has been determined. Interestly, several of these organisms utilize Factor Xa inhibitors to<br />

prevent coagulation [13–15]; the tick anticoagulant peptide (TAP), a small protein isolated from the<br />

Ornithidoros moubata tick [13], and antistasin isolated from the Haementeria officinalis leech [14] are<br />

both potent and selective inhibitors of Factor Xa. As expected, these molecules are effective<br />

antithrombotics in several animal models of thrombosis (Table 1) and provide an important proof of<br />

principle with regard to the potential effectiveness of Factor Xa inhibitors as therapeutic anticoagulants.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_266.html [4/5/2004 5:11:49 PM]


Document<br />

Table 1 Factor Xa Inhibitors (TAP, Antistasin) in Experimental Models of Thrombosis<br />

Rat and rabbit models of venous thrombosis [25]<br />

Canine model of high shear, coronary arterial thrombosis [6,26]<br />

Canine model of femoral arterial thrombosis [27,28]<br />

Rhesus monkey model of acute disseminated intravascular coagulation [29,30]<br />

Baboon model of platelet dependent arterial thrombosis [9,31,32]<br />

Page 267<br />

In an alternate approach, it has been shown that a covalently blocked, activesite-modified Factor Xa<br />

(DEGR-Xa) [16] as well as a catalytically impaired recombinant form [17] can be effective<br />

anticoagulants in models of deep-vein thrombosis [18] and in canine arterial thrombosis models [8]. In<br />

these examples, the active-site-inactivated Factor Xa competes with the active form for incorporation<br />

into the active prothrombinase complex since the binding of Factor Xa to Factor Va in this complex is<br />

independent of the active site [20]. These studies with Factor Xa inhibitors suggest that inhibiting earlier<br />

in the coagulation cascade, as well as inhibiting the production of thrombin <strong>by</strong> inactivating Factor Xa in<br />

the prothrombinase complex, may have certain therapeutic advantages.<br />

IV. Factor Xa—<strong>Structure</strong> and Function<br />

Factor Xa is a 59 kilodalton protein synthesized in the liver and secreted into the blood as an inactive<br />

zymogen (Figure 2) [21]. Prior to secretion the singlechain molecule undergoes co- and posttranslational<br />

modifications including removal of a signal sequence [22–24], gamma carboxylation of<br />

several glutamic acids (Gla) in the N-terminus [33], beta hydroxylation of Asp 63 [34], N-glycosylation at<br />

two sites [35], and cleavage at two sites, Arg 139 and Arg 142, to give a two-chain molecule [36]. The<br />

mature form of Factor X consists of a light chain (139 amino acids) and a heavy chain (303 amino acids)<br />

held together <strong>by</strong> a single disulfide (Figure 2). The Gla residues are responsible for calcium and<br />

phospholipid binding and the second EGF domain is thought to mediate binding to Factor VIIIa and<br />

Factor Va [37,38]. The heavy chain contains the catalytic domain with the prototypic serine protease<br />

active site triad, His 226, Asp 279, and Ser 376. During coagulation, Factor X is converted to the active<br />

protease, Factor Xa, <strong>by</strong> a complex of Factor VIIa/tissue factor or a complex of Factor IXa/Factor,<br />

VIIIa/phospholipid, and calcium, both of which cleave a specific Arg-Ile bond to release an activation<br />

peptide (Figure 2) [39]. Similar to the activation of chymotrypsin, trypsin, and thrombin, the newly<br />

formed N-terminal Ile folds into the interior of the protein to form an ion pair at the active site with<br />

Asp] 375 [39,40]. In the presence of calcium ions the newly formed Factor Xa associates with Factor Va<br />

on a phospholipid membrane surface to form the<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_267.html (1 of 2) [4/5/2004 5:11:51 PM]


Document<br />

Figure 2<br />

Factor Xa structure. Residues of the catalytic triad (His 226 , Asp 279 , Ser 376 ) are circled.<br />

Page 268<br />

prothrombinase complex that rapidly converts prothrombin (PT) to thrombin <strong>by</strong> cleavage at two sites in<br />

PT, Arg 271–Thr 272 and Arg 320–Ile 321 [41]. The x-ray structure of human des(1–45) Factor Xa at 2.2 Å<br />

resolution has now been reported [4] (Figure 3).<br />

V. Natural Inhibitors of Factor Xa<br />

Several small, potent, and naturally occurring Factor Xa inhibitors—tick antico-agulant protein<br />

(TAP)[3], tissue factor pathway inhibitor (TFPI) [42], antistasin (ATS) [43], Ecotin [44,45]—have been<br />

isolated and characterized (Table 2). All but TAP apparently inhibit the enzyme in the extended<br />

substrate conformation referred to as the standard mechanism of inhibition (Figure 4) [47]. In this<br />

standard mechanism the inhibitor presents a conformationally constrained binding loop with a partial<br />

beta sheet motif to the target enzyme that mimics the required substrate conformation and, after binding<br />

to the enzyme, can undergo a reversible proteolytic hydrolysis at the reactive site peptide bond (P1–P1').<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_268.html [4/5/2004 5:11:59 PM]


Document<br />

Figure 3<br />

Stereo ribbon diagram of Factor Xa [4]. Residues of the catalytic triad are shown (His 57 , Asp 102 , Ser 195 ) as well as residues<br />

of specific interest for the binding of small molecules to the active site: Glu 192 ; S4 pocket residues Tyr 99 , Phe 174 , Trp 215 ; and<br />

S1 pocket residue, Asp 189 . Residues are designated with the chymotrypsin numbering.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_269.html [4/5/2004 5:12:37 PM]<br />

Page 269


Document<br />

Table 2 Naturally Occurring Inhibitors of Factor Xa<br />

Inhibitor Source K i Structural information<br />

TFPI Human 3 pM [49]<br />

90 nM [50]<br />

Ecotin Escherichia coli 50 pM [44] X-ray; complex with trypsin [46]<br />

TAP Ornithidoros moubata (tick) 135 pM [13] 2D-NMR [57,58]<br />

Antistasin Haementeria officinalis (leech) 61 pM [51] (X-ray in progress) [52]<br />

AcAP5 Ancylostoma caninum 43 pM [19] homology to Ascaris lumbricoides<br />

var.suum [79,80]<br />

Page 270<br />

Studies of these natural inhibitors can be useful in defining the active site requirements for Factor Xa<br />

inhibition, and importantly, can indicate the level of inhibition that may be necessary for an effective<br />

Factor Xa inhibitor, recognizing that TAP and antistasin have evolved to yield functional, in vivo<br />

antithrombotics. Table 3 shows the reactive-site sequences of these substratelike inhibitors as well as the<br />

cleavage site sequences recognized <strong>by</strong> Factor Xa in the activation of prothrombin (PT), Factor VII, and<br />

Factor V.<br />

A. Tissue Factor Pathway Inhibitor (TFPI)<br />

The mature tissue factor pathway inhibitor (TFPI) is a 276-residue protein consisting of three tandom<br />

domains with homology to the Kunitz-like protease<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_270.html (1 of 2) [4/5/2004 5:12:45 PM]


Document<br />

Figure 4<br />

Extended binding modes for substrates and inhibitors. Sites in the enzyme (S) and in the<br />

inhibitor (P) are designated <strong>by</strong> the Schecter-Berger notation [48].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_270.html (2 of 2) [4/5/2004 5:12:45 PM]


Document<br />

Table 3 Active Site Sequences for Factor Xa Substrates and Inhibitors<br />

Substrates P4 P3 P2 P1 P'1 P'2 P'3 P'4<br />

PT 271,272 Ile Glu Gly Arg Thr Ala Thr Ser<br />

PT 320,321 Ile Asp Gly Arg Ile Val Glu Gly<br />

FVII Pro Gln Gly Arg Ile Val Gly Gly<br />

FV Lys Lys Tyr Arg Ser Leu His Leu<br />

Inhibitors<br />

Antistasin Val Arg Cys Arg Val His Cys Pro<br />

Ecotin Val Ser Thr Met Met Ala Cys Pro<br />

TFPI-II Gly Ile Cys Arg Gly Tyr Ile Thr<br />

AcAP5 Cys Arg Ser Arg Gly Cys Leu Leu<br />

AcAP6 Cys Arg Ser Phe Ser Cys Pro Gly<br />

Page 271<br />

inhibitors [42]. A potent inhibitor of both Factor VIIa and Xa as well as trypsin, TFPI does not,<br />

however, have significant activity against leukocyte elastase, urokinase, activated Protein C, tissue<br />

factor plasminogen activator, thrombin, or kallikrein [53,54]. The second Kunitz domain from the Nterminus<br />

of TFPI has been identified as primarily responsible for the Factor Xa inhibition while both the<br />

first and second domains contribute to inhibition of Factor VIIa [42] (Figure 5). The proposed<br />

mechanism for this Factor-Xa-dependent inhibition of FVIIa/tissue factor involves the formation of a<br />

quarternary FXa-TFPI-FVIIa/TF complex [42]. The recombinant, isolated second domain, TFPI-II, has<br />

a K i for Factor Xa of 90 nM [50] compared to 3 pM [49] for the intact protein. The sequence of the<br />

P4–P5' region of the Factor Xa inhibitory second Kunitz domain (Table 3) has been incorporated into a<br />

prototypic Kunitz inhibitor, bovine pancreatic inhibitor (BPTI), to produce potent and selective Factor<br />

Xa inhibitors [75,76].<br />

B. Antistasin (ATS)<br />

Antistasin is one of several anticoagulants isolated from the Mexican leech, Haementeria officinalis<br />

[14]. It is a 119-amino-acid cysteine-rich protein with a primary structure that shows a two-fold<br />

sequence symmetry suggesting the molecule possesses two separate and distinct domains [51].<br />

Mutagenesis studies have shown that ATS binds to Factor Xa in a substratelike manner in the<br />

P3(Arg 32–P'3(Cys 37) regions and is cleaved only in the first domain [55].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_271.html (1 of 2) [4/5/2004 5:12:47 PM]


Document<br />

While the x-ray structure of antistasin has not been reported, it is known that a Factor-Xa-induced<br />

cleavage occurs between Arg 34 and Val 35 suggesting this peptide loop conforms to the conformationally<br />

rigid substratelike conformation suggested <strong>by</strong> other known protein inhibitors of serine proteases [55]. A<br />

hallmark of this mode of inhibition is the rigid structure around the cleaved<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_271.html (2 of 2) [4/5/2004 5:12:47 PM]


Document<br />

Figure 5<br />

Predicted secondary structure of tissue factor pathway inhibitor (TFPI) showing the Factor<br />

Xa and Factor VIIa inhibitory domains. The arrows point to the P1 sites.<br />

Page 272<br />

bond, often imposed <strong>by</strong> cysteine crosslinks constraining the two ends of the cleavage site to be in close<br />

proximity even after cleavage [47]. Antistasin has cysteines at the P2(Cys 33) and P'3 (Cys 37) positions.<br />

C. Tick Anticoagulant Peptide (TAP)<br />

The tick anticoagulant peptide (TAP) is a 60-amino-acid polypeptide isolated from the soft tick<br />

Ornithodorus Moubata and is a potent (K i = 2–200 pM) and selective inhibitor of Factor Xa, both as the<br />

free enzyme and in the prothrombinase complex [13]. The TAP anticoagulant does not inhibit trypsin or<br />

other trypsinlike serine proteases and, importantly, is not cleaved <strong>by</strong> Factor Xa. The mechanism <strong>by</strong><br />

which TAP inhibits Factor Xa appears to be unique and it apparently does not utilize the substratelike<br />

binding modes characteristic of antistasin and the Kunitz inhibitors. Mutagenesis studies have shown<br />

that the primary interaction of TAP with Factor Xa occurs at the N-terminous where Arg 3 appears to<br />

play a key role [56]. The solution structure of TAP has been deter-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_272.html [4/5/2004 5:12:58 PM]


Document<br />

Figure 6<br />

Stereo diagram of the NMR solution structure of the tick anticoagulant peptide. The crucial<br />

N-terminus Arg 3 is indicated along with the pattern of cysteine bonds.<br />

Page 273<br />

mined <strong>by</strong> 2D-NMR studies [57,58] (Figure 6). With the exception of the region neighboring the<br />

Cys 15—Cys 39 bond in TAP, these studies support the originally proposed idea that there are significant<br />

structural similarities between TAP and Kunitz proteinase inhibitors [10]. Nevertheless, it is clear that<br />

TAP inhibitors FXa <strong>by</strong> a fundamentally different, and as yet, not fully understood mechanism.<br />

D. AcAP's<br />

In addition to ticks and leeches, other hematophagous organisms such as hook- worms have also evolved<br />

potent and selective Factor Xa inhibitors as anticoagulant strategies. Two such proteins, AcAP5 and<br />

AcAP6, have been isolated from the Ancylostoma caninum hookworm [19]. The AcAP5 protein is a 77amino-acid<br />

polypeptide with 10 cysteine residues. It inhibits the amidolytic activity of Factor Xa with a<br />

K i of 43 ± 5 pM. Incubation of rAcAP5 with its target enzyme Factor Xa results in partial cleavage of<br />

the Arg 40–Gly 41 peptide bond suggesting the sequence around this cleavage site can adopt the restricted<br />

conformational requirements of substrates [47].<br />

The AcAP6 protein is a 75-amino-acid polypeptide, also with cysteines, with a K i for Factor Xa<br />

inhibition of 996 ± 65 pM. Alignment of the sequences of AcAP5 and AcAP6 suggest the P1 residue in<br />

AcAP6 is Phe 38 and not the basic residue usually associated with Factor Xa specificity [19]. Substitution<br />

of Phe 38 in AcAP6 with Arg resulted in a mutant that inhibited Factor Xa with a potency similar to<br />

rAcAP5. Both AcAP6 and ecotin suggest that an Arg or<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_273.html (1 of 2) [4/5/2004 5:13:03 PM]


Document<br />

Lys in the P1 position is not an absolute requirement for potent and selective activity.<br />

E. Ecotin<br />

Page 274<br />

Ecotin, a protein isolated from Escherichia coli, is a promiscuous protease inhibitor that potently inhibits<br />

kallikrein, urokinase, Factor XIIa, granzyme B, trypsin, chymotrypsin, and elastase (reviewed in<br />

Reference 46). As with most protein inhibitors (hirudin [59] and TAP [10] being the exceptions) ecotin<br />

presents a ‘baitlike’ substrate sequence to the target protease resulting in a reversible cleavage of the<br />

P1–P1' peptide bond. However, unlike other Factor Xa inhibitors that require basic residues such as Arg<br />

or Lys in the P1 position, ecotin is cleaved between two hydrophobic amino acids, Met 84 and Met 85 [44].<br />

Three preliminary crystal structures of ecotin with the serine proteases chymotrypsin, trypsin, and<br />

fiddler crab collagenase have been described [46]. The conformation of the sequence around the reactive<br />

site is similar to the bovine pancreatic trypsin inhibitor (BPTI) but differs in that the Cys at P' (not P2 as<br />

in BPTI) provides the rigidifying function for the reactive-loop sequence. Interestingly, antistasin has<br />

cysteines at both the P2 and P3' positions. The Pro at P4' is commonly found in FXa inhibitors including<br />

antistasin.<br />

In its complexes with the serine proteases for which structures are available [46] there is a sub van der<br />

Waals contact between Met 84-C and the enzyme Ser 195-O. The Met 84-O faces the oxyanion hole and<br />

forms hydrogen bonds with Ser 195 and Gly 193. In the trypsin structure the Met 84 side chain extends into<br />

the S1 site in a manner similar to Lys 15 in the BPTI-trypsin complex. Ecotin also forms both beta sheet<br />

hydrogen bonds to the enzyme Gly 216, Ser 82-N and O to Gly 216 O and N.<br />

VI. Small Molecule Inhibitors of Factor Xa<br />

While the x-ray structure of native Factor Xa has been reported [4] the nature of its crystal packing,<br />

specifically the fact that the active site of one Factor Xa molecule is blocked <strong>by</strong> the N-terminus of a<br />

second resulting in a “continuous polymeric structure,” apparently has precluded diffusing inhibitors<br />

into the preformed crystals to obtain complexes. Complexes with inhibitors cocrystallized with Factor<br />

Xa also have not been reported [81]. Thus, efforts to do structure-<strong>based</strong> design with this enzyme have<br />

relied on molecular modeling.<br />

Since, to date, it has not been possible to directly obtain x-ray structures of inhibitor complexes with<br />

Factor Xa, the substantial information available with respect to how serine proteases, particularly<br />

thrombin, bind inhibitors can<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_274.html [4/5/2004 5:13:05 PM]


Document<br />

Page 275<br />

be utilized to model known Factor Xa inhibitors using the x-ray coordinates of native Factor Xa.<br />

Preliminary modeling of the several inhibitors described below into the Factor Xa active site was<br />

accomplished <strong>by</strong> the author [69] <strong>by</strong> first superimposing the backbone atoms (N, Ca, C) of the catalytic<br />

triads (His 57, Asp 102, Ser 195 in the benzamidine:thrombin and PPACK:thrombin x-ray structures on the<br />

corresponding residues in Factor Xa. This allowed an excellent fit of the P1 basic groups,<br />

arylbenzamidine in the case of benzamidine and arginine for PPACK, into the S1 pocket of Factor Xa.<br />

These groups were then used as templates for positioning the appropriate P1 basic groups of the various<br />

synthetic inhibitors [69]. Holding these docked P1 groups fixed, the remaining rotatable bonds were<br />

manipulated to allow a reasonable and complementary fit of the inhibitor atoms to the solvent accessible<br />

surface of the Factor Xa active site. In those cases where it was possible, hydrogen bonds, particularly to<br />

Gly 216, were formed. To fit extended peptide sequences such as that for antistasin and the antistasinderived<br />

peptides described below, the backbone atoms of the residues around the cleavage site (e.g.,<br />

P4–P4') were positioned using the corresponding BPTI backbone atoms as a template. This was done<br />

after first aligning the trypsin catalytic triad backbone in the BPTI:trypsin x-ray complex (vida infra) to<br />

Factor Xa. Where the side chains of the bound peptide segments differed from BPTI, their orientation<br />

was either modeled for maximum complementarity to the Factor Xa molecular surface or set <strong>by</strong> an<br />

algorithmic approach [78]. In some cases the structures obtained were energy minimized initially <strong>by</strong><br />

steepest descent followed <strong>by</strong> conjugate gradient minimization.<br />

Recently, compounds <strong>based</strong> on a bisamidine motif (e.g., DX-9065a) have been reported as potent and<br />

selective Factor Xa inhibitors (1, DX-9065a) [60–63].<br />

The position of the amidino group makes little difference to the Factor Xa potency of these compounds<br />

but, interestingly, has a dramatic effect on the selectivity towards thrombin [62]. It was also observed<br />

that one carboxylic acid isomer (CX-9065a) was 7 times more potent on Factor Xa than the other. A<br />

second set of analogs shows a similar SAR [60].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_275.html [4/5/2004 5:13:08 PM]


Document<br />

Page 276<br />

In both cases the acids are much more selective for Factor Xa over thrombin. An initial modeling study<br />

using a homology-built Factor Xa structure proposed a fit of DX9065a to Factor Xa in which the<br />

amidinoary1 group occupies the S1 pocket and the acetimidoyl group is directed out of the S4 pocket<br />

[60].<br />

Lin et al. [64] have provided a more systematic study of possible fits of compound 2 and DX9065a<br />

using the recently available Factor Xa coordinates [4]. After aligning the His 57, Ser 195, and Asp 102<br />

backbone atoms for Factor Xa and thrombin (in the benzamidine:thrombin x-ray structure [65]) the<br />

arylamidino group of 2 was superimposed on the benzamidine template and a systematic conformational<br />

search was performed on the rotatable bonds of inhibitor 2. Energetics and complementarity to the<br />

Factor Xa surface determined a saved set, about 300 low-energy conformations, for further study. The<br />

final result was an optimized structure in which the acetimino group of 2 fits into<br />

Figure 7<br />

Molecular modeling fit of compound 2 with the arylamidino group positioned in the S1<br />

pocket and the acetimino group in the S4 cation-π site [69].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_276.html (1 of 2) [4/5/2004 5:13:12 PM]


Document<br />

Page 277<br />

the S4 pocket formed <strong>by</strong> the aromatic residues Trp 215, Tyr 99, and Phe 174 (Figure 7). It was proposed that<br />

this collection of aryl residues forms the basis for a cation-π interaction that has recently been well<br />

documented in other cases [66]. Interestingly, this cation-π site is absent in thrombin's equivalent “aryl<br />

binding site” where the corresponding residues are Trp 215, Leu 99, and Ile 174—residues that cannot<br />

provide the π-electrons necessary for stabilization of the cation in the inhibitor. The apparent preference<br />

for FXa inhibitors to have cations in the P3,P4 position may be directly related to the ability of the<br />

Factor Xa S4 site to stabilize these cations via a cation-π interaction. Factor Xa appears to be unique<br />

among the coagulation factors in providing this electron-rich S4 pocket (Table 4).<br />

The initial discovery of bisamidine structures as potential Factor Xa inhibitors was actually made much<br />

earlier with the finding that compounds such as 4 showed an almost 300-fold preference for Factor Xa<br />

over thrombin with a K i of 13 nM (FXa) [67],<br />

As with the DX-9065 analogs and compounds 2 and 3, the Factor Xa potency was relatively insensitive<br />

to the positioning of the amidino groups (4,4' versus 3,3') while replacing the 7-membered cycloalkyl<br />

ring with the 5 or 6 membered ring analogs reduced potency <strong>by</strong> about 10 fold. Model building<br />

compound 4 and docking into Factor Xa [69], again using the x-ray benzamidine:thrombin complex [65]<br />

as a template, shows that the second aryl amidino group can be positioned into the S4 aromatic pocket of<br />

Factor Xa in a conformation closely related to the mode of binding proposed for DX9065a (Figure 8)<br />

[64].<br />

In an effort to compare the relative efficacy of thrombin versus Factor Xa inhibitors, Markwardt et al.<br />

[67] synthesized a set of amidinoaryl compounds with moderate potency as Factor Xa inhibitors (5).<br />

Table 4 S4 Residues in Selected Serine Proteases<br />

Residue Position a Factor Xa Thrombin Factor VIIa Trypsin<br />

99 Tyr Leu Thr Leu<br />

174 Phe Ile Pro Gly<br />

215 Trp Trp Trp Trp<br />

aChymotrypsin numbering system. Sequence alignments <strong>by</strong> comparison of x-ray structures (sequence<br />

for Factor VIIa).<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_277.html (1 of 2) [4/5/2004 5:13:15 PM]


Document<br />

Page 278<br />

The n=3 chain length was optimal while the nature of the tosyl group on the α- nitrogen was relatively<br />

nonspecific, with tosylGly and Nα-β-naphthylsulfonyl- Gly being of similar potency. While the phenyl<br />

group on the amide nitrogen was best, other groups were also tolerated. Although the authors did not<br />

speculate on how these compounds were bound to Factor Xa, it is reasonable to suggest that the amidino<br />

phenyl group fits in the S1 pocket similar to the orientation determined <strong>by</strong> x-ray crystallography for<br />

benzamidine in the benzamidine:thrombin x-ray crystal structure. If the aryl amidino group of 5 is<br />

matched to that of benzamidine after alignment of the catalytic triad backbone atoms (N,Cα,C) for<br />

thrombin and Factor Xa, a proposed fit of 5 to Factor Xa can be made (Figure 9) [69]. This mode of<br />

binding is consistent with the structure activity relationships observed but does not suggest the reasons<br />

for the observed small preference for Factor Xa over thrombin.<br />

Figure 8<br />

Molecular modeling fit of compound 4 in the Factor Xa active site [69]. The carbonyl of the<br />

cycloheptanone makes a hydrogen bond (3.12 Å) to N-Gly 216 .<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_278.html (1 of 2) [4/5/2004 5:13:20 PM]


Document<br />

Figure 9<br />

Molecular modeling fit of compound 5 in the Factor Xa active site.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_279.html [4/5/2004 5:13:25 PM]<br />

Page 279


Document<br />

Figure 10<br />

Proposed model of dansyl-Glu-Gly-Arg-chloromethyl ketone in Factor Xa [4,69].<br />

Page 280<br />

Tulinsky and coworkers [4] have proposed a model for the complex of dansyl-Glu-Gly-Argchloromethyl<br />

ketone using the thrombin-PPACK crystal structure as a template for the fit to Factor Xa<br />

(Figure 10).<br />

Using antistasin as a starting point, Ohta et al. [68] have synthesized a series of cyclic peptides <strong>based</strong> on<br />

the antistasin sequence. Three of these peptides are shown below and represent the most potent in the<br />

series.<br />

ATS29-47 NH2-Ser-Gly-Val-Arg-Cys * -Arg-Val-His-Cys * -Pro-His-Gly-Phe-Gln-Arg-<br />

Ser-Arg-Tyr-Gly-OH<br />

K i (FXa)<br />

0.035 µM<br />

ATS29-40 NH2-Ser-Gly-Val-Arg-Cys * -Arg-Val-His-Cys * -Pro-His-Gly-OH 11.8 µM<br />

dR-ATS32-38 NH2-dArg-Cys-Arg-Val-His-Cys-Pro-OH 0.96 µM<br />

(The Cys * -Cys * are joined in disulfide bonds to form cyclic structures)<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_280.html (1 of 2) [4/5/2004 5:13:29 PM]


Document<br />

These molecules are cleaved <strong>by</strong> Factor Xa suggesting they bind in a similar manner to antistasin itself.<br />

Assuming the sequence around the cleavage site occupies the FXa active site locally in a manner similar<br />

to BPTI in the BPTI:trypsin complex, a modeled structure of the complex can be constructed<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_280.html (2 of 2) [4/5/2004 5:13:29 PM]


Document<br />

Figure 11<br />

dArg-ATS 32-38 modeled into the active site of Factor Xa utilizing the BPTI: trypsin x-ray structure as a template [69].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_281.html [4/5/2004 5:13:34 PM]<br />

Page 281


Document<br />

Page 282<br />

using BPTI residues 11–19 as a template [69]. The modeled dR-ATS 32-38:FXa complex is shown in<br />

Figure 11. Interestingly, these peptides do not inhibit trypsin even at 1000-fold higher concentrations<br />

than their FXa inhibitory concentrations. Antistasin, on the other hand, inhibits trypsin with a K i of 10<br />

nM.<br />

Preliminary reports have appeared describing a pentapeptide that is a potent (K i=3 nM) and selective<br />

inhibitor of Factor Xa (SEL2711) [70,71].<br />

Two possible modes of binding can be envisioned for this compound with either the methylpyridinium<br />

group occupying the S1 site in a “substratelike mode” or the p-amidinophenyl group in the S1 pocket,<br />

which would require a reversed binding reminiscent of the hirudin-thrombin interaction [59]. Figure 12<br />

shows the case for the amidinophenyl group in the Factor Xa S1 pocket [69]. In this mode of binding the<br />

methylpyridinium group easily fits the S4-aryl binding site and is well positioned for a π-cation<br />

interaction.<br />

Of interest from a drug-design viewpoint is the finding that cyclotheonamide, a compound isolated from<br />

a marine sponge and originally reported as a thrombin inhibitor, has been found to also inhibit Factor Xa<br />

with a K i of 50 nM [72]. Cyclotheonamide possesses a novel α-ketoamide transition state functionality<br />

and x-ray structures of cyclotheonamide with trypsin [73] and thrombin [74] provide templates for<br />

modeling this inhibitor into Factor Xa [69]. In the resulting fit (Figure 13) cyclotheonamide does not<br />

project functionality into the S4-cation-π site and would not be expected to show Factor Xa selectivity.<br />

VII. Defining the Requirements for Factor Xa Inhibition <strong>by</strong> Mutagenesis of BPTI<br />

It has been known for some time that many examples of naturally occurring Kunitz inhibitors exist, both<br />

isolated and as domains in larger proteins, which inhibit a variety of serine proteases [47]. This strongly<br />

suggests that this molecular framework is compatible with inhibition of this general class. The contact<br />

region between these inhibitors and their protease targets is known from a<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_282.html (1 of 2) [4/5/2004 5:13:38 PM]


Document<br />

Figure 12<br />

Modeled fit of SEL2711 in Factor Xa with the arylamidino group positioned in the S1 pocket and the methylpyridinium<br />

group in the cation-π S4 site.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_283.html [4/5/2004 5:13:43 PM]<br />

Page 283


Document<br />

Figure 13<br />

Cyclotheonamide modeled into Factor Xa utilizing the x-ray structure of cyclotheonamide:trypsin [73] and<br />

cyclotheonamide:thrombin [74] as templates [69].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_284.html [4/5/2004 5:13:51 PM]<br />

Page 284


Document<br />

Page 285<br />

number of x-ray structures of complexes. Appropriate site-specific and random mutagenesis, particularly<br />

of the P3-P4' residues of a prototypic Kunitz inhibitor, bovine pancreatic trypsin inhibitor (BPTI), has<br />

been shown to result in potent and selective Factor Xa inhibitors [75,76].<br />

A potent inhibitor of trypsin, kallikrein, and plasmin, BPTI does not inhibit Factor Xa. It binds to serine<br />

proteases such as trypsin in an extended substrate mode from residue 13 (P3) through 17 (P'2) [47]. A<br />

second loop from BPTI also extends into the active site bringing residues 34, 39, and 46 into contact<br />

with the protease-active site. In terms of spatial proximity of residues three clusters can be defined:<br />

cluster 1 (13,39); cluster 2 (11,17,19,34); and cluster 3 (16,18,20,46). While residue 39 is approximately<br />

in the same region of space as residue 13 (9.4 Å CB-CB) the CA rarrow.gif CB vectors are directed in<br />

different directions and substitution at 39 would not be expected to have a cooperative effect with<br />

residue 13. Residue 34 on the other hand is in a key position. It is centrally located between residues<br />

11,17, and 19 with CB-CB distances of 5.6, 5.7, and 6.5 Å respectively, and its CA rarrow.gif CB<br />

vector converges with the corresponding vectors from these residues to a common point in space. This<br />

residue is therefore expected to have a substantial cooperative effect with the other residues of cluster 2.<br />

Finally residue 46 is close to residue 20 (CB—CB of 6.5 Å) although the CA rarrow.gif CB vectors are<br />

approximately parallel and cooperative effects are expected to be minimal. The BPTI residues<br />

11,12,13,15–20, 34,39, and 46 were therefore the focus of the site-directed and random mutagenesis<br />

studies. Residue 14 is Cys in BPTI and was not modified in the mutants since it is required for structural<br />

reasons.<br />

A. Site Specific Mutagenesis<br />

As a starting point for the design of BPTI-<strong>based</strong> Factor Xa inhibitors, the second domain of TFPI (TFPI-<br />

II) was used as a template [75,76]. Table 5 shows the results of site-directed mutagenesis of BPTI.<br />

Mutant 50cl is a direct analog of TFPI-II with the exception of the Lys at position 46. The finding that<br />

4c2 and 4c10 are essentially equivalent in potency (K i 2.8 versus 1.8 nM) and are identical<br />

Table 5 Site-Directed BPTI Mutants with Factor Xa Inhibition<br />

K i(nM) 12 13 14 15 16 17 18 19 20 34 39 46<br />

r-TFPI-II 90 Gly Ile Cys Arg Gly Tyr Ile Thr Arg Lys Leu Glu<br />

50cl 205 Gly Ile Cys Arg Ala Tyr Ile Thr Arg Lys Leu Lys<br />

4c2 2.8 Gly Ile Cys Arg Ala Tyr Ile Thr Arg Val Leu Glu<br />

4c10 1.8 Gly Ile Cys Arg Ala Tyr Ile Thr Arg Val Leu Lys<br />

57c1 1.6 Gly Ile Cys Arg Ala Tyr Ile Ile Arg Val Leu Lys<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_285.html (1 of 2) [4/5/2004 5:13:54 PM]


Document<br />

w.t.<br />

BPTI<br />

>1 mM Gly Pro Cys Lys Ala Arg Ile Ile Arg Val Arg Lys<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_285.html (2 of 2) [4/5/2004 5:13:54 PM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_286.html [4/5/2004 5:14:20 PM]<br />

Page 286


Document<br />

Page 287<br />

cal in sequence with the exception of the Glu 46 Lys switch, suggests this residue is of minor importance.<br />

Therefore, the potency of 50cl (within a factor of 3) is consistent with r-TFPI-II. On the other hand, the<br />

incorporation of Val for Lys at position 34 leads to a dramatic increase in potency (˜100 fold; cf. 4c10<br />

and 4c2 with 50c1). From cluster 2 above this suggests the importance of P5, P'2,P'4 for potency. The<br />

changes in wild type BPTI that result in a potent Factor Xa inhibitor are Lys 15 rarrow.gif Arg 15, Arg 17,<br />

rarrow.gif Tyr 17, and Arg 39 rarrow.gif Leu 39.<br />

B. Random Mutagenesis<br />

11Libraries of mutant BPTI were created <strong>by</strong> inserting mutagenic cassettes in the BPTI gene of<br />

filamentous phage PIII coat proteins [76]. These libraries produced large numbers of mutants (~10 6)<br />

with randomized amino acids in positions 11, 13, 16, 17, 18, 19, 20, 34, and 39. The mutants were<br />

panned against Factor Xa, which was affixed to a solid support <strong>by</strong> a nonneutralizing antibody and the<br />

most potent inhibitors were separately expressed as soluble proteins. By this process it was possible to<br />

determine consensus sequences at the reactive sites and to define the pharmacophore requirements of<br />

inhibitors of Factor Xa in both a functional and conformational sense from the P4 to the P5 positions.<br />

Inhibitor amino acid preferences from both site directed and random mutagenesis studies are shown in<br />

Table 6.<br />

VIII. Positional Requirements of Factor Xa Inhibitors (Table 6)<br />

Examination of models of BPTI-mutants bound to Factor Xa show the L-amino acids in the P3 position<br />

project into solvent. In the Factor Xa cleavage sites in thrombin these residues are polar and acidic (Glu,<br />

Asp); they are polar and basic in antistasin (Arg), and polar and neutral in Ecotin (Ser). The exception is<br />

TFPI- II, with this position occupied <strong>by</strong> Ile. The BPTI random mutant results are consistent with the<br />

TFPI-II case and show a preference for aliphatics or aromatics in this position. There is a hydrophobic<br />

pocket in the enzyme, formed <strong>by</strong> Trp 215, Tyr 99, and Phe 174, that would be accessible to a D-residue in<br />

this position.<br />

The accessibility of the S2 pocket of the enzyme <strong>by</strong> P2 groups would be expected to be influenced <strong>by</strong><br />

the orientation of Tyr 99. As the x-ray structure shows its position this residue puts severe limitations on<br />

the size of the P2 group. Consistent with this is the observation that Gly is the sole residue in the<br />

FXa:thrombin cleavage sites. Little information is available from the BPTI mutants, TFPI-II, or<br />

antistasin, which all have a structural requirement for Cys at this position. Synthetic compounds show,<br />

however, that in potent inhibitors large bulky aromatics are, in fact, allowed at this position, a situation<br />

that requires Tyr 99 to move out of the way [75].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_287.html [4/5/2004 5:14:31 PM]


Document<br />

Page 288<br />

Additionally, the binding of BPTI mutants also requires Tyr 99 to move because of potential severe<br />

interactions with the Cys 14–Cys 38 bridge of the BPTI mutants. While larger groups can be<br />

accommodated, structural requirements are fairly rigid in agreement with the limited mobility expected<br />

of the Tyr 99. The apparent requirement for a P2-Gly in the larger substrates may suggest that in these<br />

cases extended binding occurs that does not permit movement of Tyr 99.<br />

In the P1 position, as expected, a basic group is preferred. Interestingly, of the two naturally occuring<br />

basic amino acids, arginine is preferred over lysine. This is seen in both the BPTI mutants as well as the<br />

Arg rarrow.gif Lys switch in antistasin. This may be due to Ala 190 in the S1 pocket of FXa, which<br />

cannot orient and stabilize the BPTI lysine analog as Ser 190 does in trypsin. An interesting exception to<br />

the need for a basic group is in Ecotin where a methionine occupies this site. The x-ray of Ecotin with<br />

trypsin clearly shows this neutral residue in the P1 pocket, aligned very closely to that seen for lysine in<br />

the BPTI:trypsin complex, and proximal to the charged Asp 189 [77]. Apparently, extended binding over<br />

the rest of the site compensates for this energetically unfavorable situation.<br />

In the P'1 position, the natural cleavage sites use Thr and Ile while Ecotin has Met. In contrast to<br />

thrombin, FXa lacks the 60-insertion loop and can accommodate large groups at this position. The BPTI<br />

mutants, however, are forced to use a small residue (Ala) because of steric hindrance from the Cys 58-<br />

Cys 42 group residue 61 in the enzyme. The inhibitor TFPI-II has a Gly at P'1.<br />

The BPTI mutants, TFPI-II, and antistasin all show a preference for aromatic groups at P'2. In the BPTI<br />

panning experiments Tyr was selected more than 80% of the time at this position. It can be seen from<br />

Table 5 that the Arg to Tyr change at position 17 is one of three significant changes that converts wild<br />

type BPTI from a non-Factor Xa inhibitor to a ˜1.6 nM inhibitor. There is a possible hydrogen-bond<br />

interaction between Tyr 17 of the inhibitor and Gln 192 of the enzyme, which may explain the strong<br />

preference.<br />

While models suggest the P'3 residue is directed at solvent and the FXa thrombin cleavage sites have<br />

polar residues at this position (Thr, Glu), the BPTI mutant results show a clear preference for a<br />

hydrophobic group. It is possible that aromatic groups can pack to Phe 41 of the enzyme. Mutant results<br />

show Ile is favored over Phe, His, which in turn is selected over Tyr.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_288.html (1 of 2) [4/5/2004 5:15:07 PM]


Document<br />

IX. Conclusions<br />

Page 289<br />

Factor Xa is clearly an important component of the coagulation process and inhibition of this enzyme<br />

can lead to potent anticoagulant effects. Recently, a number of naturally occurring anti-Factor-Xa<br />

polypeptides have been isolated from several hematophagous organisms including ticks, leeches, and<br />

hook- worms. With the exception of TAP, these molecules appear to bind to Factor Xa <strong>by</strong> the standard<br />

mechanism of inhibition proposed earlier [49]. The sequence information derived from these inhibitors<br />

as well as the natural cleavage sites of substrates of Factor Xa can be used along with the conformational<br />

constraints imposed <strong>by</strong> the proposed substrate-like binding to define the pharmacophore requirements of<br />

the active site of Factor Xa. The structurally rigid BPTI mutants, which have been found to be potent<br />

Factor Xa inhibitors, also provide important conformational information particularly with regard to the<br />

specific binding interactions on the P' side of the Factor Xa active site. A number of small molecule<br />

inhibitors have also recently been reported which appear to take advantage of a unique cation-π S4-site<br />

available in Factor Xa to achieve good selectivity with moderate potency. The availbility of the X-ray<br />

structure of native Factor Xa has allowed molecular modeling approaches to suggest possible fits of<br />

these inhibitors to the Factor Xa active site.<br />

Note Added in Proof<br />

After this review was written, the x-ray structure of Factor Xa with DX-9065a was reported [81].<br />

References<br />

1. Proteinase inhibitors. In: Barrett AJ, Salvesen G, eds. Research Monographs in Cell and Tissue<br />

Physiology. Vol 12. New York: Elsevier, 1986. <strong>Design</strong> of Enzyme Inhibitors as <strong>Drug</strong>s. Sandler M,<br />

Smith HJ, eds. New York: Oxford University Press, 1989.<br />

2. Colman RW, Hirsh J, Marder VJ, Salzman EW. Hemostasis and Thrombosis. Basic Principles and<br />

Clinical Practice. Second Edition. Philadelphia: J. B. Lippincott Company, 1987.<br />

3. Hathaway, WE, Goodnight, Jr SH. Disorders of Hemostasis and Thrombosis. New York: McGraw-<br />

Hill, 1993.<br />

4. Padmanabhan K, Padmanabhan KP, Tulinsky A, Park CH, Bode W, Huber R, Blankenship DT,<br />

Cardin AD, Kisiel W. <strong>Structure</strong> of human Des(1–45) factor Xa at 2.2 Å resolution. J Mol Biol 1993;<br />

232:947–966.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_289.html [4/5/2004 5:15:45 PM]


Document<br />

Page 290<br />

5. Kaiser B, Hauptmann J. Factor Xa inhibitors as novel antithrombotic agents: facts and perspectives.<br />

Cardiovascular <strong>Drug</strong> Reviews 1994; 12 (3):225–236.<br />

6. Lynch JJ,<br />

Sitko GR,<br />

Mellott MJ,<br />

Nutt EM,<br />

Lehman ED,<br />

Friedman PA,<br />

Dunwiddie CT,<br />

Vlasuk GP.<br />

Maintenance of<br />

canine coronary<br />

artery patency<br />

following<br />

thrombolysis<br />

with front<br />

loaded plus low<br />

dose<br />

maintenance<br />

conjunctive<br />

therapy. A<br />

comparison of<br />

factor Xa<br />

versus thrombin<br />

inhibition.<br />

Cardiovasc Res<br />

1994;<br />

28:78–85.<br />

7. Hollenback S, Sinha U, Lin P-H, Needham K, Frey L, Hancock T, Wong A, Wolf D. A comparative<br />

study of prothrombinase and thrombin inhibitors in a novel rabbit model of non-occlusive deep vein<br />

thrombosis. Thromb Haemost 1994; 71:357–362.<br />

8. Benedict CR, Ryan J, Todd J, Kuwabara K, Tijburg P, Cartwright Jr J, Stern D. Active site-blocked<br />

factor Xa prevents thrombus formation in the coronaryvasculature in parallel with inhibition of<br />

extravascular coagulation in a canine thrombosis model. Blood 1993; 81:2059–2066.<br />

9. Schaffer LW, Davidson JT, Vlasuk GP, and Siegl PKS. Antithrombotic efficacy of recombinant tick<br />

anticoagulant peptide. A potent inhibitor of coagulation factor Xa in a primate model of arterial<br />

thrombosis. Circulation 1991; 84:1741– 1748.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_290.html (1 of 2) [4/5/2004 5:16:10 PM]


Document<br />

10. Vlasuk GP. Structural and functional characterization of tick anticoagulant peptide (TAP): a potent<br />

and selective inhibitor of blood coagulation factor Xa. Thrombosis and Haemostasis 1993; 70:212–216.<br />

11. Scarborough RM. Anticoagulant strategies targeting thrombin and factor Xa. Ann Reports Med<br />

Chem 1995; 30:71–80.<br />

12. Wallis RB. Inhibitors of coagulation factor Xa: from macromolecular beginnings to small molecules.<br />

Current Opinion in Therapeutic Patents Aug, 1993.<br />

13. Waxman L, Smith DE, Arcuri KE, Vlasuk GP. Tick anticoagulant peptide (TAP) is a novel inhibitor<br />

of blood coagulation factor Xa, Science 1990; 248:593.<br />

14. Hauptmann J, Kaiser B, Vowak G, Struzebecher J, Markwardt F. Comparison of the anticoagulant<br />

and antithrombotic effects of synthetic thrombin and factor Xa inhibitors. Throm Haemostas 1990;<br />

63:220–223.<br />

15. Jacobs JW, Cupp EW, Sardana M, Friedman PA. Isolation and characterization of a coagulation<br />

factor Xa inhibitor from black fly salivary glands. Thromb Haemost (GERMANY) 1990; 64:235–238.<br />

16. Taylor Jr FB, Chang ACK, Peer GT, Mather T, Blick K, Catlett R, Lockhart MS, Esmon CT Blood<br />

1991; 78:364.<br />

17. Sinha U, Hancock T, Lin P-H, Hollenback S, Wolf D. Expression, purification, and characterization<br />

of inactive human coagulation factor Xa (Asn322Ala419). Protein Expr Purif 1992;3:518–524.<br />

18. Dunwiddie CT, Waxman L, Vlasuk GP, Friedman PA. Purification and characterization of inhibitors<br />

of blood coagulation factor Xa from hematophagous organisms. Methods in Enzymol 1993;<br />

233:291–312.<br />

19. Stanssens P, Bergum PW, Gansemans Y, Jespers L, LaRoche Y, Huang S, Maki S, Messeno J,<br />

Lauwereys M, Cappello M, Hotez PJ, Lasters I, Vlasuk GP. Anticoagulant repertoire of the hookworm<br />

ancyclostoma caninum. Proc Natl Acad Sci 1996; 93:2149–2154.<br />

20. Nesheim ME, Kettner C, Shaw E, Mann KG. Cofactor dependence of factor Xa incorporation into<br />

the prothrombinase complex. J Biol Chem 1981; 256:6537– 6540.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_290.html (2 of 2) [4/5/2004 5:16:10 PM]


Document<br />

Page 291<br />

21. Davie EW,<br />

Fujikawa K,<br />

Kisiel W. The<br />

coagulation<br />

cascade:<br />

interaction,<br />

maintenance, and<br />

regulation.<br />

Biochemistry<br />

1991;<br />

30:10363–10370.<br />

22. Leytus SP, Foster DC, Kurachi K, Davie EW. Gene for human factor X: a blood coagulation factor<br />

whose gene organization is Essentially identical with that of factor IX and protein C. Biochemistry<br />

1986; 25:5098–5102.<br />

23. Fung MR, Hay CW, MacGillivray RTA. Characterization of an almost full length cDNA coding for<br />

human blood coagulation Factor X. Proc Nat Acad Sci USA 1985; 82:3591–3595.<br />

24. Blanchard RA, Faye KLM, Barrett JM, William B. Isolation and characterization of profactor X<br />

from the liver of a steer treated with sodium warfarin. Blood 1985; 66(suppl.1):331a.<br />

25. Vlasuk GP, Ramjit D, Fujita T, Dumwiddie CT, Nutt EM, Smith DE, Shebuski RJ. Comparison of<br />

the in vivo anticoagulant properties of standard heparin and the highly selective factor Xa inhibitors<br />

antistasin and tick anticoagulant peptide (TAP) in a rabbit model of venous thrombosis. Thromb<br />

Haemostas 1991; 65:257– 262.<br />

26. Sitko GR, Ramjit DR, Stabilito II, Lehman D, Lynch JJ, Vlasuk GP. Conjunctive enhancement of<br />

enzymatic thrombolysis and prevention of thrombotic reocclusion with the selective factor Xa inhibitor,<br />

tick anticoagulant peptide. Comparison to hirudin and heparin in a canine model of acute coronary artery<br />

thrombosis. Circulation 1992; 85:805–815.<br />

27. Mellot MJ, Strainieri MT, Sitko GR, Stabilito II, Lynch JJ, Vlasuk GP. Enhancement of recombinant<br />

tissue plasminogen activator-induced reperfusion <strong>by</strong> recombinant tick anticoagulant peptide, a selective<br />

factor Xa inhibitor, in a canine model of femoral arterial thrombosis. Fibrinolysis 1993; 7:195–202.<br />

28. Mellot JJ, Holahan MA, Lynch JJ, Vlasuk GP, Dunwiddie CT. Acceleration of recombinant tissuetype<br />

plasminogen activator-induced reperfusion and prevention of reocclusion <strong>by</strong> recombinant<br />

antistasin, a selective factor Xa inhibitor, in a canine model of femoral arterial thrombosis. Circ Res<br />

1992; 70:1152–1160.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_291.html (1 of 2) [4/5/2004 5:16:13 PM]


Document<br />

29. Neeper MP, Waxman L, Smith DE, Schulman CA, Sardana M, Ellis RE, Schaffer LW, Siegel PKS,<br />

Valsuk GP. Characterization of recombinant tick anticoagulant peptide. A highly selective inhibitor of<br />

blood coagulation factor Xa. J Biol Chem 1990; 265:17746–17752.<br />

30. Dunwiddie CT, Nutt EM, Vlasuk GP, Siegel PDS, Schaffer LW. Anticoagulant efficacy and<br />

immunogenicity of the selective factor Xa inhibitor antistasin following subcutaneous administration in<br />

the rhesus monkey. Thromb Haemostas 1992; 67:371–376.<br />

31. Schaffer LW, Davidson JT, Vlasuk GP, Dunwiddie CT, Siegel PKS. Selective factor Xa inhibition<br />

<strong>by</strong> recombinant antistasin prevents vascular graft thrombosis in baboons. Arteriosclerosis and Thromb<br />

1992; 12:879–885.<br />

32. Kelly AP, Hanson SR, Dunwiddie CT, Harker LA. Circulation 1992; 86:411.<br />

33. Suttie JW. Vitamin K-dependent carboxylase. Annu Rev Biochem 1985; 54:459– 477.<br />

34. Fernlund P, Stenflo J. β-Hydroxy-aspartic acid in vitamin K-dependent proteins. J Biol Chem 1983;<br />

258:12509–12512.<br />

35. DiScipio RG, Hermodson MA, Davie EW. Activation of human factor X (Stuart factor) <strong>by</strong> a<br />

protease from Russell's viper venom. Biochemistry 1977; 16:5253– 5260.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_291.html (2 of 2) [4/5/2004 5:16:13 PM]


Document<br />

36. Leytus SP, Chung DW, Kisiel W, Kurachi K, Davie EW. Characterization of a cDNA coding for<br />

human factor X. Proc Nat Acad Sci USA 1984; 81:3699–3702.<br />

Page 292<br />

37. Skogen WF, Esmon CT, Cox AC. Comparison of coagulation factor Xa and des(1–44) factor Xa in<br />

the assembly of prothrombinase. J Biol Chem 1984; 259:2306–2310.<br />

38. Hertzberg MS, Ben-Tal O, Furie BC. Construction, expression, and characterization of a chimera of<br />

Factor IX and Factor X: the role of the second epidermal growth factor domain and serine protease<br />

domain in factor Xa binding. J Biol Chem 1992; 267:14759–14766.<br />

39. Sigler PB, Blow DM, Matthews BW, Henderson R. <strong>Structure</strong> of crystalline α- chymotrypsin. II. A<br />

preliminary report including a hypothesis for the activation mechanism. J Mol Biol 1968; 35:143–164.<br />

40. Bode W, Turk D, Karshikov A. The refined 1.9 Å X-ray crystal structure of D- Phe-Pro-Arg<br />

chloromethylketone-inhibited human α-thrombin. <strong>Structure</strong> analysis, overall structure, electrostatic<br />

properties, detailed active site geometry, structure function relationships. Protein Sci 1992; 1:426–471.<br />

41. Mann KG, Nesheim ME, Church WR, Haley P, Krishnaswamy S. Surface-dependent reactions of<br />

the vitamin K-dependent enzyme complexes. Blood 1990; 76:1– 16.<br />

42. Girard TJ, Warren LA, Novotny WF, Likert KM, Brown SG, Miletich JP, Broze Jr GJ. Functional<br />

significance of the Kunitz-type inhibitory domains of lipoprotein- associated coagulation inhibitor.<br />

Nature 1989; 338:518–520.<br />

43. Nutt E, Gasic T, Rodkey J, Gasic G, Jacobs J, Friedman P, Simpson E. The amino acid sequence of<br />

antistasin. J Biol Chem 1988; 263:10162–10167.<br />

44. Lauwereys M, Stanssens P, Lambier AM, Messens J, Dempsey E, Vlasuk GP. Ecotin as a potent<br />

factor Xa inhibitor. Thromb Haemostasis 1993; 69:864.<br />

45. Seymour JL, Lindquist RN, Dennis MA, Moffat B, Yansura D, Reilly D, Wessinger ME, Lazurus<br />

RA. Ecotin is a potent anticoagulant and reversible tightbinding inhibitor of factor Xa. Biochemistry<br />

1994; 33:3949–3958.<br />

46. McGrath ME, Gillmor SA, Fletterick RJ. Ecotin: lessons on survival in a proteasefilled world.<br />

Protein Sci 1995;4:141–148.<br />

47. Laskowski M, Kato I. Protein inhibitors of proteinases. Annu Rev Biochem 1980; 49:593–626.<br />

48. Schechter I, Berger A. On the size of the active site in proteases. I. Papain. Biochem Biophys Res<br />

Commun 1967; 27:157–162.<br />

49. Peterson LC. Progress in vascular biology, haemostasis and thrombosis. Abstracts, 1992<br />

Zimmerman Conference. San Diego, California, Feb. 27–29, 1993.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_292.html (1 of 2) [4/5/2004 5:16:15 PM]


Document<br />

50. Broze Jr GJ. Trends Cardiovasc Med 1994; 2:72.<br />

51. Hofmann KJ, Nutt EM, Dunwiddie CT. Site-directed mutagenesis of the leechderived factor Xa<br />

inhibitor antistasin. Biochem J 1992; 287:943.<br />

52. Schreuder H, Arkema A, deBoer B, Kalk K, Dijkema R, Mulders J, Theunissen H, Hol W.<br />

Crystallization and preliminary crystallographic analysis of antistasin, a leech-derived inhibitor of blood<br />

coagulation factor Xa. J Mol Biol 1993; 231:1137–1138.<br />

53. Broze Jr GJ, Warren LA, Novotny WF, Huguchi DA, Girard JJ, Miletich JP. The lipoproteinassociated<br />

coagulation inhibitor that inhibits the factor VII-tissue factor complex also inhibits factor Xa:<br />

insight into its possible mechanism of action. Blood 1984; 71:335–343.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_292.html (2 of 2) [4/5/2004 5:16:15 PM]


Document<br />

54. Broze Jr GJ, Girard TJ, Novotny WF. Regulation of coagulation <strong>by</strong> a multivalent Kunitz-type<br />

inhibitor. Biochemistry 1990; 29:7539.<br />

Page 293<br />

55. Dunwiddie CT, Vlasuk GP, Nutt EM. The hydrolysis and resynthesis of a single reactive site peptide<br />

bond in recombinant antistasin <strong>by</strong> coagulation factor Xa. Arch Biochem Biophys 1992; 294:647–653.<br />

56. Dunwiddie CT, Neeper MP, Nutt EM, Waxman L, Smith DE, Hoffman KJ, Lumma PK, Garsky<br />

VM, Vlasuk GP. Site-directed analysis of the functional domains in the factor Xa inhibitor tick<br />

anticoagulant peptide: identification of two distinct regions that constitute the enzyme recognition sites.<br />

Biochemistry 1992; 31:12126–12131.<br />

57. Lim-Wil<strong>by</strong> MSL, Hallenga K, DeMaeyer M, Lasters I, Vlasuk GP, Brunck TK. NMR structure<br />

determination of tick anticoagulant peptide (TAP). Protein Sci 1995; 4:178–186.<br />

58. Antuch W, Guntert P, Billeter M, Hawthorne T, Grossenbacher H, Wuthrich K. NMR solution<br />

structure of the recombinant tick anticoagulant protein (rTAP), a factor Xa inhibitor from the tick<br />

ornithodoros moubata. FEBS Lett 1994; 325:251–257.<br />

59. Rydel TJ, Tulinsky A, Bode W, Huber R. Refined structure of the hirudin-thrombin complex. J Mol<br />

Biol 1991;221:583–601.<br />

60. Katakura S-I, Nagahara T, Hara T, Iwamoto M. A novel Factor Xa inhibitor: structure-activity<br />

relationships and selectivity between Factor Xa and thrombin. Biochem Biophys Res Comm 1993;<br />

197:965–972.<br />

61. Hara T, Yokoyama A, Ishihara H, Yokoyama Y, Nagahara T, Iwamoto M. DX- 9065a, a new<br />

synthetic, potent anticoagulant and selective inhibitor for Factor Xa. Thrombosis and Haemostasis 1994;<br />

71:314–319.<br />

62. Nagahara T, Yokoyama Y, Inamura K, Katakura S-I, Komoriya S, Yamaguchi H, Hara T, Iwamoto<br />

M. J Med Chem 1994; 37:1200–1207.<br />

63. Nagahara T, Kanaya N, Inamura K, Yokoyama Y. Aromatic amidine derivatives and salts thereof.<br />

Eur Pat App 0-540-051-A1.<br />

64. Lin Z, Johnson ME. Proposed cation-π mediated binding <strong>by</strong> Factor Xa: a novel enzymatic<br />

mechanism for molecular recognition. FEBS Lett 1995; 370:1–5.<br />

65. Banner DW, Hadvary P. Crystallographic analysis of 3.0 Å resolution of the binding to human<br />

thrombin of four active-site directed inhibitors. J Biol Chem 1991; 266:20085–20093.<br />

66. Dougherty DA. Cation-π interactions in chemistry and biology: a new view of benzene, Phe, Tyr,<br />

and Trp. Science 1996; 271:163–167.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_293.html (1 of 2) [4/5/2004 5:17:35 PM]


Document<br />

67. Sturzebecher J, Sturzebecher U, Vieweg H, Wagner G, Hauptmann J, Markwardt F. Synthetic<br />

inhibitors of bovine factor Xa and thrombin comparison of their anticoagulant efficiency. Thrombosis<br />

Res 1989; 54:245–252.<br />

68. Ohta N, Brush M, Jacobs JW. Interaction of antistasin-related peptides with factor Xa: identification<br />

of a core inhibitory sequence. Thromb Haemostasis (GERMANY) 1994; 72:825–830.<br />

69. The preliminary modeled structures of the synthetic inhibitors described in this review were<br />

constructed and energy minimized <strong>by</strong> the author using HyperChem (1995, Hypercube, Inc., Release<br />

4.5). Docking of these inhibitors to the active site of Factor Xa was accomplished <strong>by</strong> the author using<br />

the x-ray coordinates of native Factor Xa [4] and INSIGHT II (Biosym Technologies, Inc.) and the<br />

approach outlined in Section 6.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_293.html (2 of 2) [4/5/2004 5:17:35 PM]


Document<br />

Page 294<br />

70. Seligmann B, Stringer SK, Ostrem JA, Al-Obeidi F, Wildgoose P, Walser A, Safar P, Safarova A,<br />

LoCascio A, Spoonamore J, Thorpe DS, Kasireddy P, Ashmore B, Strop P. SEL 2711: a specific, orally<br />

available, active-site inhibitor of Factor Xa discovered using synthetic combinatorial chemistry.<br />

Abstract, Sixth IBC International Symposium on Advances in Anticoagulants and Antithrombotics.<br />

Washington, D. C., Oct. 23–24, 1995.<br />

71. Al-obeidi F, Lebl M, Safar P, Stierandova A, Strop P, Walser A. Factor Xa inhibitors, Patent Appl.<br />

WO 95/29189; 1995.<br />

72. Lewis SD, Ng AS, Balwin JJ, Fusetani N, Naylor AM, Shafer JA. Inhibition of thrombin and other<br />

trypsin-like serine proteinases <strong>by</strong> cyclotheonamide A Thrombosis Research 1993; 70:173–190.<br />

73. Lee AY, Hagihara M, Karmacharya R, Albers MW, Schreiber, SL, Clardy J. Atomic structure of the<br />

trypsin-cyclotheonamide A complex: lessons for the design of serine protease inhibitors. J Am Chem<br />

Soc 1993; 115:12619.<br />

74. Marynoff BE, Qui X, Padmanabhan KP, Tulinsky A, Almond Jr HR, Andrade- Gordon P, Greco<br />

MN, Kauffman JA, Nicolaou KC, Liu A, Brungs PH, Fusetani N. Proc Natl Acad Sci USA 1993;<br />

90:8048.<br />

75. Ripka W, Brunck T, Stanssens P, LaRoche Y, Lauwereys M, Lambeir A-M, Lasters I, DeMaeyer M,<br />

Vlasuk G, Levy O, Miller T, Webb T, Tamura S, Pearson D. Strategies in the design of inhibitors of<br />

serine proteases of the coagulation cascade—factor Xa. Eur J Med Chem 1995; 30 (Suppl):88s–100s.<br />

76. Lasters I, DeMaeyer M, Ripka W. Bovine pancreatic trypsin inhibitor derived inhibitors of Factor<br />

Xa. Pat Appl WO 94/01461; 1994.<br />

77. McGrath ME, Erpel T, Bystroff C, Fletterick RJ. Macromolecular chelation as an improved<br />

mechanism of protease inhibition: structure of the ecotin-trypsin complex. EMBO 1994; 13:1502–1507.<br />

78.<br />

Desmet J,<br />

DeMaeyer<br />

M, Hazes<br />

B, Lasters<br />

I. Nature<br />

1992;<br />

356:539.<br />

79. Grasberger BL, Clore AM, Gronenborn GM. <strong>Structure</strong> 1994; 2:669–678.<br />

80. Huang K, Strynadka NCJ, Bernard VD, Peanasky RJ, James MNG. <strong>Structure</strong> 1994; 2:679–689.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_294.html (1 of 2) [4/5/2004 5:17:42 PM]


Document<br />

81. Brandstetter H, Kuhne A, Bode W, Huber R, von der Saal W, Wirthensohn K, Engh RA. J Biol<br />

Chem 1996; 47:29988–29992.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_294.html (2 of 2) [4/5/2004 5:17:42 PM]


Document<br />

12<br />

Polypeptide Modulators of Sodium Channel Function as a Basis for the<br />

Development of Novel Cardiac Stimulants<br />

Raymond S. Norton<br />

Biomolecular Research Institute, Parkville, Victoria, Australia<br />

I. Introduction<br />

Page 295<br />

Cardiovascular diseases remain one of the major causes of premature death in western societies. Chronic<br />

congestive heart failure (CHF) in particular is a common disease with a poor prognosis, median survival<br />

times after the onset of heart failure being 1.7 years in men and 3.2 years in women [1]. Current<br />

treatment relies on diuretics to reduce fluid volume, vasodilators to decrease the work load of the heart,<br />

and positive inotropic agents to increase cardiac contractility [2]. The most commonly prescribed of the<br />

positive inotropes is the cardiac glycoside digoxin (Figure 1) [3]. Although this drug has been in<br />

therapeutic use for over two hundred years, its efficacy in patients with a sinus rhythm has remained<br />

controversial, and evidence for its beneficial effects is quite recent [3–5]. It is also possible that these<br />

beneficial effects are not due solely to the positive inotropic activity of digoxin and that its<br />

neurohormonal effects may also be important [2, 5–7] Nevertheless, digoxin remains a widely used drug<br />

[3] and it follows that a suitable replacement or adjunct would find access to a significant market<br />

worldwide.<br />

The incentive to develop such a replacement follows from the low therapeutic index of digoxin [8,9] and<br />

the relatively common occurrence of side effects due to digitalis toxicity. In the 1960s and 1970s,<br />

20–30% of patients receiving digitalis experienced serious toxicity and about one quarter of this group<br />

died [6, 10]. Digitalis toxicity is manifest in CNS side-effects such as<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_295.html [4/5/2004 5:17:59 PM]


Document<br />

Figure 1<br />

<strong>Structure</strong>s of the positive inotropes digoxin [3,4], DPI 201-106 [15], and BDF 9148<br />

[15–17]. In digoxin the R group is (O-2,6-dideoxy-β-D-ribo-hexopyranosyl-(1 rarrow.gif 4)-O-2,<br />

6-dideoxy-β-D-ribo-hexopyranosyl-(1 rarrow.gif 4)-2, 6-dideoxy-β-D-ribo-hexopyranosyl)oxy. In<br />

DPI 201-106 the configuration at the hydroxyl-bearing carbon influences cardiac<br />

activity.<br />

Page 296<br />

fatigue, visual disturbances, and anorexia, and in cardiac side-effects that depend on the nature and<br />

extent of the underlying heart disease [3]. Careful monitoring of digoxin serum levels and bioavailability<br />

have reduced the incidence of digitalis toxicity [3] and the recent introduction of digoxin-binding<br />

antibodies or antibody fragments has provided an effective means of treating severe digitalis toxicity<br />

[3,7]. Nevertheless the quest continues for a substitute for the cardiac glycosides in the treatment of<br />

chronic CHF.<br />

Positive inotropic compounds can be classified into three groups: cAMP generators, intracellular<br />

calcium regulators, and modulators of ion channels or pumps [11]. The cAMP generators such as<br />

dopamine, dobutamine, and milrinone (a phosphodiesterase inhibitor) may worsen ischemia, cause<br />

arrhythmias, and increase mortality [2,6]. Intracellular calcium modulators have not reached clinical use,<br />

possibly because of additional effects such as vasoconstriction,<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_296.html (1 of 2) [4/5/2004 5:18:07 PM]


Document<br />

Page 297<br />

whereas, calcium sensitizers such as EMD 57033 may be useful positive inotropic compounds, even in<br />

the diseased myocardium [12].<br />

Ion channel modulation represents another approach to positive inotropy [13]. Sodium channel<br />

modulators increase Na + influx and prolong the plateau phase of the action potential; sodium/calcium<br />

exchange then leads to an increase in the level of calcium available to the contractile elements, thus<br />

increasing the force of cardiac contraction [13,14]. Synthetic compounds such as DPI 201-106 and BDF<br />

9148 (Figure 1) increase the mean open time of the sodium channel <strong>by</strong> inhibiting channel inactivation<br />

[15]. Importantly, BDF 9148 remains an effective positive inotropic compound even in severely failing<br />

human myocardium [16] and in rat models of cardiovascular disease [17]. Modulators of calcium and<br />

potassium channel activities also function as positive inotropes [13], but in the remainder of this article<br />

we shall focus on sodium channel modulators.<br />

II. The Anthopleurins<br />

Two decades ago “drugs from the sea” were the subject of high expectations and a good deal of effort in<br />

various centers around the world. The number of therapeutically useful compounds to have emerged<br />

from that effort has been rather limited, but with the advent of high-throughput screening it is likely that<br />

useful new leads will be found, even from species investigated previously. Notwithstanding, some<br />

valuable leads did emerge from work carried out in the 1970s, amongst which were the polypeptide<br />

cardiac stimulants known as the anthopleurins. These were isolated from sea anemones, where they are<br />

components of the animal's venom and are believed to have a function in defense and the capture of<br />

prey. The work that led to the isolation and characterization of these and related polypeptides from sea<br />

anemones is covered in earlier reviews [18,19] and will not be reiterated here.<br />

The best characterized of the anthopleurins is anthopleurin-A (AP-A), which was isolated from the<br />

northern Pacific sea anemone Anthopleura xan- thogrammica and consists of 49 residues cross-linked<br />

<strong>by</strong> three disulfide bonds [18,20]. It is active as a cardiac stimulant at nanomolar concentrations in vitro,<br />

making it some 200 fold more potent on a molar basis that digoxin. Its positive inotropic activity is not<br />

associated with any significant effects on heart rate or blood pressure [21], and in conscious dogs its<br />

therapeutic index is 7.5, which is about three-fold higher than that of digoxin [8]. Anthopleurin-A is<br />

active under conditions of stress and hypocalcaemia [18,22], as well as in ischemic myocardium where<br />

many other positive inotropes give equivocal results [23]. The profile of activity for AP-A suggests that<br />

it is a potentially valuable lead in the development of an alternative positive inotrope to digoxin<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_297.html [4/5/2004 5:18:14 PM]


Document<br />

for use in the treatment of chronic CHF. This chapter describes how this development is being tackled<br />

using the approach of structure-<strong>based</strong> drug design.<br />

A. Related Sea Anemone Toxins<br />

Page 298<br />

Anthopleurin-A is a member of a family of sea anemone polypeptides [24] (Figure 2) that is steadily<br />

increasing in number. These polypeptides have been classified into two groups, designated Types 1 and<br />

2 [27], which are similar with respect to the locations of their disulfide bridges and a number of residues<br />

thought to play a role in biological activity or maintenance of the tertiary structure [24,27], but are<br />

distinguishable on the basis of sequence similarity (>>60% within each type but


Document<br />

Figure 2<br />

Amino acid sequences of Type 1 sea anemone polypeptides. The residue numbering system is <strong>based</strong> on the sequences of AP-A and<br />

AP-B. Literature references are from Norton [24], except in the case of recently published sequences for Bc III [25], Bg II and Bg III<br />

[26]. Identical residues are shaded in grey and conserved residues are boxed. The sequences of Bg II and Bg III around residue 28<br />

could also be aligned to bring the Gly-Cys sequence into register with the other toxins if the Arg were treated as an insertion. Toxins are<br />

named in this figure according to their genus and species; common names relevant to the text are Ax I = AP-A; Ax II = AP-B; As Ia =<br />

ATX Ia.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_299.html [4/5/2004 5:19:48 PM]<br />

Page 299


Document<br />

Page 300<br />

polypeptides BDS I and II [40,41], which were claimed to have antihypertensive and antiviral activity,<br />

and the α-scorpion toxins [24,27,42]. As discussed below, we may expect to obtain some useful<br />

information from comparisons of the molecular surfaces of these different classes of polypeptides, but<br />

the utility of this approach depends on the extent to which their binding sites are identical and not just<br />

partially overlapping, as well as the issue of channel sub-type specificity of the different toxins.<br />

The synthetic agent DPI 201-106 has been evaluated extensively as a potential replacement for digoxin<br />

[43]. It is a potent positive inotrope that also acts <strong>by</strong> delaying inactivation of the sodium channel, but its<br />

binding site appears to be distinct from that of ATX II [44]. Moreover, it exerts antihypertensive and<br />

local anesthetic effects and may also antagonise the calcium channel [43]. At present we are not aware<br />

of any low molecular mass compound that binds to the same site as the anthopleurins. This offers the<br />

prospect that a mimetic <strong>based</strong> on the anthopleurins might have a pharmacological profile distinct from<br />

other positive inotropes.<br />

The structure of the receptor for the anthopleurins, the α-subunit of the voltage-gated sodium channel, is<br />

known only in schematic form [35–37]. As illustrated in Figure 3, it contains four homologous domains,<br />

each consisting of six transmembrane regions (assumed to be helices) designated S1 to S6. The S4<br />

segments are thought to act as the voltage sensors of the channel. All four<br />

Figure 3<br />

Schematic of the α-subunit of the voltage-gated sodium channel, <strong>based</strong> on its amino<br />

acid sequence [35–37]. The transmembrane segments S1–S6 in each domain are<br />

thought to form helices—with the positively charged S4 segment acting as a<br />

voltage sensor—and the S5–S6 loop of each domain is thought to contribute to the<br />

transmem-brane pore. Site 3 includes regions of the S5–S6 loops of domains I<br />

and IV, and the inactivation gate (h) is located on the intracellular segment linking<br />

domains III and IV.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_300.html (1 of 2) [4/5/2004 5:20:56 PM]


Document<br />

Page 301<br />

domains contribute to formation of the transmembrane pore, which is believed to be lined <strong>by</strong> short<br />

segments from the loops linking S5 and S6 in each domain. Binding site 3 is located on the extracellular<br />

surface of the channel and involves regions of the loops between S5 and S6 in domains I and IV [37,45].<br />

The binding sites for several modulators of sodium-channel activity, including the blocker tetrodotoxin,<br />

have been mapped quite precisely onto the sequence (and thus the structural model of the channel) <strong>by</strong><br />

combining information derived from comparisons of naturally occurring variants of the channel and<br />

from site-directed mutagenesis [37]. A similar approach could be applied to the definition of site 3, but<br />

in the absence of a crystal structure for the α-subunit the three- dimensional structure of this site will<br />

remain speculative. Therefore current attempts to design a low molecular mass analogue of the<br />

anthopleurins must be <strong>based</strong> on the structure of the ligand rather than the receptor.<br />

C. Development of a New Positive Inotrope<br />

Being polypeptides, the anthopleurins have limited therapeutic potential in their own right, as they are<br />

not active following oral administration and are antigenic in experimental animals [18]. Recent advances<br />

in the field of peptide mimetics, however, lend credence to the concept of harnessing the favorable<br />

cardiotonic properties of the anthopleurins in a low molecular mass, nonpeptide, synthetic compound.<br />

The goal of such a development would be to retain the activity of the parent polypeptides in a molecule<br />

that had good bioavailability and was not antigenic. Ideally, it might also be possible to increase the<br />

cardiac selectivity of such a compound.<br />

In order to achieve this goal, a knowledge of the three-dimensional structures of the anthopleurins and<br />

their structure-function relationships is essential. Significant progress has been made towards these goals<br />

over the past few years and we are now in a position to commence analogue design and synthesis. The<br />

following sections in this chapter summarize our knowledge of the cardioactive pharmacophore of the<br />

anthopleurins and the prospects for mimicking this in a nonpeptide moiety. Aspects of the<br />

pharmacological profile which such a compound would need to display in order to be useful in CHF<br />

therapy are also discussed.<br />

III. 3D <strong>Structure</strong><br />

The structures in aqueous solution of both AP-A [46] and AP-B [47] have been solved using highresolution<br />

1H NMR data. <strong>Structure</strong>s have also been determined for the Type 1 toxin ATX Ia [48] and the<br />

Type 2 toxin Sh I [49,50] from NMR data. The main secondary structure element in each of these<br />

structures is a<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_301.html [4/5/2004 5:21:01 PM]


Document<br />

Figure 4<br />

Richardson-style diagram of the polypeptide backbone of the<br />

individual structure of AP-A [46] that is closest to the average over the<br />

whole molecule. The locations of the sulfurs in the three disulfide bonds<br />

(4–46, 6–36, and 29–47) are shown in CPK format. The locations of<br />

reverse turns found in more than half the NMR-derived structures<br />

(6–9, 25–28, and 30–33) are indicated <strong>by</strong> darker backbone shading.<br />

The diagram was generated using MOLSCRIPT [51].<br />

Page 302<br />

four-stranded, antiparallel β-sheet linked <strong>by</strong> three loops, as illustrated for AP-A in Figure 4. The first of<br />

these loops, spanning residues 8–16 in AP-A and 8–17 in AP-B, is the largest and least well defined in<br />

solution (Figure 5), although it contains several residues that are essential for activity, as described in the<br />

next section.<br />

Differences between the structures of AP-A, ATX Ia, and Sh I have been noted [46] but the overall<br />

picture that emerges is one of similar backbone folds for all three molecules, making it likely that<br />

differences among the potencies and species-specificities of these toxins are due to the presence or<br />

absence of particular side chains rather than significant structural differences. Given that ATX Ia and Sh<br />

I have weak or negligible activities on mammalian nerve and heart tissue [24,27], we have focused on<br />

the anthopleurins in an effort to define the structure of the cardioactive pharmacophore of the sea<br />

anemone polypeptides. There are several challenges in this endeavor. One is that the structures are <strong>based</strong><br />

on NMR data that, because of the paucity of NOE restraints for surface residues compared with those in<br />

the core of the structure, do not define the locations of the solvent-exposed side chains very precisely<br />

(although it should be borne in mind that this may be a more accurate picture of the actual structure in<br />

solution than one in which the side chains are fixed in a single<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_302.html (1 of 2) [4/5/2004 5:21:58 PM]


Document<br />

Figure 5<br />

Stereo views of 20 structures of AP-B [47] superimposed over the backbone heavy<br />

atoms (N, C α , C) of residues 2–7 and 17–49. The three disulfide bonds are shown in<br />

lighter shading. The lower view is related to the upper one <strong>by</strong> an approximately 180°<br />

rotation about the vertical axis.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_303.html (1 of 2) [4/5/2004 5:23:22 PM]<br />

Page 303


Document<br />

Page 304<br />

location). Another is that both the backbone and the side chains of residues 8– 16 are less well defined<br />

than the bulk of the structure due to a lack of medium- and long-range NMR restraints between residues<br />

in this loop and the rest of the molecule, as illustrated in Figure 5. These problems are exacerbated in the<br />

case of AP-A and AP-B <strong>by</strong> the presence of multiple conformers in solution, one cause of which is cistrans<br />

isomerism about the Gly40-Pro41 peptide bond [52]. The additional peak overlap caused <strong>by</strong> these<br />

conformers limited the number of NOE restraints that could be obtained from the spectra. Finally, the<br />

structures available at present are for the free ligand and we have no information on how much these<br />

structures might change upon binding to the sodium channel.<br />

One way of addressing the issue of a lack of precision in the locations of functionally important side<br />

chains is to determine the range of conformational space available to them in different ligands. To this<br />

end, we undertook a detailed comparison of the structures of AP-A and AP-B in solution [47]. This<br />

proved to be a useful exercise both in terms of defining the positions of side chains known to be<br />

important for cardiotonic activity and identifying neighboring residues which might also be involved<br />

[47].<br />

Models have been described in the literature for AP-B [53] and Bunodosoma granulifera toxins II (Bg<br />

II) [26]. The AP-B model was derived from the structure of Sh I using energy minimization and the Bg<br />

II model from that of BDS I using energy minimization and 10 ps of dynamics. In both cases the<br />

calculations appear to have been carried out for the molecule in vacuo without the use of a distancedependent<br />

dielectric, under which conditions the positions of the charged side chains on the surface are<br />

likely to be distorted. Visual comparison of the model of AP-B [53] with the experimentally determined<br />

solution structure [47] indicates significant differences in side-chain orientations.<br />

IV. Residues Essential For Cardiotonic Activity<br />

Information about which residues are essential for the cardiac stimulatory activity of the sea anemone<br />

toxins has been obtained from selective chemical modification and proteolysis studies, comparisons<br />

among naturally occurring sequences, and, most recently, site-directed mutagenesis. Although there are<br />

some discrepancies among the inferences drawn from different studies and different techniques, a<br />

consensus is emerging regarding the location of the cardioactive pharmacophore. Ideally, only effects on<br />

cardiac tissue should be considered, but doing so would exclude some useful data on activity against<br />

mammalian nerve preparations. However, data obtained on the Type 2 toxins or on the activity of Type<br />

1 toxins on nonmammalian tissues will not be discussed in detail.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_304.html [4/5/2004 5:23:24 PM]


Document<br />

A. Chemical Modification<br />

Page 305<br />

A number of chemical modification studies have been carried out on the sea anemone toxins. As<br />

discussed previously [24], the results of these studies have to be interpreted with some caution because<br />

of less than rigorous characterization of the reaction products in many cases. Nevertheless, in the Type 1<br />

toxins it appears that one or both of the Asp7 and Asp9 carboxylates are required for activity, as well as<br />

one or both of the Lys37 and Lys48 ε-ammonium groups [24,27,54]. The C-terminal carboxylate<br />

appears not to be essential, whereas the N-terminal ammonium appears to have some role, although the<br />

various studies give a confusing view of its importance. There are also conflicting data on the<br />

importance of His34 and His39 but it seems that at least one of them might be important. Both are<br />

located in the vicinity of other residues found to be necessary for activity, but His39 is in close contact<br />

with Asp7 and Lys37 and on this basis may be expected to be the more important.<br />

Although the available evidence points to a role for one or both of the Asp7 and Asp9 carboxylates in<br />

cardiotonic activity, it has not been established that either residue makes contact with the sodium<br />

channel. As indicated above, the carboxylate of Asp9 participates in a hydrogen bond to the backbone<br />

amide of Cys6, so its role may be structural. The carboxylate of Asp7 is close to the side chains of<br />

Lys37 and His39 and is exposed to the solvent, making it a more likely candidate for direct interactions<br />

with the sodium channel. The only evidence for its importance, however, is indirect, coming from the<br />

observation that its replacement <strong>by</strong> Asn in synthetic Sh I abolished toxicity to crabs [55].<br />

Considerable confusion has surrounded the role of Arg14, which is conserved throughout the Type 1 and<br />

Type 2 toxins. A recent study has shown, however, that modification of Arg14 in AP-A with 1,2cyclohexanedione<br />

under conditions where the positive charge is maintained did not affect positive<br />

inotropic activity [54]. This study also showed indirectly that any contact the Arg14 side chain makes<br />

with the sodium channel must be relatively loose: although the adduct is active, it is no longer<br />

susceptible to tryptic proteolysis, indicating that the modified side chain cannot be accommodated in the<br />

substrate binding site of the protease. The conclusion that the positive charge on Arg14, but not its exact<br />

spatial location, might be important for activity is consistent with the results of site-directed mutagenesis<br />

experiments discussed below.<br />

B. Selective Proteolysis<br />

When AP-A was treated with trypsin only the Arg14 to Gly15 peptide bond was cleaved [56]. The<br />

resulting derivative lacked cardiotonic activity but its binding affinity for the rat brain sodium channel<br />

was reduced <strong>by</strong> less than an order of magnitude (Llewellyn LE et al., unpublished results). Its overall<br />

structure, as<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_305.html [4/5/2004 5:23:26 PM]


Document<br />

Page 306<br />

monitored <strong>by</strong> NMR, was unchanged, although it should be noted that the structure of the loop containing<br />

Arg14 was not well defined in either native or trypsinised AP-A. That the backbone structure was<br />

largely unaffected was confirmed <strong>by</strong> the observation that the NH and CαH chemical shifts were<br />

unaltered except in the immediate vicinity of the cleavage (which is the site of a reverse turn in AP-A<br />

[46]) and the N-terminal regions of the loop and the second strand of the sheet. It is possible that<br />

perturbations of functionally important groups near the start of the loop may be responsible for the lack<br />

of activity of the trypsinised derivative, and now that a high-resolution structure is available for AP-A a<br />

more detailed comparison with the cleavage product would be useful. Other possible explanations for<br />

the lack of activity are that the position of the Arg14 side chain relative to other key residues in the<br />

molecule is more important than suggested <strong>by</strong> chemical modification and site-directed mutagenesis data,<br />

or that the introduction of additional charges associated with the new termini affects activity.<br />

Endoproteinase LysC cleaved AP-A between Lys37 and Ala38 to yield a derivative with cardiotonic<br />

activity an order of magnitude lower than that of the parent molecule (Monks SA and Norton RS,<br />

unpublished results). This reduction in activity could be a consequence of local conformational<br />

perturbations. Treatment of AP-B with carboxypeptidase B removed Lys49, resulting in only a two-fold<br />

reduction in cardiotonic activity (Monks SA and Norton RS, unpublished results).<br />

C. Sequence Comparisons<br />

Among the Type 1 toxins shown in Figure 2, the Bc and Bg toxins (from the genus Bunodosuma) form a<br />

subgroup with characteristic differences from the Anemonia and Anthopleura toxins at residues 5, 12–13<br />

and 37–42. In addition, the Bg toxins also have Asp7 rarrow.gif Lys and Gly27 rarrow.gif Arg<br />

substitutions. A potent toxin in mice [26], the cardiac stimulatory activity of Bg II has not been reported.<br />

The potent activity of Bg II was ascribed to its abundance of positive charges [26]. Ignoring the<br />

histidines, which at least in AP-A and ATX II would be predominantly in their neutral forms at<br />

physiological pH [57], Bg II has six positively charged side chains and only one negatively charged side<br />

chain, whereas, AP-A has three and two respectively, and AP-B has five and two. As discussed below,<br />

however, it may be that an abundance of positive charge is associated with a lack of discrimination<br />

between the neuronal and cardiac sodium channels, as found for the scorpion α-toxins.<br />

Comparison of the activities of Af I and Af II is useful because of their close similarity. Differences<br />

between Af I and Af II are Ala3 rarrow.gif Pro, Asn12 rarrow.gif Ser, Thr21 rarrow.gif Ile, and an<br />

additional Gly at the N-terminus. Inspection of the sequences in Figures 2 shows that the identity of<br />

residue 3 correlates with that of residue 21, Ala or Ser at 3 co-occurring with Thr21, and Pro 3 with<br />

Ile21<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_306.html [4/5/2004 5:23:44 PM]


Document<br />

Page 307<br />

(the exception is ATX Ib [19,24], which has Pro 3 and Thr21, but it is not known if this has the same<br />

local structure). Residues 3 and 21 are juxtaposed in the β sheet [46–48] and the effect of switching<br />

Ser3/Thr21 to Pro3/Ile21 can be assessed <strong>by</strong> comparing the structures of AP-A [46] and AP-B [47]. It<br />

appears that this leads to a distortion of the sheet between the second and third residues but causes no<br />

significant perturbations to the overall structure. In the calculated structures of AP-A a Va12 NH<br />

rarrow.gif Leu22 CO hydrogen bond was found but this was not present in the AP-B structures,<br />

presumably as a result of the local differences. Comparing Af I and Af II, the combination of the<br />

Ala3/Thr21 to Pro3/Ile21 switch plus the addition of an extra Gly at the N-terminus of Af II would be<br />

expected to alter the local conformation at the N-terminus. The fact that the cardiac stimulatory activities<br />

of the two are the same [58] therefore implies that the N-terminus is not important functionally. This<br />

inference is conditional on the effect on activity of the only other difference between Af I and Af II,<br />

Asn12 rarrow.gif Ser, being minimal. In the Type 1 toxins, Ser is found more often than Asn at position<br />

12, while in the Type 2 toxins, replacement of Asn12 <strong>by</strong> Tyr increases toxicity in mice <strong>by</strong> a factor of<br />

two [24,27]. Thus, it is possible that the presence of Ser in Af II slightly favors cardiac activity while the<br />

changes at the N-terminus might reduce it slightly; the main conclusion to be drawn, however, is that<br />

neither change has a significant effect. The lack of effect of changes at the N-terminus is consistent with<br />

the results from expression of AP-B [59], where a Gly-Arg extension at the N-terminus had no effect on<br />

activity.<br />

At seven locations AP-B differs from AP-A. Two of these, Ser3 rarrow.gif Pro and Thr21 rarrow.gif<br />

Ile, were discussed above. Two more, Leu24 rarrow.gif Phe and Thr42 rarrow.gif Asn, are<br />

conservative changes located in or at the start of loops that are not in the immediate vicinity of the<br />

sodium channel binding surface and would not be expected to have a direct effect on activity. This<br />

leaves Ser12 rarrow.gif Arg, Val13 rarrow.gif Pro, and Gln49 rarrow.gif Lys, which do have<br />

functional significance, as discussed in the following section.<br />

It is important to note that there are several residues that are common to all of the long toxins and serve<br />

to maintain the biologically active conformations of these molecules. The clearest examples of residues<br />

in this category are the six half-cystines, although we believe that the Gly10-Pro11 sequence may<br />

influence the structure and flexibility of the Arg14-containing loop [46] (at the other end of this loop<br />

Thr17 and/or Ser19-Gly20 may also be important). Also conserved, Trp33 is probably important<br />

structurally even though its surroundings are different in the Type 1 and Type 2 toxins.<br />

It is also interesting that in ATX Ia, which is a potent crustacean neurotoxin but a poor mammalian<br />

cardiac stimulant, Lys37 and both histidines are missing, suggesting that one or more of these side<br />

chains may be important in promoting specificity for the mammalian cardiac sodium channel.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_307.html [4/5/2004 5:23:45 PM]


Document<br />

D. Site-Directed Mutagenesis<br />

Page 308<br />

The complete synthesis of AP-A has been reported [60] but so far this approach has not been pursued to<br />

generate analogues. More productive has been the analysis of analogues produced <strong>by</strong> site-directed<br />

mutagenesis, following the successful cloning and expression of AP-B [59]. A series of single-site<br />

mutations [61,62] showed that R14A, K48A, and K49A had affinities for the cardiac sodium channel<br />

that were, respectively, only 3.2-, 2.9-, and 2.4-fold lower than that of native AP-B, while R12A had an<br />

8.5-fold lower affinity. These results suggested that amongst the cationic side chains only Arg12 was<br />

significant for activity. Recent data on double mutations [53] indicates that the situation is not quite that<br />

simple. For example, the R12S-R14Q double mutant had a 72-fold lower affinity for the neuronal<br />

sodium channel whereas the R12S and R14Q mutants individually had 5.9- and 2.4-fold lower affinities,<br />

respectively, which should combine to produce only a 14-fold effect. It appears that the presence of one<br />

cationic side chain in the Arg14 loop may be sufficient for activity and that its exact location can vary<br />

somewhat. It would be interesting to know if the same applies to the C-terminus, where Lys48 and<br />

Lys49 may be able to compensate for one another. However, the proposal that the cationic side chains of<br />

residues 12, 14, and 49 form a cluster [53] is not consistent with the solution structure of AP-B [47].<br />

A similar situation appears to exist in the ω-conotoxins, which possess 5–7 net positive charges.<br />

Substitution of individual cationic groups had a relatively minor effect on affinity for the voltage-gated<br />

calcium channel, whereas replacement of several had a significant effect, greater than that expected from<br />

the sum of the individual effects [63,64].<br />

A further outcome of analyses of the double mutants was that the cationic side chains at positions 12 and<br />

49 in AP-B seem to favor binding to the neuronal over the cardiac sodium channel. Thus, the R12S-<br />

R49Q double mutant, in which residues 12 and 49 in AP-A, had a 37-fold lower affinity for the neuronal<br />

channel but only a 5-fold lower affinity for the cardiac channel relative to native AP-B [53]. In fact, the<br />

affinity of this double mutant for the cardiac channel was lower than that of AP-A, implying that one or<br />

more of the other five differences between AP-A and AP-B might decrease affinity for the cardiac<br />

channel. It seems, therefore, that while AP-A is slightly less potent than AP-B on cardiac channels, it<br />

has greater selectivity for the cardiac channel when measured in sodium flux experiments. In voltageclamp<br />

experiments AP-A and AP-B favor the cardiac channel to similar degrees [53].<br />

V. Other Ligands for Site 3 on the Sodium Channel<br />

A. Sea Anemone Toxins<br />

Apart from the “long” sea anemone polypeptides that are the main focus of our interest, there are two<br />

other classes of anemone polypeptides that bind at or near<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_308.html [4/5/2004 5:23:47 PM]


Document<br />

Page 309<br />

site 3 on the sodium channel. The first of these are the short anemone polypeptides [19,24] such as ATX<br />

III [38] and PaTX [39], which are neurotoxic to crustacea. The polypeptide ATX III, which consists of<br />

27 residues cross-linked <strong>by</strong> three disulfide bonds, and PaTX, which has 31 residues and four disulfides,<br />

can be aligned such that 16 residues are identical. The three disulfides in ATX III are linked in a<br />

1–5/2–4/3–6 pattern in the same way as in the long polypeptides [65], but the only other similarity at the<br />

level of primary structure is a GCPXG sequence corresponding to residues 28–32 of AP-A. Although<br />

neurotoxic to crustacea, ATX III is inactive as a positive inotrope [28], suggesting that it possesses an<br />

appropriate structural scaffold to interact with site 3 but lacks key side chains required for interaction<br />

with the cardiac channel. Nevertheless, the smaller size of ATX III makes it an attractive candidate for<br />

further study, with the aim of engineering into it the ability to bind to the cardiac channel. The welldefined<br />

solution structure for this toxin [65] provides and essential basis for such an effort.<br />

The Anemonia sulcata polypeptides BDS I and II [40], which were claimed to have antihypertensive and<br />

antiviral activity, also bind to site 3 on neuronal sodium channels and have weak negative inotropic<br />

activity [41]. The points of similarity and difference between the solution structures of BDS I [40] and<br />

the long anemone polypeptides have been discussed previously [40,41] and will not be reiterated here;<br />

suffice to say that the overall folds are similar but the Arg14 loop in the long polypeptides is truncated in<br />

BDS I and the molecule lacks several residues that have been shown to be important for activity.<br />

B. Scorpion Toxins<br />

The scorpion α-toxins have been shown to bind to site 3 on the voltage-gated sodium channel<br />

[24,27,42]. These polypeptides contain up to 70 residues crosslinked <strong>by</strong> four disulfide bonds, but show<br />

no sequence similarity to the anemone polypeptides. Possible structural similarities have been discussed<br />

[24], and in a theoretical model of the anemone toxin Bg II, some of the cationic residues were in similar<br />

locations to those in the crystal structure of the scorpion toxin Aah II [26].<br />

It is clear that positively charged residues play an important role in the interactions of sea anemone<br />

toxins and scorpion toxins with site 3 on the sodium channel (as indeed they do with other polypeptide<br />

toxins binding to other ion channels) but this role may be relatively more important for the TTXsensitive<br />

sodium channel in nerve and muscle than for the TTX-insensitive channel of the heart. For<br />

example, Bg II, which is more positively charged than AP-A and AP-B (see above), has a higher affinity<br />

for neuronal sodium channels [26], and replacement of Arg12 and Lys49 in AP-B with uncharged<br />

residues favors its binding to the cardiac channel [53]. Similarly, the scorpion α-toxins bind more tightly<br />

to the neuronal channel than the anemone toxins but, as with the Type 2<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_309.html [4/5/2004 5:23:49 PM]


Document<br />

Page 310<br />

anemone toxins, do not discriminate between TTX-sensitive and TTX-insensitive channels. In fact the<br />

scorpion toxins are less potent cardiac stimulants than the Type 1 anemone toxins [29]. Thus, while it<br />

will be interesting to compare the sodium channel binding surfaces of the two classes of toxin as these<br />

surfaces become better defined in each case, more useful input to the design of a positive inotrope acting<br />

at site 3 is likely to come from direct studies on the anthopleurins.<br />

VI. Defining the Cardioactive Pharmacophore<br />

Now that high-resolution structures are available for both AP-A [46] and AP-B [47], we can begin to<br />

interpret the results described above in terms of an emerging picture of the cardioactive pharmacophore<br />

of these molecules. It is encouraging that most of the residues that have been shown hitherto to be<br />

important for activity lie on one face of the structures, as shown in Figure 6. Of the residues highlighted<br />

in Figure 6a, evidence to support their role in activity on the neuronal or cardiac channels (or both) has<br />

come from chemical modification or site-directed mutagenesis studies except in the case of Asn35. The<br />

reason for including this residue is that in AP-B it is close to the Asp7/Lys37/His39 region and its side<br />

chain is hydrogen bonded to the backbone carbonyl of Lys37 [47]. Moreover, it is conserved throughout<br />

the Type 1 toxins (Figure 2).<br />

We anticipate that many of the residues highlighted in Figure 6 will participate in interactions with the<br />

sodium channel binding site. One residue which may not is Asp9, the side-chain carboxylate of which<br />

hydrogen bonds with the amide of Cys6 in AP-A and AP-B. Thus, it is possible that this residue has a<br />

“structural” role rather than a “functional” one. Other residues in the vicinity of the pharmacophore that<br />

may also have a structural role are Gly10 and Pro11, as discussed above. The side chain of Ser8 is<br />

exposed and on the same surface of the molecule, placing it in a position potentially to interact with the<br />

sodium channel; it is also conserved throughout the Type 1 toxins.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_310.html [4/5/2004 5:23:50 PM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_311.html (1 of 2) [4/5/2004 5:24:01 PM]<br />

Page 311


Document<br />

Figure 6<br />

(a) CPK representation of the individual structure of<br />

AP-B [47], which is closest to the average over the<br />

whole molecule, showing residues thought to contribute<br />

to the cardioactive pharmacophore. The surfaces of<br />

residues 7 and 9 are shaded black, those of 14, 37, 39, and<br />

48 dark grey, and that of 35 lighter grey. As discussed in<br />

the text, the primary functions of Asp9 (the side chain of<br />

which is hardly visible in this view) and Asn35 may be to<br />

maintain the local structure in an active conformation,<br />

but it cannot be excluded that they also interact directly<br />

with the sodium channel. In AP-B the cationic side<br />

chains of Arg12 and Lys49 are also important, but it<br />

appears that their roles can be compensated for <strong>by</strong> near<strong>by</strong><br />

cationic side chains (Arg14 and Lys48, respectively) and<br />

that they favor binding to the neuronal sodium channel<br />

rather than the cardiac channel [53]. (b) Connolly surface<br />

of AP-B in the same orientation as in part a, with the<br />

charged residues Asp7, Asp9, Arg14, Lys37, and Lys48<br />

highlighted. This figure was generated using Insight<br />

(Biosym Technologies).<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_311.html (2 of 2) [4/5/2004 5:24:01 PM]


Document<br />

Page 312<br />

In peptide—protein and protein—protein interactions the size of the buried surface area ranges from 400<br />

Å 2 to 1400 Å 2 [66]. In the potassium channel blocker charybdotoxin, 5–8 residues with surface areas of<br />

530–850 Å 2 were found to be essential for binding, depending on the type of potassium channel<br />

investigated [67,68]. In the calcium channel blocker ω-conotoxin GVIA, an alanine scan identified only<br />

two residues, Lys2 and Tyr13, that were important for activity [69]. It is likely, however, that the<br />

number of residues contributing to the binding surface is greater than this, particularly given the high<br />

affinity of this toxin for its receptor. If we consider a larger ligand such as human growth hormone, eight<br />

of the 31 residues in the growth hormone-receptor interface contribute 85% of the binding energy and<br />

more than half make no significant contribution to the affinity [70]. A similar result was found for a<br />

growth hormone-monoclonal antibody complex, where only five residues were critical for binding [71].<br />

In fact, the example of growth hormone may be more relevant to the anthopleurins than those of<br />

charybdotoxin and conotoxin, which function simply as ion channel blockers. Thus, we expect the<br />

essential residues in the anthopleurins to number between five and ten, a total somewhat less than<br />

previously anticipated [24].<br />

If we ignore Asp9 and assume that only one of Arg12 or Arg14 and one of Lys48 or Lys49 are<br />

necessary, then it is likely that the residues highlighted in Figure 6 make a significant contribution to the<br />

sodium channel binding surface of the anthopleurins. They span a larger area on the surface than the<br />

essential residues in charybdotoxin, but it is important to note that the conformation of the Arg14 loop in<br />

solution is not fixed and that conformations in which the Arg14 side chain is closer to the region around<br />

Asp7 could be more representative of the sodium-channel-bound structure. For example, the distance<br />

between the Arg14 guanidino group and the Asp7 carboxylate varies from 13 to 26 Å over the family of<br />

structures of AP-A and 10 to 22 Å in AP-B.<br />

By analogy with other protein—protein interactions, it is likely that the sodium channel binding surface<br />

of the anthopleurins will include side chains that mutagenesis studies will not identify as having a<br />

significant role in binding or activity. As indicated above, alanine scanning of residues in the human<br />

growth hormone—receptor interface indicated that less than a quarter of the contact residues provided<br />

most of the binding energy [70]. Thus, we believe that the residues identified above will contribute to<br />

the sodium channel binding surface of the anthopleurins <strong>by</strong> virtue of their location on the protein surface<br />

in the immediate vicinity of residues, which clearly are important for activity. Nevertheless, some of<br />

them may make only modest contributions to binding affinity and could be altered without destroying<br />

activity. Other residues, as yet unidentified, may also make significant contributions. By analogy with<br />

other protein—protein interactions characterized to date, it is reasonable to anticipate that some of these<br />

residues will be hydrophobic, in contrast to the charged<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_312.html [4/5/2004 5:24:03 PM]


Document<br />

Page 313<br />

residues that have been the main focus of attention hitherto. A key requirement now is to identify those<br />

residues that provide the majority of the binding energy and to differentiate these from others whose<br />

main role is to stabilize the binding residues in the active conformation.<br />

VII. Mimicking The Pharmacophore<br />

Significant progress has been made over the past few years in the field of peptide mimetics, although the<br />

most successful examples are those where a small peptide ligand or a linear segment of a larger protein<br />

has been the target [72–74]. An alternative approach to de novo design is to optimize a lead compound<br />

obtained <strong>by</strong> screening chemical libraries on the basis of a knowledge of the conformation of the<br />

polypeptide ligand, as in the case of the endothelin receptor antagonist SB 209670 [75].<br />

The task of mimicking a pharmacophore is simplified where the contributing residues are contiguous in<br />

the amino acid sequence. This is not the case in the anthopleurins, with residues from at least four<br />

different regions of the sequence contributing to affinity. In charybdotoxin the essential residues come<br />

from two or three regions of the sequence, depending on which potassium channel is considered, while<br />

in growth hormone, binding site I is comprised of residues from three different regions of the protein<br />

and site II from two regions. Mimicking the pharmacophore of the anthopleurins therefore represents a<br />

task at least as challenging as those presented <strong>by</strong> these two examples. Strategies for achieving this goal<br />

include de novo design, conformationally directed data base searching and screening chemical libraries<br />

(synthetic and naturally occurring) for leads, which could then be optimized on the basis of our<br />

knowledge of the structure. Our approach is <strong>based</strong> on the first two of these.<br />

Initial attempts to mimic the pharmacophore of AP-A were <strong>based</strong> on linear and disulfide-cyclized<br />

versions of the Arg14-containing loop [76]. At that time, our level of understanding of the<br />

pharmacophore was inadequate and it is clear in retrospect that not enough of the key elements were<br />

present. Nevertheless, conformational analysis of these peptides <strong>by</strong> NMR was useful in showing that<br />

they retained several elements of local structure observed in the corresponding region of the native<br />

protein, there<strong>by</strong> emphasizing the independence of this loop from the rest of the structure in solution.<br />

VIII. Conclusions<br />

In this chapter I have attempted to summarize the current state of our understanding of the structure and<br />

structure-function relationships of the type 1 sea<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_313.html [4/5/2004 5:24:04 PM]


Document<br />

Page 314<br />

anemone toxins and to indicate how the favorable cardiotonic properties of the anthopleurins might be<br />

mimicked in a low molecular weight analog.<br />

Progress in defining the cardiotonic pharmacophore has been hampered <strong>by</strong> difficulties in determining<br />

high-resolution structures of the anthopleurins in solution due to the presence of multiple conformers,<br />

and <strong>by</strong> uncertainties concerning the exact location of some of the key residues in the Arg14 loop. Both<br />

of these problems would have been alleviated <strong>by</strong> isotopic labeling of the molecules in a high-yield<br />

expression system, which would have allowed for better definition of the structures in solution. The<br />

question of how the structure in solution might change upon binding to the sodium channel remains<br />

open and is particularly relevant to residues in the Arg14-containing loop. Conformationally constrained<br />

analogs of the ligand offer one means of addressing this problem. The definitive solution would be<br />

provided <strong>by</strong> a high-resolution structure for the sodium channel, determined <strong>by</strong> x-ray or electron<br />

crystallography, but this will not be available in the near future. In the meantime, the approach of<br />

complementary mutagenesis (of both ligand and receptor), which has been very informative in defining<br />

the charybdotoxin-potassium channel interface [68,77], can be employed to produce a crude model of<br />

the binding site.<br />

Once a lead compound is obtained, further development will almost certainly be required to optimize<br />

properties such as bioavailability and stability in vivo. A key requirement will also be good selectivity<br />

for the cardiac over other sodium channels, but the results of mutagenesis studies carried out so far on<br />

the anthopleurins suggest that this should be achievable. Lead compounds will also have to be rigorously<br />

evaluated in terms of their effects on cardiac arrhythmias to ensure that they ameliorate rather than<br />

exacerbate this problem, especially in the failing heart. Finally, the possibility that the beneficial effects<br />

of positive inotropes in vivo may be the result of inotropic and noninotropic activities [6] would need to<br />

be evaluated for mimetics of the anthopleurins. These requirements notwithstanding, there is good<br />

reason to be optimistic that a mimetic of the anthopleurins can be developed and that it may have<br />

significant benefits in the treatment of congestive heart failure.<br />

Acknowledgments<br />

I am grateful to all the colleagues who have contributed to our sea anemone toxin work over the years,<br />

and, in particular, to Steve Monks, Paul Pallaghy, and Jane Tudor for assistance with the figures and for<br />

helpful discussions. I also thank Ken Blumenthal for communicating results prior to publication.<br />

Note Added in Proof<br />

Since completion of this chapter, two additional papers on site-directed mutations of anthopleurin-B<br />

have been published. In the first (Khera PK, Blumenthal<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_314.html [4/5/2004 5:24:06 PM]


Document<br />

KM. Importance of highly conserved anionic residues and electrostatic interactions in the activity and<br />

structure of the cardiotonic polypeptide anthopleurin B. Biochemistry 1996; 35:3503–3507), it was<br />

concluded that Asp7 may be important for folding, Asp9 may be important for protein folding and<br />

interaction with the sodium channel, and Lys37 for channel interaction. In the second (Dias-Kadambi<br />

BL, Drum CL, Hanck DA, Blumenthal KM. Leucine 18, a hydrophobic residue essential for high<br />

affinity binding of anthopleurin-B to the voltage-sensitive sodium channel. J Biol Chem 1996;<br />

271:9422–9428), Leu18 was shown to be important for binding, with several hundred fold losses in<br />

affinity being associated with its mutation to Val or Ala. This residue is adjacent to the surface<br />

highlighted in Figure 6.<br />

References<br />

2. van Zwieten PA. Pharmacotherapy of congestive heart failure. Currently used and experimental<br />

drugs. Pharmacy World and Science 1994; 16:234–242.<br />

Page 315<br />

1. Ho KKL,<br />

Anderson<br />

KM, Kannel<br />

WB,<br />

Grossman<br />

W, Levy D.<br />

Survival<br />

after the<br />

onset of<br />

congestive<br />

heart failure<br />

in<br />

Framingham<br />

heart study<br />

subjects.<br />

Circulation<br />

1993;<br />

88:107–115.<br />

3. Lewis RJ. Digitalis: a drug that refuses to die. Critical Care Medicine 1990; 18:S5–S13.<br />

4. DiBianco R, Shabetai R, Kostuk W, Moran J, Schlant RC, Wright R. A comparison of oral milrinone,<br />

digoxin, and their combination in the treatment of patients with chronic heart failure. N Engl J Med<br />

1989; 320:677–683.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_315.html (1 of 2) [4/5/2004 5:24:09 PM]


Document<br />

5. Packer M, Gheorghiade M, Young JB, Costantini PJ, Adams KF, Cody RJ, Smith LK, Van Voorhees<br />

L, Gourley LA, Jolly MK. Withdrawal of digoxin from patients with chronic heart failure treated with<br />

angiotensin-converting-enzyme inhibitors. N Engl J Med 1993; 329:1–7.<br />

6. Packer M. The development of positive inotropic agents for chronic heart failure: how have we gone<br />

astray? J Am Coll Cardiol 1993; 22(suppA):119A–126A.<br />

7. Lederer WJ, Hadley RW, Kir<strong>by</strong> MS, Eisner DA. Inotropic mechanisms in heart muscle: cardiotonic<br />

steroids—how do they work? In: Gwathmey JK, Briggs GM, Allen PD, eds. Heart failure. Basic science<br />

and clinical aspects. New York: Marcel Dekker, 1993; 349–365.<br />

8. Scriabine A, Van Arman CG, Morgan G, Morris AA, Bennett CD, Bohidar NR. Cardiotonic effects of<br />

anthopleurin-A, a polypeptide from a sea anemone. J Cardiovasc Pharmacol 1979; 1:571–583.<br />

9. Marsh JD, Smith TW. Epidemiology and general considerations of digitalis toxicity. In: Smith TW,<br />

ed. Digitalis glycosides. Orlando, Fla: Grune and Stratton, 1986: 217–225.<br />

10. Beller GA, Smith Tw, Abelmann WH, Haber E, Hood WB Jr. Digitalis intoxication: a prospective<br />

clinical study with serum level correlations. N Engl J Med 1971; 284:989–997.<br />

11. Feldman AM. Classification of positive inotropic agents. J Am Coll Cardiol 1993; 22:1223–1227.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_315.html (2 of 2) [4/5/2004 5:24:09 PM]


Document<br />

Page 316<br />

12. Nankervis R, Lues I, Brown L. Calcium sensitization as a positive inotropic mechanism in diseased<br />

rat and human heart. J Cardiovasc Pharmacol 1994; 24:612–617.<br />

13. Doggrell S, Hoey A, Brown L. Ion channel modulators as potential positive inotropic compounds for<br />

treatment of heart failure. Clin Exp Pharmacol Physiol 1988; 21:833–843.<br />

14. Briggs GM, Gwathmey JK. Role of the sodium channel in the development of force in<br />

myocardium. In: Gwathmey JK, Briggs GM, Allen PD, eds. Heart failure. Basic science and<br />

clinical aspects. New York: Marcel Dekker, 1993:597–612.<br />

15. Hoey A, Amos GJ, Wettwer E, Ravens U. Differential effects of BDF 9148 and DPI 201-106 on<br />

action potential and contractility in rat and guinea-pig myocardium. J Cardiovasc Pharmacol 1994;<br />

23:907–915.<br />

16. Schwinger RHG, Böhm M, Mittmann C, La Rosee K, Erdmann E. Evidence for a sustained<br />

effectiveness of sodium-channel activators in failing human myocardium. J Mol Cell Cardiol 1991;<br />

23:461–471.<br />

17. Hoey A, Nankervis R, Brown L. Positive inotropic responses of the sodium channel modulator BDF<br />

9148 in diseased rat myocardium. Clin Exp Pharmacol Physiol 1995; 22:418–422.<br />

18. Norton TR. Cardiotonic polypeptides from Anthopleura xanthogrammica (Brandt) and A.<br />

elegantissima (Brandt). Fed Proc 1981; 40:21–25.<br />

19. Beress L. Biologically active compounds from coelenterates. Pure Appl Chem 1982; 54:1981–1994.<br />

20. Tanaka M, Haniu M, Yasunobu KT, Norton TR. Amino acid sequence of the Anthopleura<br />

xanthogrammica heart stimulant anthopleurin-A. Biochemistry 1977; 16:204–208.<br />

21. Blair RW, Peterson DF, Bishop VS. The effects of anthopleurin-A on cardiac dynamics in conscious<br />

dogs. J Pharmacol Exp Ther 1978; 207:271–276.<br />

22. Kodama I, Toyama J, Shibata S, Norton TR. Electrical and mechanical effects of anthopleurin-A, a<br />

polypeptide from a sea anemone, on isolated rabbit ventricular muscle under conditions of hypoxia and<br />

glucose free medium. J. Cardiovasc Pharmacol 1981; 3:75–86.<br />

23. Gross GJ, Warltier DC, Hardman HF, Shibata S. Cardiotonic effects of anthopleurin-A (AP-A), a<br />

polypeptide from a sea anemone, in dogs with a coronary artery stenosis. Eur J Pharmacol 1985;<br />

110:271–276.<br />

24. Norton RS. <strong>Structure</strong> and structure-function relationships of sea anemone proteins that interact with<br />

the sodium channel. Toxicon 1991; 29:1051–1084.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_316.html (1 of 2) [4/5/2004 5:24:11 PM]


Document<br />

25. Malpezzi ELA, De Freitas JC, Muramoto K, Kamiya H. Characterization of peptides in sea anemone<br />

venom collected <strong>by</strong> a novel procedure. Toxicon 1993; 31:853–864.<br />

26. Loret EP, Menendez Soto del Valle R, Mansuelle P, Sampieri F, Rochat H. Positively charged amino<br />

acid residues located similarly in sea anemone and scorpion toxins. J Biol Chem 1994;<br />

269:16785–16788.<br />

27. Kem WR. Sea anemone toxins: structure and action. In: Hessinger D, Lenhoff H, eds. The Biology<br />

of Nematocysts. New York: Academic Press, 1988:375–405.<br />

28. Alsen C. Biological significance of peptides from Anemonia sulcata. Fed Proc 1983;42:101–108.<br />

29. Renaud J-F, Fosset M, Schweitz H, Lazdunski M. The interaction of polypeptide neurotoxins with<br />

tetrodotoxin-resistant Na + channels in mammalian cardiac cells.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_316.html (2 of 2) [4/5/2004 5:24:11 PM]


Document<br />

Correlation with inotropic and arrhythmic effects. Eur J Pharmacol 1986; 120:161–170.<br />

Page 317<br />

30. Schweitz H, Bidard J-N, Frelin C, Pauron D, Vijverberg HPM, Mahasneh DM, Lazdunski M,<br />

Vilbois F, Tsugita A. Purification, sequence and pharmacological properties of sea anemone toxins from<br />

Radianthus paumotensis. A new class of sea anemone toxins acting on the sodium channel.<br />

Biochemistry 1985; 24:3554–3561.<br />

31. Frelin C, Vigne P, Schweitz H, Lazdunski M. The interaction of sea anemone and scorpion<br />

neurotoxins with tetrodotoxin-resistant Na + channels in rat myoblasts. A comparison with Na + channels<br />

in other excitable and non-excitable cells. Mol Pharmacol 1984;26:70–74.<br />

32. Platou ES, Refsum H, Hotvedt R. Class III antiarrhythmic action linked with positive inotropy:<br />

antiarrhythmic, electrophysiological, and hemodynamic effects of the sea anemone polypeptide ATX II<br />

in the dog heart in situ. J Cardiovasc Pharmacol 1986; 8:459–465.<br />

33. Vaughan Williams EM. Classification of antidysrhythmic drugs. Pharmac Therap B 1975;<br />

1:115–138.<br />

34. Sasayama S. What do the newer inotropic drugs have to offer? Cardiovasc <strong>Drug</strong>s Ther 1992;<br />

6:15–18.<br />

35. Catterall WA. <strong>Structure</strong> and function of voltage-sensitive ion channels. Science 1988; 242:50–61.<br />

36. Wann KT. Neuronal sodium and potassium channels: structure and function. Br J Anaesthes 1993;<br />

71:2–14.<br />

37. Catterall WA. <strong>Structure</strong> and function of voltage-gated ion channels. Ann Rev Biochem 1995;<br />

64:493–531.<br />

38. Warashina A, Jiang Z-Y, Ogura T. Potential-dependent action of Anemonia sulcata toxins III and IV<br />

on sodium channels in crayfish giant axons. Eur J Physiol 1988; 411:88–93.<br />

39. Warashina A, Ogura T, Fujita S. Binding properties of sea anemone toxins to sodium channels in the<br />

crayfish giant axon. Comp Biochem Physiol 1988; 90C:351–358.<br />

40. Driscoll PC, Gronenborn AM, Beress L, Clore GM. Determination of the three-dimensional solution<br />

structure of the antihypertensive and antiviral protein BDS-1 from the sea anemone Anemonia sulcata: a<br />

study using nuclear magnetic resonance and hybrid distance geometry—dynamical simulated annealing.<br />

Biochemistry 1989; 28:2188–2198.<br />

41. Llewellyn LE, Norton RS. Binding of the sea anemone polypeptide BDS II to the voltage-gated<br />

sodium channel. Biochem Intl 1991; 24:937–946.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_317.html (1 of 2) [4/5/2004 5:24:13 PM]


Document<br />

42. Catterall WA, Beress L. Sea anemone toxin and scorpion toxin share a common receptor site<br />

associated with the action potential sodium ionophore. J Biol Chem 1978; 253:7393–7396.<br />

43. Scholtysik G. Cardiac Na + channel activation as a positive inotropic principle. J Cardiovasc<br />

Pharmacol 1989; 14 (Suppl 3):S24–S29.<br />

44. Scholtysik G, Quast U, Schaad A. Evidence for different receptor sites for the novel cardiotonic S-<br />

DPI 201-106, ATX II, and veratridine at the cardiac sodium channel. Eur J Pharmacol 1986;<br />

125:111–118.<br />

45. Thomsen WJ, Catterall WA. Localization of the receptor site for α-scorpion toxins <strong>by</strong> antibody<br />

mapping: implications for sodium channel topology. Proc Natl Acad Sci USA 1989; 86:10161–10165.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_317.html (2 of 2) [4/5/2004 5:24:13 PM]


Document<br />

46. Pallaghy PK, Scanlon MJ, Monks SA, Norton RS. Three-dimensional structure in solution of the<br />

polypeptide cardiac stimulant anthopleurin-A. Biochemistry 1995; 34:3782–3794.<br />

47. Monks SA, Pallaghy PK, Scanlon MJ, Norton RS. Solution structure of the cardiostimulant<br />

polypeptide anthopleurin-B and comparison with anthopleurin-A. <strong>Structure</strong> 1995; 3:791–803.<br />

Page 318<br />

48. Widmer H, Billeter M, W¨thrich K. The three-dimensional structure of the neurotoxin ATX Ia from<br />

Anemonia sulcata in aqueous solution <strong>by</strong> nuclear magnetic resonance spectroscopy. Proteins 1989;<br />

6:357–371.<br />

49. Fogh RH, Kem WR, Norton RS. Solution structure of neurotoxin I from the sea anemone<br />

Stichodactyla helianthus. A nuclear magnetic resonance, distance geometry and restrained molecular<br />

dynamics study. J Biol Chem 1990; 265:13016–13028.<br />

50. Wilcox GR, Fogh RH, Norton RS. Refined structure in solution of the sea anemone neurotoxin Sh I.<br />

J Biol Chem 1993; 268:24707–24719.<br />

51. Kraulis P. MOLSCRIPT: a program to produce both detailed and schematic plots of protein<br />

structures. J Appl Crystallogr 1991; 24:946–950.<br />

52. Scanlon MJ, Norton RS. Multiple conformations of the sea anemone polypeptide anthopleurin-A in<br />

solution. Protein Sci 1994; 3:1121–1124.<br />

53. Khera PK, Benzinger GR, Lipkind G, Drum CL, Hanck DA, Blumenthal KM. Multiple cationic<br />

residues of anthopleurin-B that determine high affinity and channel isoform discrimination.<br />

Biochemistry 1995; 34:8533–8541.<br />

54. Gould AR, Norton RS. Chemical modification of cationic groups in the polypeptide cardiac<br />

stimulant anthopleurin-A. Toxicon 1995; 33:187–199.<br />

55. Pennington MW, Kem WR, Dunn BM. Synthesis and biological activity of six monosubstituted<br />

analogs of a sea anemone polypeptide neurotoxin. Peptide Res 1990; 3:228–232.<br />

56. Gould AR, Mabbutt BC, Norton RS. <strong>Structure</strong>-function relationships in the polypeptide cardiac<br />

stimulant, anthopleurin-A. Effects of limited proteolysis <strong>by</strong> trypsin. Eur J Biochem 1990; 189:145–153.<br />

57. Gooley PR, Blunt JW, Beress L, Norton RS. Effects of pH and temperature on cardioactive<br />

polypeptides from sea anemones: a 1H-NMR study. Biopolymers 1988; 27:1143–1157.<br />

58. Sunahara S, Muramoto K, Tenma K, Kamiya H. Amino acid sequence of two sea anemone toxins<br />

from Anthopleura fuscoviridis. Toxicon 1987; 25:211–219.<br />

59. Gallagher MJ, Blumenthal KM. Cloning and expression of wild-type and mutant forms of the<br />

cardiotonic polypeptide anthopleurin-B. J Biol Chem 1992; 267:13958–13963.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_318.html (1 of 2) [4/5/2004 5:24:15 PM]


Document<br />

60. Pennington MW, Zadenberg I, Byrnes ME, Norton RS, Kem WR. Synthesis of the cardiac inotropic<br />

polypeptide anthopleurin-A. Intl J Pept Prot Res 1994; 43:463–470.<br />

61. Gallagher MJ, Blumenthal KM. Importance of the unique cationic residues arginine-12 and lysine-<br />

49 in the function of the cardiotonic polypeptide anthopleurin-B. J Biol Chem 1994; 269:254–259.<br />

62. Khera PK, Blumenthal KM. Role of the cationic residues arginine-14 and lysine-48 in the function<br />

of the cardiotonic polypeptide anthopleurin-B. J Biol Chem 269:921–925.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_318.html (2 of 2) [4/5/2004 5:24:15 PM]


Document<br />

Page 319<br />

63. Lampe RA, Lo MMS, Keith RA, Horn MB, McLane MW, Herman JL, Spreen RC. Effects of sitespecific<br />

acetylation on ω-conotoxin GVIA binding and function. Biochemistry 1993; 32:3255–3260.<br />

64. Basus VJ, Nadasdi L, Ramachandran J, Miljanich GP. Solution structure of ω-conotoxin MVIIA<br />

using 2D NMR spectroscopy. FEBS Lett 1995; 370:163–169.<br />

65. Manoleras N, Norton RS. Three-dimensional structure in solution of neurotoxin III from the sea<br />

anemone Anemonia sulcata. Biochemistry 1994; 33:11051–11061.<br />

66. Stanfield RN, Wilson IA. Protein-peptide interactions. Curr Opinion Str Biol 1995; 5:103–113.<br />

67. Stampe P, Kolmakova-Partensky L, Miller C. Intimations of K + channel structure from a complete<br />

functional map of the molecular surface of charybdotoxin. Biochemistry 1994; 33:443–450.<br />

68. Goldstein SAN, Pheasant DJ, Miller DJ. The charybdotoxin receptor of a Shaker K + channel:<br />

peptide and channel residues mediating molecular recognition. Neuron 1994; 12:1377–1388.<br />

69. Kim JI, Takahashi M, Ogura A, Kohno T, Kudo Y, Sato K. Hydroxyl group of Tyr 13 is essential for<br />

the activity of ω-conotoxin GVIA, a peptide toxin for N-type calcium channel. J Biol Chem 1994;<br />

269:23876–23878.<br />

70. Clackson T, Wells JA. A hot spot of binding energy in a hormone-receptor interface. Science<br />

1995;267:383–386.<br />

71. Jin L, Wells JA. Dissecting the energetics of an antibody-antigen interface <strong>by</strong> alanine shaving and<br />

molecular grafting. Protein Sci 1994; 3:2351–2357.<br />

72. Marshall GR. A hierarchical approach to peptidomimetic design. Tetrahedron 1993; 49:3547–3558.<br />

73. Chen S, Chrusciel RA, Nakanishi H, Raktabutr A, Johnson ME, Sato A, Wiener D, Hoxie J,<br />

Saragovi HU, Greene MI, Kahn M. <strong>Design</strong> and synthesis of a CD4 β-turn mimetic that inhibits human<br />

immunodeficiency virus envelope glycoprotein gp120 binding and infection of human lymphocytes.<br />

Proc Natl Acad Sci USA 1992; 89:5872–5876.<br />

74. Jackson S, De Grado W, Dwivedi A, Parthasarathy A, Higley A, Krywko J, Rock-well, A,<br />

Markwalder J, Wells G, Wexler R, Mousa S, Harlow R. Template-constrained cyclic peptides: design of<br />

high-affinity ligands for GPIIb/IIIa. J Am Chem Soc 1994; 116:3220–3230.<br />

75. Ohlstein EH, Nambi P, Douglas SA, Edwards RM, Gellai M, Lago A, Leber JD, Cousins RD, Gao<br />

A, Frazee JS, Peishoff CE, Bean JW, Eggleston DS, Elshourbagy NA, Kumar C, Lee JA, Yue T-L,<br />

Louden C, Brooks DP, Weinstock J, Feuerstein G, Poste G, Ruffolo RR, Gleason JG, Elliot JD. SB<br />

209670, a rationally designed potent nonpeptide endothelin receptor antagonist. Proc Natl Acad Sci<br />

USA 1994; 91:8052–8056.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_319.html (1 of 2) [4/5/2004 5:24:17 PM]


Document<br />

76. Gould AR, Mabbutt BC, Llewellyn LE, Goss NH, Norton RS. Linear and cyclic peptide analogues<br />

of the polypeptide cardiac stimulant anthopleurin-A. 1H-NMR and biological activity studies. Eur J<br />

Biochem 1992; 206:641–651.<br />

77. Stocker M, Miller C. Electrostatic distance geometry in a K + channel vestibule. Proc Natl Acad Sci<br />

USA 1994; 91:9509–9513.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_319.html (2 of 2) [4/5/2004 5:24:17 PM]


http://legacy.netlibrary.com/reader/message.asp?message=811&BookID=12640&FileName=Page_320.html<br />

The requested page could not be found.<br />

Return to previous page<br />

http://legacy.netlibrary.com/reader/message.asp?message=811&BookID=12640&FileName=Page_320.html [4/5/2004 5:24:22 PM]


Document<br />

13<br />

Rational <strong>Design</strong> of Renin Inhibitors<br />

V. Dhanaraj * and J.B. Cooper† **<br />

Birkbeck College, London, England<br />

I. Introduction<br />

Page 321<br />

There has been much interest in the development of therapies for hypertension and associated heart<br />

failure, which is a major cause of death in the western world. One of the key mediators in primary<br />

hypertension is the plasma octapeptide angiotensin II (AII), which plays a major role <strong>by</strong> causing<br />

vasoconstriction and stimulating aldosterone release, there<strong>by</strong> increasing blood volume <strong>by</strong> its action on<br />

the kidneys. Angiotensin II is produced <strong>by</strong> a proteolytic cascade—known as the renin-angiotensin<br />

system—in which the aspartic proteinase renin catalyses the rate-limiting cleavage of angiotensinogen<br />

produced <strong>by</strong> the liver to yield the decapeptide angiotensin I (AI). The subsequent removal of the<br />

carboxy-terminal dipeptide from AI <strong>by</strong> angiotensin-converting enzyme (ACE), yielding AII, is the target<br />

for a number of drugs that are effective for treating hypertension, hyperaldosteronism, and congestive<br />

heart failure [1].<br />

The development of potent low-molecular-weight orally active ACE inhibitors from natural and<br />

synthetic metalloproteinase inhibitors has been rapid, due in part to the relative lack of specificity of this<br />

enzyme. In contrast, renin cleaves only its natural substrate or very close analogs and although inhibition<br />

of an enzyme more specific than ACE may be desirable for reducing side effects in vivo, the selectivity<br />

of renin meant that during the early stages of drug development, potent inhibition required the use of<br />

large peptide-<strong>based</strong> compounds. These were often poorly absorbed and susceptible to gastric proteolysis<br />

and biliary excretion. Nevertheless, the commercial and clinical success of ACE inhibitors fueled<br />

interest in the search for therapeutic renin drugs. Most inhibitors have been developed <strong>by</strong> elaboration of<br />

the minimal substrate sequence (residues 6–13 of angiotensinogen), which exhibits weak competitive<br />

inhibition, and replacement of the scissile bond with various nonhydrolysable surrogates, some of which<br />

may be transition state analogues [2].<br />

Current affiliation: University of Cambridge, Cambridge, England.<br />

Current affiliation: University of Southhampton, Southhampton, England.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_321.html [4/5/2004 5:24:23 PM]


Document<br />

Figure 1<br />

The three-dimensional structures of human (left) and mouse renins (right) showing<br />

oligopeptide inhibitors bound in the active site cleft. The cleft lies between the N- and<br />

C-terminal domains of the enzyme and is approximately perpendicular to the plane<br />

of the page. It can accommodate 9–10 residues with the substrate/inhibitor bound<br />

in an extended conformation. The catalytic aspartic acid residues (not shown)<br />

are centrally placed at the base of the cleft.<br />

Page 322<br />

Renin is a member of the homologous group of enzymes known as aspartic proteinases that includes<br />

pepsin and a group of fungal enzymes such as endothiapepsin, penicillopepsin, and rhizopuspepsin.<br />

Their sequences all contain two aspartates (at positions 32 and 215 in porcine pepsin) that are essential<br />

for catalytic activity. The crystal structures of several aspartic proteinases have been solved <strong>by</strong> x-ray<br />

diffraction at high resolution, revealing a common bilobal structure with a large cleft between the N- and<br />

C-terminal domains that can accommodate up to nine residues of a substrate (Figure 1) [3]. The two<br />

essential carboxyls of Asp 32 and Asp 215 are within hydrogen-bonding distance and are approximately<br />

co-planar due to the constraints of a hydrogen-bonding network involving residues of the two highly<br />

conserved loops that contain the essential aspartates. The three-dimensional structures of the two<br />

domains are related <strong>by</strong> a topological two-fold axis passing between the catalytic residues where the<br />

pseudosymmetry happens to be strongest. Modeling studies <strong>based</strong> on the homology with other aspartic<br />

proteinases showed that human renin assumes a tertiary structure that is similar to the other enzymes and<br />

that the homology is greatest for the binding cleft region [4]. Subsequent x-ray analysis of the structure<br />

of renin (described later) revealed the specific interactions made with inhibitors and implicated certain<br />

loop regions covering the active site as being important for tight binding of peptides.<br />

One enigmatic feature of renin is its extreme substrate specificity, its only known natural substrate being<br />

a single Leu-Val peptide bond of angiotensinogen. The minimal synthetic analog is the 6–13 octapeptide<br />

that encompasses the scissile peptide bond of angiotensinogen between residues 10 and 11 [5]. It has<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_322.html [4/5/2004 5:24:30 PM]


Document<br />

Page 323<br />

been suggested that specificity of proteinases in general is due to rigidity in the binding pockets [6];<br />

broad specificity results from the ability of the pockets to change shape in response to different ligand<br />

side chains. However, renin is known to be inhibited <strong>by</strong> a wide variety of peptide analogs of different<br />

length and sequence indicating either that the active site may be somewhat flexible or that the strength<br />

of binding of a substrate does not determine the rate of subsequent turnover. This is emphasised <strong>by</strong> the<br />

existence of substrates that act as competitive inhibitors, e.g., RIP of Haber and Burton [7], which has a<br />

K i that is lower than the K m. Therefore hydrolysis of bound substrate appears to be more specific than<br />

the binding step. This may be because only certain substrate sequences allow correct positioning of the<br />

scissile bond for hydrolysis. Evidence for this effect was provided <strong>by</strong> comparison of 21 inhibitor<br />

structures of endothiapepsin [8], where it was shown that for inhibitors with different sequences but with<br />

the same transition state analog, the scissile bond analog can be disposed somewhat differently with<br />

respect to the catalytic carboxyls in each case.<br />

The availability of crystal structures of a number of renin inhibitors complexed with fungal aspartic<br />

proteinases [8, 9] allowed new compounds to be designed and modeled <strong>by</strong> such techniques as computer<br />

graphics, energy minimization, and molecular dynamics [10]. X-ray crystallographic analysis of aspartic<br />

proteinase inhibitor complexes has made a significant contribution to rationalizing the activity data for<br />

many of these compounds as well as understanding the catalytic mechanism of this class of proteinase.<br />

II. Strategies for <strong>Design</strong> of Renin Inhibitors<br />

Some of the parameters that have been varied in the search for therapeutically active renin inhibitors are<br />

outlined below.<br />

A. Elaboration of the Transition State Analog<br />

There is no evidence that aspartic proteinase catalysis involves a covalently bound intermediate [11] and<br />

major advances in the design of nonhydrolysable analogs have stemmed from attempts to mimic an<br />

intermediate of the following form.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_323.html [4/5/2004 5:24:33 PM]


Document<br />

Page 324<br />

This intermediate is derived <strong>by</strong> nucleophilic attack of a water molecule on the scissile-bond carbonyl.<br />

Szelke pioneered the use of reduced-bond analog (-CH 2-NH-), which were incorporated into the 6–13<br />

peptide of angiotensinogen [12]. Cocrystallisation with endothiapepsin revealed that the reduced-bond<br />

analog associates tightly with aspartate carboxyls (32 and 215) probably via a salt link. The naturally<br />

occurring transition-state analogue statine (-CHOH-CH 2-CO-NH-) [9] is a closer analogue of the<br />

putative intermediate and has been incorporated into many inhibitors [13]. The scissile bond has also<br />

been replaced <strong>by</strong> the ketone analogue (-CO-CH 2-) [12] or <strong>by</strong> a C-terminal aldehyde group in a series of<br />

tetrapeptides [14,15]. Although these appear to mimic the substrate more closely than the intermediate,<br />

the carbonyl probably binds to the enzyme in the hydrated gem-diol form (-C(OH) 2-CH 2) [16]. Use of<br />

the hydroxyethylene analog (-CHOH-CH 2-) has led to exceptionally potent inhibitors [17] as has<br />

substitution of fluorines into ketone analogs [16,18] giving, for example, -CO-CF 2- which undergoes<br />

hydration of the carbonyl to form -C(OH) 2-CF 2- and exhibits tight binding to renin. The hydrated gemdiol<br />

is thought to closely mimic the putative transition state -C(OH) 2-NH-.<br />

All inhibitors solved in complex with aspartic proteinases <strong>by</strong> x-ray diffraction are observed to adopt<br />

similar main-chain conformations and form a conserved set of hydrogen bonds involving the inhibitor's<br />

main-chain groups interacting with enzyme moieties. The inhibitors bind in extended conformations and<br />

residues in the P 3-P 1 region form antiparallel β-sheet-like interactions with residues 217–219 on the<br />

enzyme. A β-hairpin turn formed <strong>by</strong> residues 74–78 lies between the inhibitor and bulk solvent and<br />

forms a number of generally conserved hydrogen bonds with the bound peptide. These interactions are<br />

indicated in Figure 2. The binding pockets for the inhibitor's side chains are shallow and contiguous with<br />

a greater hydrophobic character towards the central region of the active site cleft.<br />

The elaboration of several classes of transition state analog are now considered in greater detail.<br />

Statine Analogs<br />

The natural transition analog statine possesses one less main chain atom than a dipeptide and has been<br />

shown <strong>by</strong> x-ray analysis of inhibitor cocrystals to occupy the S 1 and S 1' sites of the enzyme [19]. The<br />

hydroxyl group of statine binds symmetrically between the catalytic carboxyl groups displacing a<br />

solvent molecule bound to the native enzyme. The carboxyl diad, therefore, provides a stereospecific<br />

binding site for statine and hydroxyethylene analogues with preference for the S-enantiomer.<br />

Replacement of the hydroxyl <strong>by</strong> an ammonium group might be expected to improve the potency of an<br />

inhibitor <strong>by</strong> introduction of a salt link with the enzyme. The corresponding deoxy-aminostatine (ASTA)<br />

analogs have been synthesised [20, 21] and were found to be nearly as potent as<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_324.html [4/5/2004 5:24:37 PM]


Document<br />

Figure 2<br />

A schematic diagram of the putative hydrogen bonds formed between oligopeptide inhibitors and the fungal aspartic<br />

proteinase endothiapepsin. The latter enzyme provided a useful model system for structural studies of interactions formed <strong>by</strong> renin<br />

inhibitors with the active site cleft of aspartic proteinases prior to the determination of the human renin structure. The inhibitor is shown<br />

horizontally with enzyme groups above and below. Intervening hydrogen bonds are indicated <strong>by</strong> dashed lines. Note the extensive<br />

hydrogen-bond interactions made between the transition state analogs and the catalytic apparatus of the enzyme.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_325.html [4/5/2004 5:24:51 PM]<br />

Page 325


Document<br />

the equivalent statine analogs with the benefit of improved solubility. Preference for the S- versus Renantiomer<br />

at the C3 amino position is observed in accordance with the statine inhibition data [22].<br />

Page 326<br />

Oxahomostatine (-CHOH-CH 2-O-CO-) and azahomostatine (-CHOH-CH 2-NR-CO-) analogs eliminate<br />

the main-chain frameshift that occurs with statine and, in addition, the azahomostatine conveniently<br />

reduces the stereochemical complexity of the dipeptide surrogate <strong>by</strong> introducing a 7-membered urea-like<br />

planar group into the S 1' binding region. The crystal structure of such a compound complexed with<br />

endothiapepsin has been solved at 1.8 Å resolution [23] and reveals that the large planar group is<br />

accommodated <strong>by</strong> the active site and that hydrogen bonds to the P 1' CO and P 2' NH groups, observed in<br />

other complexes, are retained.<br />

One shortened analog of statine -CHOH-CO-O-R referred to as norstatine (where R is a C-terminal alkyl<br />

group) has been found to be more potent than some equivalent statine analog [24]. The x-ray structure of<br />

such an inhibitor complexed with endothiapepsin reveals that the carbonyl oxygen of this analog is held<br />

<strong>by</strong> a hydrogen bond to the active site flap region of the enzyme involving the Gly76 >NH group (pepsin<br />

numbering) in much the same way as the P 1' >C=O group of other isosteres (Figure 2).<br />

Aminoalcohols<br />

In principle, a good analog of the putative intermediate would be the aminal –CHOH–NH– group but<br />

this would be in equilibrium with the aldehyde and amino fragments. Interposing a methylene group<br />

between the hydroxymethyl and amino groups stabilizes the analog and may still allow tight binding to<br />

the enzyme. Such aminoalcohols (-CHOH-CH 2-NH-) have been synthesized [25] and were shown to be<br />

potent inhibitors.<br />

Cocrystallisation of two such compounds extending from P 1 to P 3' with endothiapepsin allowed their<br />

bound structures to be solved at high resolution [26]. The bound structures revealed that despite the<br />

insertion of a methylene group in the analog a frameshift in the binding mode does not occur since the<br />

residue following the aminoalcohol occupies the S 1' pocket. In contrast, the single amino acid, statine,<br />

replaces two residues of the substrate. The hydroxyl of the aminoalcohol (S-enantiomer) is bound<br />

symmetrically to both essential carboxyls as is the case for the hydroxyl of the statine and<br />

hydroxyethylene analogs.<br />

Glycols<br />

Incorporation of glycol or vicinal diol analogs of the peptide bond (-CH(OH)-CH(OH)-) has led to<br />

potent inhibitors and the x-ray structure for one such compound complexed with endothiapepsin is<br />

available [27]. The first hydroxyl in<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_326.html [4/5/2004 5:24:56 PM]


Document<br />

Page 327<br />

this analog interacts with the catalytic aspartate carboxyls in the same manner as statine or<br />

hydroxyethylene moieties; whereas, the second hydroxyl forms a hydrogen bond with the NH group of<br />

Gly 76, there<strong>by</strong> mimicing the carbonyl oxygen at P 1' of other analogs (see Figure 2).<br />

Phosphorus-Containing Analogs<br />

Aspartic proteinase inhibitors in which the scissile bond is replaced <strong>by</strong> a phosphinic acid group (shown<br />

below) have been reported [28].<br />

These may mimic the tetrahedral intermediate more closely than statine or hydroxyethylene analogs.<br />

One of the oxygens binds to the carboxyl diad and the other resides adjacent to Tyr75 (pepsin<br />

numbering) forming a hydrogen bond with the outer oxygen of Asp32 [27]. This isostere is very<br />

effective against pepsin. However, it ionizes at physiological pH and the resulting anion is ineffective as<br />

an inhibitor of renin [29].<br />

Fluoroketone Analogs and Implications for Catalysis<br />

Fluoroketone analogs (-CO-CF 2-) have been reported [16, 30] and found to be substantially more potent<br />

than the unhalogenated statone molecules, presumably due to the ease of hydration and greater<br />

complementarity of the resulting hydrated gem-diol with the catalytic site. The structure of a<br />

difluorostatone inhibitor complexed with endothiapepsin [31] revealed interactions that indicate how the<br />

catalytic intermediate is stabilized <strong>by</strong> the enzyme (Figure 3). One hydroxyl of the hydrated fluoroketone<br />

associates tightly with the aspartate diad in the same position as the statine hydroxyl or the native<br />

solvent molecule and the other hydroxyl is positioned such that it hydrogen bonds to the outer carboxyl<br />

oxygen of Asp32. It has been suggested that the tetrahedral intermediate is uncharged, because if the<br />

carboxyl of Asp32 carries a negative charge instead, the latter can be stabilized <strong>by</strong> a full complement of<br />

hydrogen bonds donated <strong>by</strong> the gem-diol intermediate and surrounding protein atoms [31]. The current<br />

mechanistic proposals are <strong>based</strong> on the key suggestion <strong>by</strong> Suguna et al. [32] that, although transition<br />

state analogs appear to displace the active-site water molecule located between the two catalytic<br />

aspartate carboxyls, the more weakly bound substrate may not. Instead as the substrate binds, the water<br />

may be partly displaced to a position appropriate for nucleophilic attack on the scissile bond carbonyl.<br />

Details of the proposed mechanism are given in Figure 3.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_327.html [4/5/2004 5:25:04 PM]


Document<br />

Figure 3<br />

The catalytic mechanism proposed <strong>by</strong> Veerapandian et al. [31] <strong>based</strong> on the x-ray<br />

structure of a difluoroketone (geminal-diol) inhibitor bound to endothiapepsin. A water<br />

molecule tightly bound to the aspartates in the native enzyme is proposed to<br />

nucleophilically attack the scissile-bond carbonyl. The resulting geminal-diol<br />

intermediate is stabilised <strong>by</strong> hydrogen bonds with the negatively charged carboxyl of aspartate 32.<br />

Fission of the scissile C-N bond is accompanied <strong>by</strong> transfer of a proton from Asp215 to<br />

the leaving amino group.<br />

B. Complementarity of the Inhibitor<br />

Page 328<br />

Optimizing the fit of a ligand to its binding site improves the potency <strong>by</strong> burying lipophilic residues and<br />

<strong>by</strong> maximizing the number of van der Waals contacts, hydrogen bonds, and charge—charge interactions.<br />

The principles that apply to ligand binding are similar to those involved in protein folding. Inhibitor<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_328.html (1 of 2) [4/5/2004 5:25:14 PM]


Document<br />

Page 329<br />

binding involves displacement of hydrogen-bonded water molecules from both the ligand and the<br />

binding cleft. This process is entropy favored since waters in the solvent lattice are more disordered than<br />

those bound to protein. The change in enthalpy on forming NH…CO bonds in the complex from<br />

>CO…HOH and >NH…OH 2 is favorable but relatively small [33]. The hydrophobic effect is therefore<br />

thought to play a dominant role in the energetics of binding with hydrogen bonds providing precise<br />

alignment of the ligand with respect to the catalytic apparatus. The main chain >CO and >NH groups<br />

from P 3 to P 3' of aspartic proteinase inhibitors are nearly always satisfied <strong>by</strong> hydrogen-bond interactions<br />

on formation of the complex. Therefore, given that polypeptides can form the same hydrogen bonds to<br />

the binding cleft regardless of amino acid sequence, differences in affinity for ligands of equal length<br />

must be due to other interactions at the specificity pockets, presumably those between the ligand's side<br />

chains and the enzyme.<br />

One example of optimizing these interactions for renin is the use of the cyclohexylmethyl side chain at<br />

P 1, which has been shown to improve the potency <strong>by</strong> two orders of magnitude relative to the equivalent<br />

leucine-containing inhibitor [13]. <strong>Structure</strong>/Activity Relationship (SAR) studies have shown that in<br />

many inhibitor types, the cyclohexylmethyl group is optimal for the S 1 pocket of human renin; whereas,<br />

other analogs such as cyclohexyl, cyclohexylethyl, and the very bulky dicyclohexyl and adamantyl rings<br />

generally have significantly reduced potency [10]. The use of a cyclohexylmethyl appears to introduce<br />

selectivity for renin versus other human aspartic proteinases. This has been partly rationalized for<br />

endothiapepsin where it was shown <strong>by</strong> x-ray analysis that the cyclohexylmethyl group at P 1 can force<br />

the Phe at P 3 to adopt a less energetically favorable X 2 angle. Hence, differences at the S 3 pocket in renin<br />

may allow the P 3 Phe to adopt a more favorable X 2 angle in the presence of a cyclohexyl at P 1. In<br />

contrast the S 2 site is able to accommodate a wide variety of side chains depending on inhibitor type,<br />

e.g., Phe and His are equipotent in some analogs [34]. The x-ray structures of a number of bound renin<br />

inhibitors complexed with endothiapepsin have shown that His at P 2 can adopt different X 1 angles<br />

separated <strong>by</strong> about 120 degrees [8]. In one conformation the imidazole is lying partly in the S 1' pocket,<br />

which has a definite hydrophobic character. In the other conformation, the His side chain is in a more<br />

polar environment. The ability of aspartic proteinases to accept a variety of both polar and hydrophobic<br />

groups at the P 2 position may be due to this bifurcation. Many inhibitors possess naphthylalanine side<br />

chains at P 3 and P 4 [14,24,35]. Compounds of this type are potent renin inhibitors with binding constants<br />

in the nanomolar range. Cocrystallisation of such an inhibitor with endothiapepsin revealed that one<br />

naphthalic ring is accommodated in the S 3 pocket <strong>by</strong> significant conformational changes of local enzyme<br />

side chains (Asp77 and Asp114). The other naphthalene lies in the S 4 binding region [36].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_329.html [4/5/2004 5:25:18 PM]


Document<br />

C. Rigidification<br />

Page 330<br />

The number of conformations that a peptide can adopt in solution is reduced <strong>by</strong> cyclization. This can be<br />

optimized, at least in theory, to lock the peptide in the conformation that has the highest affinity for the<br />

receptor, resulting in a gain of affinity, primarily for entropic reasons. The structures of the fungal<br />

aspartic proteinases reveal that the binding cleft is a wide channel with no obvious division between the<br />

pockets, e.g., S 1 and S 3 are contiguous. The bound structures of numerous inhibitors have shown that<br />

alternate side chains are in van der Waals contact due to the extended conformation that these ligands<br />

adopt. In addition, at certain positions, e.g., P 2, the side chains are allowed very different conformations<br />

due to the permissiveness of the pocket. Hence, the cross-linking of certain side chains may, at least, not<br />

be detrimental to inhibitory potency and may also reduce the susceptibility to degradation in the gut or<br />

plasma. It might therefore be expected that oligopeptide renin inhibitors would be suitable' candidates<br />

for cross-linking experiments. A similar philosophy of rigidification was pursued in the development of<br />

the ACE inhibitor cilazapril [37].<br />

A number of statine-containing inhibitors possessing disulphide links between P 2 and P 5, and P 2 and P 4'<br />

have been synthesized [13] although the best potencies were slightly less than for the linear peptides. An<br />

alkyl cycle of varying length was introduced between the hydroxyl of a serine residue at P 1 and the main<br />

chain nitrogen of P 2 in a series of reduced-bond inhibitors [53]. Potencies similar to the uncrosslinked<br />

molecule were achieved but none were greater. This was attributed to the cis isomerisation of the P 3—P 2<br />

peptide bond giving a conformation that cannot fit the active site of the enzyme. Difficulties in<br />

achieving more potent cyclic inhibitors may be due to the tight binding environment provided <strong>by</strong> some<br />

pockets (especially S 1 and S 3), and the possibility that other unproductive conformations of the inhibitor<br />

become favorable. More recently similar findings have been reported for cyclic analogs of pepstatincontaining<br />

alkyl crosslinks of variable length between the P 1 and P 3 side chains [38].<br />

D. In Vivo Stability<br />

Peptides, when administered orally, are susceptible to degradation in the stomach <strong>by</strong> gastric enzymes<br />

and the proteinases of the pancreas and brush border of the small intestine. Their lifetimes in the plasma<br />

are often short due to rapid proteolysis and other metabolic processes. Early efforts were made to<br />

improve the resistance of renin inhibitors to hydrolysis in vivo <strong>by</strong> the use of blocking groups at the N-<br />

and C-terminii [39] and replacement of susceptible peptide bonds other than the renin cleavage site.<br />

Studies of SAR have shown that various N- and C-terminal groups, some <strong>based</strong> on the morpholine<br />

nucleus and derivatives of it, have a favorable effect on the duration of inhibition in the<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_330.html [4/5/2004 5:25:25 PM]


Document<br />

plasma. This may arise from reduced nonspecific plasma binding due to the relatively polar nature of<br />

these blocking groups [24].<br />

Page 331<br />

Resistance of inhibitors to gut proteolysis has been improved <strong>by</strong> various methods such as replacement of<br />

phenylalanine at P 3 with O-methyl tyrosine (or naphthylalanine), which was shown to abolish<br />

chymotrypsin cleavage and yet retain high inhibitory potency for renin [40].<br />

III. Structural Studies of Rennin Complexed With Inhibitors<br />

The three-dimensional structures of renin-inhibitor complexes had long been sought as an aid to the<br />

discovery of clinically effective antihypertensives [41]. X-ray analyses of recombinant human renin [42]<br />

and mouse submandibulary renin [43] have given an accurate picture of active-site interactions and<br />

largely confirm the predictions of models <strong>based</strong> on homologous aspartic proteinases [4]. A large number<br />

of questions concerning the specificities of renins have been answered <strong>by</strong> these x-ray analyses. The<br />

renin-inhibitor structures also make an important contribution towards the rational design of effective<br />

antihypertensive agents.<br />

A. X-Ray Analysis of Mouse and Human Renin Complexes<br />

For both of these renins multiple copies of the molecules have been independently defined in the x-ray<br />

analysis and shown to have very similar structures. These x-ray structures were refined to final<br />

agreement factors and correlation coefficients of 0.19 and 0.91 for human renin at 2.8 Å resolution and<br />

0.18 and 0.95 for mouse renin at 1.9 Å resolution. As expected from the high degree of sequence<br />

identity of human and mouse renins (approximately 70%), they have very similar three-dimensional<br />

structures as shown in Figure 1.<br />

The active-site cleft has a less open arrangement in renins than in the other aspartic proteinases. Many<br />

loops as well as the helix h c (residues 224–236) belonging to the C-domain (residues 190–302) are<br />

significantly closer to the active site in the renin structures compared to those of endothiapepsininhibitor<br />

complexes. This is partly due to a difference in relative position of the rigid body comprising<br />

the C-domain. For instance, there is a domain rotation of ~4 ° and translation of ~0.1 Å in the human<br />

renin complex with respect to the endothiapepsin-difluorostatone complex.<br />

The entrance to the active-site cleft is made even narrower in renins as a consequence of differences in<br />

the positions and composition of several well-defined loops and secondary structure elements. Unique to<br />

the renins is a cis proline, Pro111, which caps a helix (h N2) and contributes to the subsites S 3 and<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_331.html [4/5/2004 5:25:30 PM]


Document<br />

Page 332<br />

S 5. This helix is nearer to the active site in renins than in other aspartic proteinases. On an equivalent<br />

loop in the C-lobe (related <strong>by</strong> the intramolecular pseudo 2-fold axis), there is a sequence of three<br />

prolines—the Pro292–Pro293– Pro 294 segment. This structure is also unique to the renins among the<br />

aspartic proteinases with Pro294 and Pro297 in a cis configuration. Such a proline-rich structure<br />

provides an effective means of constructing well-defined pockets from loops that would otherwise be<br />

more flexible.<br />

This rather rigid poly-proline loop, together with the loop comprised of residues 241–250, lies on either<br />

side of the active site “flap” formed <strong>by</strong> residues 72–81. Hence, in the renins, the cleft is covered <strong>by</strong> the<br />

flaps from both lobes rather than from the N-lobe alone as in other pepsin-like aspartic proteinases. This<br />

gives renin a superficial similarity to the dimeric, retroviral proteinases where each subunit provides an<br />

equivalent flap that closes down on top of the inhibitor [44,45].<br />

B. The Role of Hydrogen Bonds in Inhibitor Recognition<br />

Whereas the mouse renin inhibitor extends from P 6 to P 4', the human renin inhibitor extends only from<br />

P 4 to P 1'. The cyclohexyl norstatine residue at P 1 in the human renin inhibitor mimics a dipeptide analog<br />

with its isopropyloxy group occupying the subsite for the side chain of P 1'. The mouse renin inhibitor<br />

(CH-66) possesses a Leu-Leu hydroxyethylene transition state analog [12]. Both inhibitors are bound in<br />

the extended conformation that is found in other aspartic proteinase-inhibitor complexes. Both inhibitors<br />

make extensive hydrogen bonds with the enzymes as shown in Figure 4. In general the two renin-<br />

inhibitor complexes described here demonstrate that a similar pattern of hydrogen bonding is probably<br />

used in the substrate recognition of all aspartic proteinases although their specificities differ<br />

substantially.<br />

There is also great similarity between aspartic proteinases in terms of interactions with the transitionstate<br />

analog inhibitors at the catalytic center. The catalytic aspartyl side chains and the inhibitor<br />

hydroxyl group are essentially superimposable in both renin complexes. The isostere C-OH bonds lie at<br />

identical positions when the structures of inhibitor complexes of several aspartic proteinases are<br />

superposed, in spite of the differences in the sequence and secondary structure. Most of the complex<br />

array of hydrogen bonds found in endothiapepsin complexes are formed in renin with the exception of<br />

that to the threonine or serine at 218, which is replaced <strong>by</strong> alanine in human renin. The similarity can be<br />

extended to all other pepsin-like aspartic proteinases and even to the retroviral proteinases [44,45]. This<br />

implies that the recognition of the transition state is conserved in evolution, and the mechanisms of this<br />

divergent group of proteinases must be very similar.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_332.html [4/5/2004 5:25:32 PM]


Document<br />

C. Specificity<br />

Figure 4<br />

The inhibitors complexed with human (top) and mouse (bottom) renin<br />

showing the putative hydrogen-bond interactions made with the enzyme moieties.<br />

Page 333<br />

If the main-chain hydrogen bonding of substrates is conserved among aspartic proteinases, how are the<br />

differences in specificities achieved? Table 1 defines the enzyme residues that line the specificity<br />

pockets for both mouse and human renin. In modeling exercises (e.g., Reference 4) it was assumed that<br />

specificities derive from differences in the sizes of the residues in the specificity pockets (S n) and their<br />

ability to complement the corresponding side chains at positions P n in the substrate/inhibitor. A detailed<br />

analysis now shows that this simple assumption only partly accounts for the steric basis of specificity.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_333.html (1 of 2) [4/5/2004 5:25:42 PM]


Document<br />

Page 334<br />

For example, in the specificity subsite S 3 the phenyl rings of Phe P 3 occupy almost identical positions in<br />

both renin inhibitor complexes. Modeling studies have predicted the specificity subsite S 3 to be larger in<br />

renins than in other aspartic proteinases [4] due to substitution of smaller residues, Pro 111, Leu114, and<br />

Ala115, in place of larger ones in mammalian and fungal proteinases. However, a compensatory<br />

movement of a helix (h N2) makes the pocket quite compact and complementary to the aromatic ring as<br />

shown in Figure 5. Thus, the positions of an element of secondary structure differ between renin and<br />

other aspartic proteinases with a consequent important difference in the specificity pocket.<br />

The differing positions of secondary structural elements may also account for the specificities at P 2'.<br />

Mouse submaxillary and other nonprimate renins do not appreciably cleave human angiotensinogen or<br />

its analogs [46], which have an isoleucine at P 2', although they do cleave substrates with a valine at this<br />

position. In contrast, human renin not only cleaves the human and nonprimate substrates but also the rat<br />

angiotensinogen with tyrosine at P 2', albeit rather slowly [47]. This can be explained in terms of the threedimensional<br />

structures. In the mouse renin complex, the P 2' tyrosyl ring is packed parallel to an<br />

adjoining helix (h 3) in a narrow pocket and there is only limited space available beyond the Cβ<br />

methylene group. This appears to be able to accommodate a valine, but not the larger isoleucine at P 2',<br />

which will suffer greater steric interference from several residues that are conserved in identity and<br />

position in the two renins. On the other hand, in human renin differences in the orientation and position<br />

of<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_334.html [4/5/2004 5:26:06 PM]


Document<br />

Figure 5<br />

The S 3 specificity pocket of human renin occupied <strong>by</strong> phenylalanine in the<br />

cyclohexylnorstatine inhibitor.<br />

Page 335<br />

helix h 3 bring it closer <strong>by</strong> (<strong>by</strong> ˜0.5Å) to the substrate-binding site than in mouse renin. It is orientated in<br />

such a fashion in human renin that, although it can accommodate the isobutyl side chain of isoleucine at<br />

P 2', aromatic rings on substituents such as phenylalanine and tyrosine will have severe short contacts<br />

with the side chain of Ile130 (valine in mouse renin). Thus the reorientation of a helix, coupled with<br />

subtle differences in the shapes of the side chains, makes significant changes in the substrate specificity<br />

at this subsite. It is interesting to note that in pepsin this helix is in a similar position with respect to the<br />

active site as in human renin. This provides a structural rationale for the negative influence of peptides<br />

containing phenylalanine [48], tyrosine, or histidine [49] at this subsite (S 2') on the rate of proteolytic<br />

pepsin cleavage, while isoleucine and valine enhance catalysis.<br />

Differences in the specificity subsites at S 1' in the human and mouse renins have a more complicated<br />

explanation. At first sight the situation appears to be explained <strong>by</strong> complementarity of the subsites to the<br />

valine and leucine at<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_335.html (1 of 2) [4/5/2004 5:26:14 PM]


Document<br />

Page 336<br />

P 1' in human and mouse angiotensinogens. Accordingly residue 213 is leucine in human renin and valine<br />

in mouse renin. The S 1' pockets of chymosin, pepsin, and endothiapepsin have an aromatic side chain at<br />

residue 189 while the renins have amino acids with smaller side chains (valine in human and serine in<br />

mouse renins). This would be expected to make the pocket larger in renins. However the structure of the<br />

mouse renin complex shows that the substrate moves closer to the enzyme in renins as a result of the<br />

smaller residue at 189 and the pocket is made even more compact due to a compensatory change in the<br />

position and composition of the polyproline loop (residues 290–297). Thus, the specificity difference at<br />

this site arises not only from a compensatory movement of a secondary structure, in this case a loop<br />

region, but also from the substitution of an enzyme residue that allows the substrate to come closer to<br />

the body of the enzyme.<br />

Elaboration of loops on the periphery of the binding cleft in renins also influences the specificity. This is<br />

most marked at P 3' and P 4', for which it has been particularly difficult to obtain complexes with welldefined<br />

conformations for other aspartic proteinases. In endothiapepsin, which has been the subject of<br />

the greatest number of studies, different conformations are adopted at P 3' and the residue at P 4' is<br />

generally disordered. In contrast these residues are clearly defined in mouse renin. This is mainly a<br />

consequence of the polyproline loop, illustrated in Figure 6, which occurs uniquely in renins. The x-ray<br />

analysis of the mouse renin complex shows that the S 3' and S 4' subsites are formed <strong>by</strong> the polyproline<br />

loop together with residues of the flap, and a similar situation is likely to occur in human renin. The welldefined<br />

interactions of P 3' described in the mouse renin complex explains the significant affinity when<br />

inhibitors have phenylalanine or tyrosine at P 3' as well as the importance of a P 3' residue for catalytic<br />

cleavage of a substrate <strong>by</strong> renin [50].<br />

Hydrogen bonds between the side chains of the inhibitor and the enzyme do not play a major role in<br />

most specificity pockets. However, S 2 is an exception. This subsite is large and contiguous with S 1', so<br />

that in human renin the S-methyl cysteine (SMC) side chain of P 2 is oriented towards the S 1' pocket,<br />

which is only partly filled <strong>by</strong> the isopropyloxy group of the putative P 1' residue. The carbonyl oxygen of<br />

P 2 accepts a hydrogen from the O γ of Ser76, which is unique to human renin; residue 76 is a highly<br />

conserved glycine in all the other aspartic proteinases, including mouse renin. In mouse renin the P 2<br />

histidyl group has a different orientation and forms a hydrogen bond with the O γ of Ser222. If such a<br />

conformation were adopted <strong>by</strong> the human angiotensinogen in complex with human renin, the two<br />

imidazole nitrogens would be hydrogen bonded to the O γ of both Ser76 and Ser222. The observed<br />

reduction in the rate of cleavage of a human angiotensinogen analog containing a 3-methyl histidine<br />

substituent at P 2 [51] could be explained on the basis of the hydrogen bonding scheme proposed above.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_336.html [4/5/2004 5:26:17 PM]


Document<br />

IV. Rational <strong>Drug</strong> <strong>Design</strong><br />

Figure 6<br />

The P 3' tyrosine residue of the mouse renin inhibitor complex showing<br />

the unique polyproline loop on the right. Specificity of this and<br />

neighboring subsites in renins must derive partly from this rigid loop<br />

region.<br />

Page 337<br />

The pioneering work of Burton, Szelke, and others in developing peptide-<strong>based</strong> renin inhibitors has been<br />

followed <strong>by</strong> a worldwide commercial effort to elaborate such compounds into therapeutically active<br />

antihypertensives. The twin problems of insufficient oral bioavailability and rapid clearance has<br />

seemingly presented major obstacles to success. In addition, the possible advantages of renin inhibitors<br />

compared with ACE inhibitors remain questionable. Never-theless information from human-renin<br />

crystallographic studies—such as the more recent high resolution analyses [52] and algorithms for<br />

analysing voids in the complexes as potential sites for elaborating the drug molecule (e.g. Figure<br />

7)—may yet provide leads for compounds with suitable therapeutic characteristics.<br />

The detailed analyses of renin-inhibitor complexes reported here confirm the general structural features<br />

thought to contribute to renin's specificity but<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_337.html (1 of 2) [4/5/2004 5:26:24 PM]


Document<br />

Figure 7<br />

Schematic illustration of the voids between enzyme and<br />

inhibitor in the crystal structure of human renin complexed with a norstatine<br />

inhibitor. The figure was produced using the GapE software<br />

(Dr. Roman Laskowski). The inhibitor (dark bonds) is<br />

enclosed <strong>by</strong> a net surface and the gaps (where the<br />

enzyme and inhibitor are not in<br />

contact) are represented <strong>by</strong> solid surfaces [42].<br />

Page 338<br />

demonstrate the need for careful, high-resolution x-ray analyses for more confidence in drug design. In<br />

particular, they show that even minor alternations in the positions of secondary structural elements can<br />

lead to major changes in the disposition of the subsites and thus the recognition of substrates. Since such<br />

molecular recognition defines the species specificity and determines the catalytic efficiency of the<br />

enzymes, a through understanding is indispensable for the synthesis of suitable inhibitors. The<br />

specificity pockets—the molecular recognition sites—are modified <strong>by</strong> elaboration, particularly of<br />

surface loops, which can be disordered in the uncomplexed enzymes and difficult to model with<br />

precision from homologous structures. These data establish a new foundation for the rational design of<br />

renin inhibitors and have provided a rational base for development of clinically successful HIV<br />

proteinase inhibitors.<br />

References<br />

1. Ondetti MA, Cushman DW. Enzymes of the renin-angiotensin system and their inhibitors. Annu Rev<br />

Biochem 1982; 51:283–308.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_338.html (1 of 2) [4/5/2004 5:27:09 PM]


Document<br />

2. Blundell TL, Cooper J, Foundling SI, Jones DM, Atrash B, Szelke M. On the rational design of renin<br />

inhibitors: X-ray studies of aspartic proteinases complexed with transition state analogues. Biochemistry<br />

1987; 26:5585–5590.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_338.html (2 of 2) [4/5/2004 5:27:09 PM]


Document<br />

Page 339<br />

3. Pearl LH, Blundell TL. The active sites of aspartic proteinases. FEBS Lett 1984; 174:96–101.<br />

4. Sibanda BL, Blundell TL, Hobart PM, Fogliano M, Bindra JS, Dominy BW, Chirgwin JM. Computer<br />

graphics modeling of human renin. FEBS Lett 1984; 174:102–111.<br />

5. Skeggs LT, Dover FE, Levine M, Lentz KE, Kahn JR. In: Johnson JA, Anderson RR, ed. The Renin-<br />

Angiotensin System. New York: Plenum.<br />

6. Bone R, Silen JL, Agard DA. Structural plasticity broadens the specificity of an engineered protease.<br />

Nature 1989; 339:191–195.<br />

7. Haber E, Burton J. Inhibitors of renin and their utility in physiologic studies. Fedn Proc 1979;<br />

38:2768–2773.<br />

8. Cooper JB, Bailey D. A structural comparison of 21 inhibitor complexes of the aspartic proteinase<br />

from Endothia parasitica. Protein Science 1994, 3:2129–2143.<br />

9. Foundling SI, Cooper, J, Watson FE, Cleas<strong>by</strong> A, Pearl LH, Sibanda BL, Hemmings A, Wood SP,<br />

Blundell TL, Valler MJ, Norey CG, Kay J, Boger J, Dunn BM, Leckie BJ, Jones DM, Atrash B, Hallett<br />

A, Szelke M. High resolution X-ray analyses of renin inhibitor-aspartic proteinase complexes. Nature<br />

(London) 1987; 327:349–352.<br />

10. Luly JR, Bolis G, Bamaung N, Soderquist J, Dellaria JF, Stein H, Cohen J, Thomas JP, Greer J,<br />

Plattner JJ. New inhibitors of human renin that contain novel replacements. Examination of the P 1 site. J<br />

Med Chem 1988; 31:532–539.<br />

11. Hofmann T, Fink AL. Cryoenzymology of penicillopepsin. Biochemistry 1984; 23:5249–5256.<br />

12. Szelke M, Leckie B, Hallett A, Jones DM, Sueiras-Diaz J, Atrash B, Lever AF. Potent new<br />

inhibitors of human renin. Nature 1982; 299:555–557.<br />

13. Boger J. Renin inhibitors. <strong>Design</strong> of angiotensin transition state analogues containing statine. In:<br />

Kostka V. ed. Aspartic Proteinases and Their Inhibitors. Berlin: Walter de Gruyter, 1985:401–420.<br />

14. Kokubu T, Hiwada K, Murakami E, Imamura Y, Matsueda R, Yabe Y, Koike H, Iijima Y. Highly<br />

potent and specific inhibitors of human renin. Hypertension 1985; 7 (suppl.1):8–11.<br />

15. Kokubu T, Hiwada K, Nagae A, Murakami E, Morisawa Y, Yabe Y, Koike H, Iijima Y. Statine<br />

containing dipeptide and tripeptide inhibitors of human renin. Hypertension (Suppl.II) 1986; 8:1–5.<br />

16. Gelb MH, Svaren JP, Abeles RH. Fluoroketone inhibitors of hydrolytic enzymes. Biochemistry<br />

1985; 24:1813–1817.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_339.html (1 of 2) [4/5/2004 5:27:13 PM]


Document<br />

17. Szelke M. Chemistry of renin inhibitors. In: Kostka V, ed. Aspartic Proteinases and Their Inhibitors.<br />

Berlin: Walter de Gruyter, 1985:421–441.<br />

18. Sham HL, Stein HH, Rempel CA, Cohen J and Plattner JJ. Highly potent and specific inhibitors of<br />

human renin. FEBS Lett 1987; 220:299–301.<br />

19. Cooper JB, Foundling SI, Blundell TL, Boger J, Jupp R, Kay J. X-ray studies of aspartic proteinasestatine<br />

inhibitor complexes. Biochemistry 1989; 28:8596–8603.<br />

20. Arrowsmith RJ, Carter K, Dann JG, Davies DE, Harris CJ, Morton JA, Lister P, Robinson JA,<br />

Williams DJ. Novel renin inhibitors: synthesis of aminostatine and comparison with statine-containing<br />

analogues. J Chem Soc Chem Commun 1986; 10:755–757.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_339.html (2 of 2) [4/5/2004 5:27:13 PM]<br />

21. Jones M,<br />

Sueiras-Diaz<br />

J, Szelke M,<br />

Leckie B,<br />

Beattie S.<br />

Renin<br />

inhibitors<br />

containing<br />

the novel<br />

amino-acid 3aminodeoxystatine.<br />

In: Deber<br />

CM, Hru<strong>by</strong><br />

VJ, Kopple


Document<br />

KD eds. Peptides: <strong>Structure</strong> and Function. Rockford:Pierce Chemical Company, 1985:759–762.<br />

22. Rich DH, Sun ETO, Ulm E. Synthesis of analogues of the carboxyl proteinase inhibitor pepstatin.<br />

Effect of structure on the inhibition of pepsin and renin. J Med Chem 1980; 23:27–33.<br />

Page 340<br />

23. Sali A, Veerapandian B, Cooper JB, Founding SI, Hoover DJ, Blundell TL. High resolution X-ray<br />

diffraction study of the complex between endothiapepsin and an oligopeptide inhibitor: the analysis of<br />

inhibitor binding and description of the rigid body shifts in the enzyme. EMBO J 1989; 8:2179–2188.<br />

24. Iizuka K, Kamijo T, Kubota T, Akahane K, Umeyama H, Kiso Y. New human renin inhibitors<br />

containing an unnatural amino acid, norstatine. J Med Chem 1988; 31:701–704.<br />

25. Dann JG, Stammers DK, Harris, CJ, Arrowsmith RJ, Davies DE, Hardy GW, Morton JA. Human<br />

renin: an new class of inhibitors. Biochem Biophys Res Commun 1986; 134:71–77.<br />

26. Cooper JB, Foundling SI, Blundell TL, Arrowsmith RJ, Harris CJ, Champness JN. A rational<br />

approach to the design of antihypertensives: X-ray studies of complexes between aspartic proteinases<br />

and aminoalcohol inhibitors. In: Leeming PR, ed. Topics in Medicinal Chemistry. London: Royal<br />

Society of Chemistry, 1988; 308–313.<br />

27. Lunney EA, Hamilton HW, Hodges JC, Kaltenbrohn JS, Repine JT, Badasso M, Cooper J, Dealwis<br />

C, Wallace B, Lowther WT, Dunn BM, Humblet C. Analyses of ligand binding in five endothiapepsin<br />

crystal complexes and their use in the design and evaluation of novel renin inhibitors. J Med Chem<br />

1993; 36:3809–3820.<br />

28. Bartlett PA, Kezer WB. Phosphinic acid dipeptide analogues: potent, slowbinding inhibitors of<br />

aspartic proteinases. J Amer Chem Soc 1984; 106:4282–4283.<br />

29. Greenlee WJ. Renin inhibitors. Pharm Res 1987; 4(5):364–374.<br />

30. Thaisrivongs S, Pals DT, Harris DW, Kati WM, Turner SR. <strong>Design</strong> and synthesis of potent and<br />

specific renin inhibitors containing difluorostatine, difluorostatone and related analogues. J Med Chem<br />

1986; 29:2088–2093.<br />

31. Veerapandian B, Cooper JB, Sali A, Blundell TL. Direct observation <strong>by</strong> X-ray analysis of the<br />

tetrahedral “intermediate” of aspartic proteinases. Protein Science 1992; 1:322–328.<br />

32. Suguna K, Padlan EA, Smith CW, Carlson WD, Davies DR. Binding of a reduced peptide inhibitor<br />

to the aspartic proteinase from Rhizopus chinensis: implications for a mechanism of action. Proc Natl<br />

Acad Sci USA 1987; 84:7009–7013.<br />

33. Ptitsyn OB. Pure Appl Chem 1973; 31:227–244.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_340.html (1 of 2) [4/5/2004 5:27:15 PM]


Document<br />

34. Rosenberg SH, Plattner JJ, Woods KW, Stein HH, Marcotte PA, Cohen J, Perun TJ. Novel renin<br />

inhibitors containing analogues of statine retro-inverted at the C-termini: specificity of the P 2 histidine<br />

site. J Med Chem 1987; 30:1224–1228.<br />

35. Luly JR, Yi N, Soderquist J, Stein H, Cohen J, Perun TJ, Plattner JJ. New inhibitors of human renin<br />

that contain novel Leu-Val replacements. J Med Chem 1987; 30:1609–1616.<br />

36. Cooper J, Quail W, Frazao C, Foundling SI, Blundell TL. X-ray crystallographic analysis of<br />

inhibition of endothiapepsin <strong>by</strong> cyclohexyl renin inhibitors. Biochemistry 1992; 31:8142–8150.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_340.html (2 of 2) [4/5/2004 5:27:15 PM]


Document<br />

Page 341<br />

37. Attwood MR, Hassall CH, Krohn A, Lawton G, Redshaw S. The design and synthesis of angiotensin<br />

converting enzyme inhibitor cilazapril and related bicyclic compounds. J Chem Soc Perkin 1986;<br />

I:1011–1019.<br />

38. Szewczuk Z, Rebholz KL, Rich DH. Synthesis and biological activity of new conformationally<br />

restricted analogues of pepstatin. Int J Pept Res 1992; 40:233–242.<br />

39. Wood JM, Fuhrer W, Buhlmayer P, Riniker B, Hofbauer KG. Protection groups increase the in vivo<br />

stability of a statine-containing renin inhibitor. In: Kostka V, ed. Aspartic Proteinases and Their<br />

Inhibitors. Berlin: Walter de Gruyter, 1985:463–466.<br />

40. Bolis G, Fung AKL, Greer J, Kleinert HD, Marcotte PA, Perun TJ, Plattner JJ, Stein HH. Renin<br />

inhibitors. Dipeptide analogues of angiotensinogen incorporating transition-state, nonpeptidic<br />

replacements at the scissile bond. J Med Chem 1987; 30:1729–1737.<br />

41. Greenlee, WJ Renin inhibitors. Med Res Rev 1990; 10:173.<br />

42. Dhanaraj V, Dealwis C, Frazao C, Badasso M, Sibanda BL, Tickle IJ, Cooper JB, Driessen HPC,<br />

Newman M, Aguilar C, Wood SP, Blundell TL, Hobart PM, Geoghegan KF, Ammirati MJ, Danley DE,<br />

O'Connor BA, Hoover DJ. X-ray analyses of peptide-inhibitor complexes define the structural basis of<br />

specificity for human and mouse renins. Nature 1992; 357:466–472.<br />

43. Dealwis CG, Frazao C, Badasso M, Cooper JB, Tickle IJ, Driessen H, Blundell TL, Murakami K,<br />

Miyazaki H, Sueiras-Diaz J, Jones DM, Szelke M. X-ray analysis at 2.0 Å resolution of mouse<br />

submaxillary renin complexed with a decapeptide inhibitor CH-66, <strong>based</strong> on the 4–16 fragment of rat<br />

angiotensinogen. J Mol Biol 1994; 236:342–360.<br />

44. Wlodawer A, Miller M, Jaskolski M, Sathyanarayana BK, Baldwin E, Weber IT, Selk LM, Clawson<br />

L, Schneider J, Kent S. Conserved folding in retroviral proteinases: crystal structure of synthetic HIV-1<br />

proteinase. Science 1989; 245:616–621.<br />

45. Lapatto R, Blundell TL, Hemmings A, Overington J, Wilderspin A, Wood SP, Merson JR, Whittle<br />

PJ, Danley DE, Geoghegan KF, Hawrylik SJ, Lee SE, Scheld KG, Hobart PM. X-ray analysis of HIV-1<br />

proteinase at 2.7 Å resolution confirms structural homology among retroviral enzymes. Nature (Lond)<br />

1989; 342:299–302.<br />

46. Poe M, Wu JK, Lin TY, Hoogsteen K, Bull HG, Slater EE. Renin cleavage of a human-kidney renin<br />

substrate analogous to human angiotensinogen that is human renin specific and resistant to cathepsin D.<br />

Analyt Biochem 1984; 140:459–467.<br />

47. Cumin F, Lenguyen D, Castro B, Menard J, Corvol P. Comparative enzymatic studies of human<br />

renin acting on pure natural or synthetic substrates. Biochim Biophys Acta 1987; 913:10–19.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_341.html (1 of 2) [4/5/2004 5:27:17 PM]


Document<br />

48. Powers JC, Harley AD, Myers DV. Subsite specificity of porcine pepsin. In: Tang J, ed. Acid<br />

Proteases-<strong>Structure</strong>, Function and Biology. New York: Plenum Press, 1977:141–157.<br />

49. Antonov VK. In: Tang J, ed. Acid Proteases-<strong>Structure</strong>, Function and Biology. New York: Plenum<br />

Press, 1977:179.<br />

50. Skeggs LT, Lentz KE, Kahn JR, Hochstrasser H. Kinetics of the reaction of renin with nine synthetic<br />

peptide substrates. J. Exp Med 1968; 120:130–34.<br />

51. Holzman TF, Chung CC, Edalji R, Egan DA, Martin M, Gubbins EJ. Krafft GA, Wang GT, Thomas<br />

AM, Rosenberg SH, Hutchins C. Characterisation of recombi-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_341.html (2 of 2) [4/5/2004 5:27:17 PM]


Document<br />

nant human renin kinetics, pH-stability, and peptidomimetic inhibitor binding. J Protein Chem<br />

1991; 10:553–563.<br />

Page 342<br />

52. Tong L, Pav S, Lamarre D, Pilote L, Laplante S, Anderson PC, Jung G. High resolution crystalstructures<br />

of recombinant human renin in complex with polyhydroxymonoamide inhibitors. J Mol Biol<br />

1995; 250:211–222.<br />

53. Sham HL, Bolis G, Stein HH, Fesik SW, Marcott PA, Plattner JJ, Rempel CA, Greer J. Renin<br />

inhibitors. <strong>Design</strong> and synthesis of a new class of conformationally restricted analogues of<br />

angiotensinogen. J Med Chem 1988; 31:284–295.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_342.html [4/5/2004 5:27:19 PM]


Document<br />

Page 343<br />

14<br />

Structural Aspects in the Inhibitor <strong>Design</strong> of Catechol O-Methyltransferase<br />

Jukka Vidgren and Martti Ovaska<br />

Orion Corporation Orion Pharma, Espoo, Finland<br />

I. Introduction<br />

Catechol O-methyltransferase (COMT) plays an important role in the catabolic inactivation of<br />

catecholamines. It is present both in extracerebral tissues and in the central nervous system. During the<br />

last few years there has been a remarkable interest in COMT. Basic biochemical and molecular biology<br />

research has given detailed insights into the function and nature of the enzyme. The knowledge of the<br />

crystallographic structure has allowed researchers to analyze the molecular mechanism of the catalytic<br />

reaction and to accomplish the structure-<strong>based</strong> design of inhibitors. The development of potent and<br />

selective inhibitors has provided effective pharmacological tools to investigate the physiological role of<br />

the enzyme. The main clinical interest has been the possible application of COMT inhibitors as adjuncts<br />

in the L-dopa therapy of Parkinson's disease. Parkinson's disease is a dopamine deficiency disorder. The<br />

dopamine-producing neurons in striatum are destroyed. The medication strategy is to replenish the<br />

missing dopamine. L-Dopa, given together with a peripheral inhibitor of dopa decarboxylase (DDC), for<br />

example, carbidopa, is a standard therapy in Parkinson's disease. While dopamine does not penetrate<br />

into the brain, L-dopa penetrates the blood-brain barrier and is decarboxylated into dopamine in the<br />

brain. The half-life of L-dopa is short and in the presence of DDC inhibitor a large amount of the drug is<br />

eliminated <strong>by</strong> COMT. The COMT enzyme produces the metabolite 3-methoxytyrosine (3-OMD), which<br />

has no benefit in the treatment of Parkinson's disease, but has a long elimination half-life and may be<br />

harmful during chronical treatment. Also a gradual loss of the efficacy of L-dopa occurs during longterm<br />

medication. Since the early 1980s active<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_343.html [4/5/2004 5:27:21 PM]


Document<br />

Figure 1<br />

The rationale of COMT inhibition as adjunct in the L-dopa therapy of<br />

Parkinson's disease (reproduced <strong>by</strong> permission from COMT News, Issue 1, Orion<br />

Corporation, Orion Pharma, 1994).<br />

Page 344<br />

research in pharmaceutical companies has been carried out to develop new potent, selective, and orally<br />

active COMT inhibitors. Some of them (e.g., entacapone), are now in final clinical trials and the results<br />

have been promising. The rationale of COMT inhibition can be seen in Figure 1. It can be concluded<br />

that COMT inhibition in peripheral tissues improves the brain entry of L-dopa and decreases the<br />

formation of 3-OMD. The dose of L-dopa can be lowered and the dose interval prolonged. Also a<br />

decrease of the fluctuations of dopamine formation has been observed. The inhibition of COMT seems<br />

to be the next step in improving the L-dopa therapy of Parkinson's disease. This paper discusses the<br />

structure-<strong>based</strong> approach for the understanding of the enzyme function and inhibitor design.<br />

II. The Enzyme<br />

A. Physiological Role of COMT<br />

Catechol O-methyltransferase (COMT, EC 2.1.1.6) was originally detected in rat liver extracts [1]. Since<br />

then, COMT has been found in many species:<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_344.html (1 of 2) [4/5/2004 5:27:34 PM]


Document<br />

Figure 2<br />

The reaction catalyzed <strong>by</strong> catechol O-methyltransferase. Dopamine:<br />

R=CH 2-CH 2-NH 2; L-dopa: R=CH 2-CH(NH 2)-COOH.<br />

animals, plants, and procaryotes [2]. In mammals the highest COMT activities have been found in the<br />

liver and kidney, but COMT is common in almost all mammalian tissues [2–4].<br />

Page 345<br />

The COMT enzyme catalyzes the transfer of the methyl group from the coenzyme S-adenosyl-Lmethionine<br />

(AdoMet) to one of the phenolic hydroxyl groups of a catechol or substituted catechol [1]<br />

(Figure 2). The presence of magnesium ions is required for the catalysis. The reaction products are Omethylated<br />

catechol and S-adenosyl-L-homocysteine (AdoHcy). Physiological substrates of COMT are<br />

catecholamine neurotransmitters, dopamine, noradrenaline, and adrenaline, and some of their<br />

metabolites. The COMT enzyme inactivates catecholic steroids such as 2-hydroxyestradiol, drugs with a<br />

catechol structure such as L-dopa, and a large number of other catechol compounds [1,2,5–7]. The<br />

general physiological function of COMT is the inactivation of biologically active or toxic catechols. A<br />

schematic view of the major catecholamine pathways in the brain is shown in Figure 3. L-Dopa is the<br />

dopamine precursor used in the treatment of Parkinson's disease [8].<br />

B. Primary <strong>Structure</strong>s<br />

There are no isoenzymes of COMT known in different mammalian tissues. Two distinct forms of<br />

COMT have been found: one is soluble (S-COMT) and the other membrane bound (MB-COMT) [9,10].<br />

Both soluble and membrane-bound COMT have been cloned and characterized [11–16]. The soluble and<br />

membrane-bound COMT are coded <strong>by</strong> one gene using two separate promoters [17]. The soluble COMT<br />

contains 221 amino acids, whereas the membrane-bound form has a 50-(human) or 43-(rat) residueslong<br />

amino-terminal extension containing the hydrophobic membrane anchor region. The sequences of<br />

COMT enzymes from different species are highly similar (see Figure 4). The soluble human protein is<br />

81% identical with the rat enzyme. The 165-amino-acids-long fragment of porcine COMT has 82%<br />

homology with the human<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_345.html [4/5/2004 5:27:37 PM]


Document<br />

Figure 3<br />

The main metabolic routes of dopamine and noradrenaline in<br />

the brain. COMT, Catechol O-methyltransferase; MAO, monoamino<br />

oxidase; DDC, dopa decarboxylase; DBH, dopamine<br />

β-hydroxylase; 3-OMD, 3-methoxytyrosine; Dopac,<br />

dihydroxyphenyl acetic acid.<br />

Page 346<br />

enzyme [13]. The existence of a thermolabile low-activity and a thermostable high-activity COMT in<br />

human population has been reported [18]. Interestingly, the two published sequences of human soluble<br />

COMT differ in only one amino acid. Recent kinetic studies have shown that this difference affects<br />

unambiguously the thermostability of the enzyme [19].<br />

C. Kinetics of Human COMT<br />

The kinetic mechanism of the methylation reaction of human COMT has been studied exhaustively<br />

using recombinant enzymes [19]. The mechanism is sequential ordered: AdoMet binding first, then<br />

Mg 2+ and the catechol substrate as the last ligand. Human S-COMT and MB-COMT have similar kinetic<br />

properties. The main difference is the one-order lower K m value of MB-COMT for dopamine as<br />

substrate (S-COMT 207 μM and MB-COMT 15 μM). The COMT enzyme is a rather slow enzyme with<br />

a low catalytic number. At saturating substrate levels S-COMT has a double efficiency compared with<br />

MB-COMT (k cat=37 and k cat =17, respectively). At low substrate concentrations (


Document<br />

Figure 4<br />

Amino acid sequence comparison of known COMT sequences (rat [11], human<br />

[14], pig [13]). Secondary structure elements in the sequence of rat soluble COMT<br />

are indicated as well as important active-site residues involved in binding of ligands<br />

(a, AdoMet; m, magnesium; s, substrate/inhibitor). The numbering of the residues<br />

corresponds to the soluble enzyme. The extension of the MB-COMT consists of the first<br />

50 amino terminal residues. In the human sequence of COMT determined <strong>by</strong> Bertocci<br />

[13], Val108 is replaced <strong>by</strong> Met108.<br />

Under physiological concentrations of catecholamines in the brain, MB-COMT may play a more<br />

important role than S-COMT [19,20].<br />

D. Three-Dimensional <strong>Structure</strong> of COMT<br />

Backbone<br />

Page 347<br />

The crystal structure of rat soluble COMT has been solved at 2.0 Å resolution [21]. The COMT enzyme<br />

has a single domain α/β-folded structure, in which eight α-helices are arranged around the the central<br />

mixed β sheet. The sheet contains five parallel β strands and one antiparallel β hairpin. An overview of<br />

COMT is illustrated in Figure 5.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_347.html (1 of 2) [4/5/2004 5:27:50 PM]


Document<br />

The AdoMet binding motif is similar to the Rossmann fold, which is well known from the nucleotide<br />

binding proteins [22]. It has been shown that the known crystal structures of methyltransferases are<br />

strikingly similar in the AdoMet-binding regions [23], which indicates that all AdoMet-utilizing<br />

enzymes may share a common divergent evolution.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_347.html (2 of 2) [4/5/2004 5:27:50 PM]


Document<br />

Figure 5<br />

Schematic stereo view of the three-dimensional structure of COMT. The ligands<br />

bound to COMT are the methyl-donating coenzyme AdoMet and the magnesium ion.<br />

Figures 5–7 and 13 were produced using the program MOLSCRIPT [50].<br />

Figure 6<br />

Stereo view of the AdoMet binding to COMT. The most important amino acid<br />

residues are shown as well as the magnesium ion and the inhibitor<br />

3, 5-dinitrocate-chol (DNC).<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_348.html (1 of 2) [4/5/2004 5:28:03 PM]<br />

Page 348


Document<br />

Active Site<br />

The active site of COMT consists of the AdoMet binding domain and the catalytic site. The structural<br />

elements and individual interactions between COMT residues and the enzymatic-action-participating<br />

ligands are demonstrated in detail in Figures 5 to 8.<br />

Page 349<br />

AdoMet Binding <strong>by</strong> COMT. The active-site residues, which have significant interactions with the<br />

coenzyme, are shown in Figure 6. The loop region between strand β1 and helix α4 forms the AdoMetbinding<br />

consensus sequence (in COMT GAxxG) that is conserved in methyltransferases [23]. In this<br />

region the terminal amino and carboxyl groups of AdoMet are bound. The last residue of the strand β2,<br />

Glu90, forms a hydrogen bond to the ribose hydroxyls. The residue Met91 has face-to-face van der<br />

Waals contacts on one side, and His142 has edge-to-face contacts on the opposite side of the adenine<br />

ring. The residue Trp143 closes the adenine of AdoMet into the protein with face-to-edge contact.<br />

Furthermore, the N-6 atom of the adenine hydrogen binds to Ser119. The Met40 residue holds the<br />

sulphur of AdoMet with the methyl group in right position towards the hydroxyl group of the catechol<br />

substrate. As a result of the various hydrogen bonds and van der Waals contacts, AdoMet has a high<br />

affinity to COMT with a dissociation constant of 23 μM [19].<br />

Catalytic Site. The catalytic site of COMT is a rather simple environment formed <strong>by</strong> the metal ion and<br />

<strong>by</strong> the amino acids important for substrate binding and catalysis of the methylation reaction.<br />

The magnesium ion plays a crucial role for the catalytic activity of COMT. Figure 7 shows the binding<br />

of magnesium to COMT as derived from the<br />

Figure 7<br />

Magnesium binding in COMT. The magnesium ligands are Asp141, Asp169, Asn170,<br />

both hydroxyls of 3, 5-dinitrocatechol (DNC) and a water molecule (W).<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_349.html (1 of 2) [4/5/2004 5:28:09 PM]


Document<br />

Figure 8<br />

The catalytic machinery of COMT. Shown are the COMT residues<br />

important for catalysis, Mg 2+ binding, the catechol as substrate, and the<br />

methyl-donating coenzyme AdoMet. The hydrophobic walls are defined<br />

<strong>by</strong> two tryptophane residues and a proline residue.<br />

Page 350<br />

crystallographic studies. Magnesium has an octahedral coordination to two aspartic acid residues<br />

(Asp141 and Asp169), to an asparagine residue (Asn170), to both catechol hydroxyls of the substrate,<br />

and to a water molecule. In addition to the Mg 2+ ion, Lys144 and Glu199 participate directly in the<br />

methylation reaction as shown in Figure 8. The “gate keeper” residues Trp38 and Pro174 form the<br />

hydrophobic “walls” and define the selectivity of the enzyme to different side chains of the substrate.<br />

They play a significant role in the binding of the substrates and inhibitors of COMT [19, 21].<br />

III. Mechanism of the Catalytic Action of COMT<br />

The catalytic site is a shallow groove with the catalytic machinery at the bottom as illustrated in Figure<br />

8. The two hydroxyl oxygens of a catechol substrate bind directly to the Mg 2+ ion. The active methyl<br />

group of AdoMet is near one of the hydroxyl groups, on one side of the catechol ring. The amino group<br />

of Lys144 is also located near this hydroxyl group, on the other side of the catechol ring from AdoMet.<br />

The Glu199 residue is near the other hydroxyl group.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_350.html [4/5/2004 5:28:14 PM]


Document<br />

Figure 9<br />

The energy profile (as calculated at the 3-21G/PM3 level [24]) for the proton<br />

transfer step A rarrow.gif B and the methyl transfer step B rarrow.gif C for catechol as a<br />

substrate.<br />

Page 351<br />

The pK a of the catecholic hydroxyl is about 9.8. The role of the Mg 2+ ion bound to the enzyme is to<br />

make the hydroxyl groups more easily ionizable. It has been shown <strong>by</strong> quantum mechanical calculations<br />

that the hydroxyl protons can be transferred to Lys144 and Glu199 [24]. The proton transfer OH<br />

rarrow.gif Lys144 activates the hydroxyl group for the methyl transfer AdoMet rarrow.gif O -. Thus<br />

the reaction coordinate for methylation of catechols <strong>by</strong> COMT consists of a proton transfer from a<br />

hydroxyl group to Lys144 and a subsequent methyl transfer from AdoMet to the hydroxyl group (Figure<br />

9). The Lys144 residue acts as a typical catalytic base in a general base-catalysed SN2-like nucleophilic<br />

substitution reaction [24].<br />

IV. Inhibitors of COMT<br />

A. First-Generation Inhibitors<br />

First generation COMT inhibitors such as pyrogallol, U-0521 (3, 4-dihydroxy-2-methylpropiophenone),<br />

tropolone, and 8-hydroxyquinoline (Figure 10) were<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_351.html (1 of 2) [4/5/2004 5:28:20 PM]


Document<br />

Figure 10<br />

<strong>Structure</strong>s of some first generation COMT inhibitors: pyrogallol,<br />

U-0521 (3, 4-dihydroxy-2-methylpropiophenone), tropolone,<br />

and 8-hydroxyquinoline.<br />

Figure 11<br />

<strong>Structure</strong>s of second-generation COMT inhibitors discussed in this<br />

paper: nitecapone (OR-462), entacapone (OR-611), tolcapone<br />

(RO-40-7592), and 2-((3, 4-dihy-droxy-2-nitrophenyl) vinyl) phenylketone.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_352.html (1 of 2) [4/5/2004 5:28:27 PM]<br />

Page 352


Document<br />

Table 1 Inhibitor Constants of Selected COMT Inhibitors<br />

K i (nM) Human b K i (nM)Pig c<br />

Nitecapone 1.0 700<br />

Entacapone 0.3 N.D.<br />

Tolcapone 0.3 N.D.<br />

Vinylphenylketone 4.0 a 200<br />

a T. Lotta, unpublished results<br />

b Reference 19.<br />

c Reference 41.<br />

N.D.: not determined<br />

Page 353<br />

used as in vitro tools to investigate COMT inhibition, but because of the lack of potency and selectivity<br />

and because they were toxic, they were not clinically useful [2]. These inhibitors have inhibition<br />

constants (K i values) in the micromolar range. Many of them contain the catechol structure and are also<br />

substrates of COMT.<br />

B. Second-Generation Inhibitors<br />

The invention of a new structural family of COMT inhibitors in the late 1980s lead the COMT research<br />

into a new active epoch [25,26]. The most potent second-generation inhibitors are nitrocatechol<br />

derivatives; some examples are shown in Figure 11. Entacapone (OR-611) and tolcapone (RO-40-7592)<br />

are now in clinical trials for the treatment of Parkinson's disease, and have been extensively studied<br />

[27–40]. Both compounds are very potent and selective tightbinding inhibitors of human COMT with K i<br />

values of 0.3 nM [19]. They differ mainly in their pharmacokinetic properties. Entacapone acts<br />

peripherally while tolcapone inhibits COMT both peripherally and centrally.<br />

Recently it was reported that 2-((3,4-dihydroxy-2-nitrophenyl)vinyl) phenylketone is a tight-binding<br />

inhibitor of pig COMT [41], and it is also a potent inhibitor of human COMT (Table 1). This inhibitor<br />

has a vinylphenylketone group at position 4 of the catechol ring, i.e., in the position ortho to the nitro<br />

group.<br />

V. Enzyme Inhibitor Interactions<br />

A. Background<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_353.html (1 of 2) [4/5/2004 5:28:31 PM]


Document<br />

From quantitative structure activity relationship studies (QSAR) of COMT inhibitors it became evident<br />

that the acidity of the catechol hydroxyl group is the most important factor that influences the inhibitory<br />

activity of catechol<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_353.html (2 of 2) [4/5/2004 5:28:31 PM]


Document<br />

derivatives [26,42]. <strong>Structure</strong>s containing a catechol ring optimally substituted with a nitro group in<br />

position 3 and other electron-withdrawing substituents in position 5 showed high-potency inhibition<br />

[25,26,42,43]. It was also detected that the side-chain hydrophobicity at position 5 correlates<br />

significantly with the inhibitor activity [42].<br />

B. X-Ray <strong>Structure</strong>s of COMT-<strong>Drug</strong> Complexes<br />

Page 354<br />

The structure of 3,5-dinitrocatechol complexed with COMT has been solved [21]. This inhibitor is a<br />

typical nitrocatechol derivative with a high affinity for COMT. The excellent electron density of the<br />

inhibitor in the active site of COMT is represented in Figure 12. The planar structure of this compound<br />

fits well into the active-site cavity of the enzyme and forms nearly ideal contacts<br />

Figure 12<br />

A portion of the structure model and the 2F 0-F c electron<br />

density map contoured at 1.0 standard deviation. The region<br />

containing the inhibitor, magnesium, parts of AdoMet, and<br />

the residue Glu199 is shown. The sphere marked with “W”<br />

represents a water molecule.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_354.html (1 of 2) [4/5/2004 5:28:46 PM]


Document<br />

Page 355<br />

with the tryptophane residues 38 and 143. The hydroxyl groups of the inhibitor are coordinated to the<br />

Mg 2+ ion. The hydroxyl group in position 1 has an important hydrogen bond to the carboxyl group of<br />

Glu199. The other hydroxyl is near the methyl group donated <strong>by</strong> AdoMet. The pK a of this hydroxyl<br />

group is low (about 3.4) [26]. The 3-nitro group of the inhibitor has favourable van der Waals<br />

interactions with Trp143. The Trp38 residue is located edge-to-face with the catechol plane, which<br />

allows an ideal aromatic hydrophobic contact. Such aromatic hydrophobic interactions have been<br />

described to be important in proteins and for the binding of ligands [44,45]. The binding mode of the<br />

catechol ring of other crystallographically determined potential nitrocatecholtype inhibitors complexed<br />

with COMT is essentially the same as with 3,5- dinitrocatechol (J.Vidgren, unpublished results).<br />

C. Differences in the Active Site of Human, Rat, and Pig COMT<br />

Models for human and pig COMT are easy to build using the experimental structure of the rat COMT,<br />

due to the high degree of homology between the rat, human, and pig COMT enzymes (Figure 4). The<br />

active sites are especially well conserved—the few differences in the active-site residues are collected in<br />

Table 2. The kinetic data show that the K m values of common substrates for rat and human COMT are<br />

very similar. Pig COMT shows, however, a considerably higher K m value for catechol [46]. The same<br />

difference is apparent for inhibitors represented <strong>by</strong> the K i values in Table 1.<br />

The model for the binding of vinylphenylketone to pig and rat COMT is shown in Figure 13. Assuming<br />

that the catechol part of the inhibitor adopts the same position as found in the crystal structure with<br />

dinitrocatechol, the vinylphenylketone substituent has enough room to bind to both enzymes. The most<br />

significant difference between these enzymes lies in residue 38, the hydrophobic tryptophan in rat (and<br />

human) COMT and the polar arginine in pig COMT. If Arg38 is directed towards the hydrophobic core<br />

of the enzyme in a similar conformation as Trp38 (shown in Figure 13), it causes repulsion with the<br />

catechol ring of the inhibitor. However, it is probable that the polar Arg38 is directed towards the<br />

solvent. In this case the substrates and inhibitors will lack the favorable contacts that exist with Trp38 in<br />

human and rat enzymes. Obvi-<br />

Table 2 Differences in the Active Sites Between Rat, Human, and Pig COMT<br />

Position Rat Human Pig<br />

38 Trp Trp Arg<br />

173 Val Cys Cys<br />

201 Met Arg Ser<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_355.html [4/5/2004 5:28:47 PM]


Document<br />

Figure 13<br />

The inhibitor 2-((3,4-dihydroxy-2-nitrophenyl)vinyl)phenylketone modeled<br />

into the active site of rat (a) and pig (b) COMT.<br />

Page 356<br />

ously this one amino acid differences in the active site of the isoenzymes can explain the significant<br />

differences in the inhibitory potency of vinylphenylketone against these enzymes. From the structural<br />

point of view it seems that the position of the side-chain substitution (for example at C5 in entacapone<br />

and C4 in vinylphenylketone) is not critical for the inhibitor binding. In both cases the substituent has<br />

sufficient space to adapt to the protein structure, and in fact, large substituents reach from the active site<br />

cavity to the solvent (Figure 13).<br />

The tenfold higher inhibitory activity of entacapone compared with vinylphenylketone against human<br />

COMT can be accounted for <strong>by</strong> the electron- withdrawing effect of the side-chain substitution. In the<br />

case of entacapone, the side chain at position C5 has a more beneficial electronic influence on the 2-<br />

hydroxyl of the inhibitor producing a better inhibition (see Section VI).<br />

VI. Mechanism of the COMT Inhibition By Nitrocatechols<br />

As described above, catechols with strong electronegative groups are potent inhibitors of COMT. These<br />

compounds seem to bind well to the active site, but in spite of that they are very poor substrates. It has<br />

been shown, with nitecapone<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_356.html (1 of 2) [4/5/2004 5:28:54 PM]


Document<br />

Figure 14<br />

The energy profile (as calculated at the 3-21G//PM3 level [24]) for the proton<br />

transfer step A rarrow.gif B and the methyl transfer step B rarrow.gif C for 3,4-dinitrocatechol as a<br />

substrate.<br />

in a rat COMT assay, that the rate of nitecapone methylation is equivalent to about 1% of the rate of<br />

dopamine methylation [47].<br />

Page 357<br />

The energy profile for the hypothetical methylation of 3,5-dinitrocatechol is shown in Figure 14 [24].<br />

The electronegative nitro groups strongly stabilize the ionized catechol–COMT complex, and the energy<br />

barrier for the methylation step is high (see Figure 9 for comparison). This can be understood as<br />

decreased nucleophilicity of the hydroxyl oxygen, due to the electron-with- drawing properties of the<br />

nitro groups.<br />

The electronic effect of the substituents of the catechol ring to the nucleophilicity of the hydroxyl group<br />

at the active site can be readily seen from the molecular electrostatic potential (MEP) surfaces of the<br />

system. The MEP surfaces were calculated at the PM3 level and plotted at –20 kcal/mol for catechol and<br />

3,5-dinitrocatechol at the active site of COMT [24]. The results are summarized in Figure 15. In the case<br />

of catechol the effect of the proton transfer form OH to Lys144 is seen as a remarkable increase in the<br />

negative potential between AdoMet and the substrate. 3,5-Dinitro substitution of the catechol ring<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_357.html (1 of 2) [4/5/2004 5:28:58 PM]


Document<br />

Figure 15<br />

The energy profiles and MEP surfaces (–20 kcal/mol) for catechol (black) and<br />

3,5-dinitrocatechol (grey).<br />

decreases the activating negative potential substantially. As a consequence, the OH group is weakly<br />

nucleophilic and 3,5-dinitrocatechol is not a substrate of COMT but a potent inhibitor.<br />

VII. Inhibitor <strong>Design</strong><br />

Page 358<br />

The drug-design process of COMT inhibitors started long before the structure of the target molecule was<br />

available. The most important results were extracted from the QSAR studies of substituted catechols<br />

[26,42]. Those investigations clearly indicated the importance of the acidity of one of the two hydroxyl<br />

groups in the catechol ring. The ionization of the hydroxyl was greatly influenced <strong>by</strong> electronwithdrawing<br />

substituents in the positions ortho and para to the hydroxyl. The lipophilicity of the side<br />

chain was predicted, but after the determination of the enzyme structure it became clear that the active<br />

site of COMT is a relatively shallow groove where ligands with longer side chains<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_358.html [4/5/2004 5:29:56 PM]


Document<br />

Page 359<br />

easily reach the surface of the enzyme. The pharmacophore model constructed with biochemical and<br />

QSAR knowledge [42] was surprisingly good and realistic in comparison to the experimental structure.<br />

The crucial role of the magnesium ion for the binding of substrates and catalysis was not given enough<br />

attention. Even without the experimental knowledge of the three-dimensional structure of COMT, the<br />

correct decision for the direction of the drug-design process was possible. Many open questions were<br />

still waiting for the determination of the molecular structure. The most important drug-design aspects<br />

under consideration included the optimization of the pharmacokinetic properties of the molecules,<br />

especially the penetration across the blood brain barrier. The structure of COMT clearly showed that<br />

with catechol-type inhibitors, one of the positions ortho to the hydroxyls is optimally substituted <strong>by</strong> a<br />

nitro group, while the other ortho position has to be unsubstituted. Sterically, the two remaining sites can<br />

be substituted quite freely, so that these substitution positions can be used to modify the<br />

physicochemical properties of the inhibitors. The question of the possibility of designing an inhibitor<br />

without the catechol structure is important, because the potent inhibitors which rely on the ionization of<br />

a catechol hydroxyl, penetrate very poorly into the brain. In clinical trials of the therapeutic use of<br />

COMT inhibitors it has become evident that the beneficial effect to L- dopa metabolism is fully reached<br />

with peripheral inhibitors such as entacapone [27–29]. As demonstrated above, <strong>based</strong> on the threedimensional<br />

structure of COMT, the design of potent noncatechol type inhibitors may be very tedious.<br />

The active site of COMT is a rather simple environment with a few catalytic residues and a magnesium<br />

ion defining the structural limits of the catechol ligands.<br />

VIII. Clinical Possibilities of Inhibitors of COMT<br />

A. Parkinson's Disease<br />

Parkinson's disease is a neurological disorder that affects voluntary movement. The symptoms are<br />

slowness of movement, rigidity, and tremor. The reason for the disease is unknown. Parkinson's disease<br />

is a progressive disorder and involves the deterioration of dopaminergic nerve fibers in substantia nigra,<br />

which leads to a striatal deficiency of dopamine. The symptoms of Parkinson's disease are detected only<br />

after about 80% of the dopamine-producing neurons are degenerated. There is no known cure but the<br />

symptoms are treated with a combination of different drugs.<br />

Current and Future Therapy of Parkinson's Disease<br />

Dopamine does not penetrate the blood-brain barrier. L-Dopa is actively transported into the brain and<br />

then converted to dopamine. L-Dopa is rapidly metabolized and only about 1% of an oral dose reaches<br />

the brain. In the current<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_359.html [4/5/2004 5:30:04 PM]


Document<br />

Page 360<br />

therapy of Parkinson's disease L-dopa (precursor of dopamine) is administered together with a<br />

peripheral dopadecarboxylase (DDC) inhibitor (such as carbidopa or benserazide). The major peripheral<br />

L-dopa metabolizing enzymes are DDC and COMT. After inhibition of DDC, COMT is responsible for<br />

the main catabolism of L-dopa. In the central nervous system COMT together with mono- amino<br />

oxidase (MAO) participitates in the metabolism of L-dopa and dopamine. Large amounts of orally<br />

administered L-dopa are converted <strong>by</strong> COMT to 3-O-methyldopa (3-OMD). Having a long plasma halflife<br />

(approximately 15 h compared with the dopamine 1-h half-life), 3-OMD accumulates in the plasma<br />

during the L-dopa treatment. L-Dopa and 3-OMD also compete for the same active transport system into<br />

the brain. It has been proposed that 3-OMD could cause some side effects of the L-dopa treatment<br />

(dyskenisia, on-off phenomenon). Inhibition of COMT enzyme decreases the 3-OMD formation and<br />

improves the brain entry and bioavailability of L-dopa. The use of COMT inhibition should prolong the<br />

L-dopa effects and permit a decreased does [27– 29].<br />

Preclinical and clinical results indicate that both entacapone and tolcapone are orally active, nontoxic<br />

and well-tolerated drugs. The adjuvant L- dopa therapy with DDC inhibitor + COMT-inhibitor (+<br />

possible MAO inhibitor) may substitute for the present double therapy in the treatment of Parkinson's<br />

disease [27–40]. Together with the development of dopamine agonists and MAO inhibitors, the<br />

inhibition of COMT will constitute major progress in the treatment of Parkinson's disease in the near<br />

future.<br />

B. Other Possible Indications of the COMT Inhibition<br />

It has been proposed that COMT inhibitors co-administered with L-dopa could have beneficial effects in<br />

the treatment of depressive illness [40]. This can be caused <strong>by</strong> either the better availability of dopamine<br />

or <strong>by</strong> the elevated noradrenaline levels in the brain. Another hypothesis suggests that the increasing level<br />

of AdoMet caused <strong>by</strong> COMT inhibition may cause an antidepressive effect [33].<br />

Dopamine has also natriuretic and diuretic effects in kidney. There has been evidence that abnormalities<br />

of the renal dopamine system can lead to salt- sensitive hypertension [48]. In rat kidney, deamination<br />

represents the major pathway in the metabolism of dopamine, but when MAO is inhibited, methylation<br />

appears to offer an alternative metabolic pathway [49]. Thus COMT inhibition may be important in the<br />

regulation of renal sodium excretion.<br />

References<br />

1. Axelrod J, Tomchick R. Enzymatic O-methylation of epinephrine and other catechols. Journal of<br />

Biological Chemistry 1958; 233:702–705.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_360.html [4/5/2004 5:30:11 PM]


Document<br />

Page 361<br />

2. Guldberg H, Marsden C. Catechol-O-methyl transferase: pharmacological aspects and Physiological<br />

role. Pharmacological Reviews 1975; 27:135–206.<br />

3. Rivett AJ, Francis A, Roth JA. Localization of membrane-bound catechol-O- methyltransferase.<br />

Journal of Neurochemistry 1983; 40:1494–1496.<br />

4. Karhunen T, Tilgmann C, Ulmanen I, Julkunen I, Panula P. Distribution of catechol-Omethyltransferase<br />

enzyme in rat tissues. Journal of Histochemistry and Cytochemistry 1994;<br />

42:1079–1090.<br />

5. Axelrod J. Methylation reactions in the formation and metabolism of catecholamines and other<br />

biogenic amines. Pharmacological Reviews 1966; 18:95– 113.<br />

6. Ball P, Knuppen R, Haupt M, Breuer H. Interactions between estrogens and catechol amines III.<br />

Studies on the methylation of catechol estrogens, catechol amines and other catechols <strong>by</strong> the catechol-Omethyltransferase<br />

of human liver. Journal of Clinical Endocrinology 1972; 34:736–746.<br />

7. Borchardt RT. N- and O-methylation. In: Jako<strong>by</strong> WB, ed. Enzymatic Basis of Detoxification. Vol. 2.<br />

New York: Academic Press, 1980:43–62.<br />

8. Nutt JG, Fellman JH. Pharmacokinetics of levodopa. Clinical Neuropharmacology 1984; 7:35–79.<br />

9. Assicot M, Bohuon C. Presence of two distinct catechol-O-methyltransferase activities in red blood<br />

cells. Biochimie 1971; 53:871–874.<br />

10. Nissinen E, Männistõ P. Determination of catechol-O-methyltransferase activity <strong>by</strong> high<br />

performance liquid chromatography with electrochemical detection. Analytical Biochemistry 1984;<br />

137:69–73.<br />

11. Salminen M, Lundstrõm K, Tilgmann C, Savolainen R, Kalkkinen N, Ulmanen I. Molecular cloning<br />

and characterization of rat liver catechol-O-methyltransferase. Gene 1990; 93:241–247.<br />

12. Tilgmann C, Kalkkinen N. Purification and partial characterization of rat liver soluble catechol-Omethyltransferase.<br />

FEBS Letters 1990; 264:95–99.<br />

13. Bertocci B, Garotta G, Da Prada M, et al. Immunoaffinity purification and partial amino acid<br />

sequence analysis of catechol-O-methyltransferase from pig liver. Biochimica et Biophysica Acta 1991;<br />

1080:103–109.<br />

14. Lundström K, Salminen M, Jalanko A, Savolainen R, Ulmanen I. Cloning and characterization of<br />

human placental catechol-O-methyltransferase cDNA. DNA and Cell Biology 1991; 10:181–189.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_361.html (1 of 2) [4/5/2004 5:30:23 PM]


Document<br />

15. Tilgmann C, Kalkkinen N. Purification and partial sequence analysis of the soluble catechol-Omethyltransferase<br />

from human placenta: Comparison to the rat liver enzyme. Biochemical and<br />

Biophysical Research Communications 1991; 174:995–1002.<br />

16. Lundström K, Tilgmann C, Peränen J, Kalkkinen N, Ulmanen I. Expression of enzymatically active<br />

rat liver and human placental catechol-O-methyltransferase in Escherichia coli; purification and partial<br />

characterizaton of the enzyme. Biochimica et Biophysica Acta 1992; 1129:149–154.<br />

17. Tenhunen J, Salminen M, Jalanko A, Ukkonen S, Ulmanen I. <strong>Structure</strong> of the rat catechol-Omethyltransferase<br />

gene: separate promoters are used to produce mRNAs for soluble and membranebound<br />

forms of the enzyme. DNA and Cell Biology 1993; 12:253–263.<br />

18. Boudikova B, Szumlanski C, Maidak B, Weinshilboum R. Human liver catechol- Omethyltransferase<br />

pharmacogenetics. Clinical Pharmacology and Therapeutics 1990; 48:381–389.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_361.html (2 of 2) [4/5/2004 5:30:23 PM]


Document<br />

Page 362<br />

19. Lotta T, Vidgren J, Tilgmann C, et al. Kinetics of human soluble and membrane- bound catechol-Omethyltransferase:<br />

a revised mechanism and description of the thermolabile variant of the enzyme.<br />

Biochemistry 1995; 34:4202–4210.<br />

20. Roth JA. Membrane-bound catechol-O-methyltransferase: A reevaluation of its role in the Omethylation<br />

of the catecholamine neurotransmitters. Reviews of Physiology, Biochemistry and<br />

Pharmacology 1992; 120:1–29.<br />

21. Vidgren J, Svensson LA, Liljas A. Crystal structure of catechol-O-methyltransferase. Nature 1994;<br />

368:354–358.<br />

22. Rossman M, Liljas A, Bränden C I, Banaszak L. Evolutionary and structural relationships among<br />

dehydrogenases. In: Boyer PD, ed. Enzymes. New York: Academic Press, 1975:61–102.<br />

23. Schluckebier G, O'Gara M, Saenger W, Cheng X. Universal catalytic domain structure of adometdependent<br />

methyltransferases. Journal of Molecular Biology 1995; 247:16–20.<br />

24. Ovaska M. The mechanism of catalysis and inhibition of catechol-O-methyltransferase. Submitted<br />

1996.<br />

25. Bäckström R, Honkanen E, Pippuri A, et al. Synthesis of some novel potent and selective catechol-Omethyltransferase<br />

inhibitors. Journal of Medicinal Chemistry 1989; 32:841–846.<br />

26. Borgulya J, Bruderer H, Bernauer K, Zurcher G, Da Prada M. Catechol-O- methyltransferaseinhibiting<br />

pyrocatechol derivatives: synthesis and structure- activity studies. Helvetica Chimica Acta<br />

1989; 72:952–968.<br />

27. Nutt JG, Woodward WR, Beckner RM, et al. Effect of peripheral catechol-O- methyltransferase<br />

inhibition on the pharmacokinetics and pharmacodynamics of levodopa in parkinsonian patients.<br />

Neurology 1994; 44:913–919.<br />

28. Ruottinen H, Rinne UK, Ahtila S, Karlsson M, Kyyrä T, Gordin A. Entacapone increases levodopa<br />

response in a one-month double-blind study in parkinsonian patients with fluctuations. Neurology 1995;<br />

45:412S.<br />

29. Ruottinen H, Rinne UK. Entacapone prolongs levodopa response in a one-month double-blind study<br />

in parkinsonian patients with levodopa related fluctuations. Journal of Neurology, Neurosurgery, and<br />

Psychiatry 1996; 60:36–40.<br />

30. Nutt JG. Effects of catechol-O-methyltransferase (COMT) inhibition on the pharmacokinetics of L-<br />

DOPA. Advances in Neurobiology. Vol. 69. Philadelphia: Lippincott-Raven Publishers, 1996:493–496.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_362.html (1 of 2) [4/5/2004 5:30:53 PM]


Document<br />

31. Törnwall M, Männistö PT. Acute toxicity of three new selective COMT inhibitors in mice with<br />

special emphasis on interaction with drugs increasing catecholaminergic neurotransmission.<br />

Pharmacology and Toxicology 1991; 69:64–70.<br />

32. Törnwall M, Männistö PT. Effects of three types of catechol-O-methylation inhibitors on 1-3,4dihydroxyphenylalanine-induced<br />

circling behaviour in rats. European Journal of Pharmacology 1993;<br />

250:77–84.<br />

33. Da Prada M, Borgulya J, Napolitano A, Zürcher G. Improved theraphy of parkinson's disease with<br />

tolcapone, a central and peripheral COMT inhibitor with an S- adenosyl-L-methionine-sparing effect.<br />

Clinical Neuropharmacology 1994; 17:26– 37.<br />

34. Kaakkola S, Gordin A, Männistö PT. General properties and clinical possibilities of new selective<br />

inhibitors of catechol-O-methyltransferase. General Pharmacology 1994; 25:813–824.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_362.html (2 of 2) [4/5/2004 5:30:53 PM]


Document<br />

35. Davis TL, Roznoski M, Burns RS. Effects of tolcapone in parkinson's patients taking Ldihydroxyphenylalanine/carbidopa<br />

and selegiline. Movement Disorders 1995; 10:349–351.<br />

Page 363<br />

36. Deleu D, Sarre S, Ebinger G, Michotte Y. The effect of carbidopa and entacapone pretreatment of<br />

the L-dopa pharmacokinetics and metabolism in blood plasma and skeletal muscle in beagle dog: an in<br />

vivo microdialysis study. Journal of Pharmacology and Experimental Therapeutics 1995;<br />

273:1323–1331.<br />

37. Napolitano A, Zürcher G, Da Prada M. Effect of tolcapone, a novel catechol-O-methyltransferase<br />

inhibitor, on striatal metabolism of L-DOPA and dopamine in rats. European Journal of Pharmacology<br />

1995; 273:215–221.<br />

38. Dingemanse J, Jorga KM, Schmitt M, et al. Integrated pharmacokinetics and pharmacodynamics of<br />

the novel catechol-O-methyltransferase inhibitor tolcapone during first administration to humans.<br />

Clinical Pharmacology and Therapeutics 1995; 57:508–517.<br />

39. Männistö PT. Clinical potential of catechol-O-methyltransferase (COMT) inhibitors as adjuvants in<br />

Parkinson's disease. CNS <strong>Drug</strong>s 1994; 1:172–179.<br />

40. Männistö PT, Lang A, Rauhala P, Vasar E. Beneficial effects of co-administration of catechol-Omethyltransferase<br />

inhibitors and l-dihydroxyphenylalanine in rat models of depression. European<br />

Journal of Pharmacology 1995; 274:229–233.<br />

41. Perez RA, Fernandez-Alvarez E, Nieto O, Piedrafita FJ. Kinetics of the reversible tight-binding<br />

inhibition of pig liver catechol-O-methyltransferase <strong>by</strong> [2-(3,4-dihydroxy-2-nitrophenyl)vinyl]phenyl<br />

ketone. Journal of Enzyme Inhibition 1994; 8:123–131.<br />

42. Taskinen J, Vidgren J, Ovaska M, Bäckström R, Pippuri A, Nissinen E. QSAR and binding model<br />

for inhibition of rat liver catechol-O-methyltransferase <strong>by</strong> 1,5- Substituted-3,4-Dihydroxybenzenes.<br />

Quantitative <strong>Structure</strong> Activity Relationships 1989; 8:210–213.<br />

43. Lotta T, Taskinen J, Bäckström R, Nissinen E. PLS Modeling of structure-activity relationships of<br />

catechol-O-methyltransferase inhibitors. Journal of Computer- Aided Molecular <strong>Design</strong> 1992;<br />

6:253–272.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_363.html (1 of 2) [4/5/2004 5:31:00 PM]


Document<br />

44. Burley SK,<br />

Petsko GA.<br />

Aromaticaromatic<br />

interaction: A<br />

mechanism of<br />

protein structure<br />

stabilization.<br />

Science 1985;<br />

229:23–28.<br />

45. Serrano L, Bycroft M, Fersht AR. Aromatic-aromatic interactions and protein stability. Journal of<br />

Molecular Biology 1991; 218:465–475.<br />

46. Piedrafita FJ, Elorriaga C, Fernandez-Alvarez E, Nieto O. Inhibition of catechol- O-<br />

methyltransferase <strong>by</strong> N-(3,4-dihydroxyphenyl) maleimide. Journal of Enzyme Inhibition 1990; 4:43–50.<br />

47. Wikberg T. Docotoral Thesis, University of Helsinki, Helsinki, Finland, 1993:29– 30.<br />

48. Aperia A. Dopamine action and metabolism in the kidney. Current Opinion in Nephrology and<br />

Hypertension 1994; 3:39–45.<br />

49. Fernandes MH, Soares-da-Silva P. Role of monoamine oxidase and catechol-O- methyltransferase in<br />

the metabolism of renal dopamine. Journal of Neural Transmission. Supplementum 1994; 41:101–105.<br />

50. Kraulis PJ. MOLSCRIPT: a program to produce both detailed and schematic plots of protein<br />

structures. Journal of Applied Crystallography 1991; 23:946–950.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_363.html (2 of 2) [4/5/2004 5:31:00 PM]


http://legacy.netlibrary.com/reader/message.asp?message=811&BookID=12640&FileName=Page_364.html<br />

The requested page could not be found.<br />

Return to previous page<br />

http://legacy.netlibrary.com/reader/message.asp?message=811&BookID=12640&FileName=Page_364.html [4/5/2004 5:31:10 PM]


Document<br />

Page 365<br />

15<br />

Antitrypanosomiasis <strong>Drug</strong> Development Based on <strong>Structure</strong>s of Glycolytic<br />

Enzymes<br />

Christophe L. M. J. Verlinde, Hidong Kim, Bradley E. Bernstein, Shekhar C. Mande, * and Wim G.J.<br />

Hol<br />

University of Washington, Seattle Washington<br />

I. Trypanosomiasis<br />

A. Disease and Treatment<br />

Human African trypanosomiasis, also called sleeping sickness, is caused <strong>by</strong> the parasites Trypanosoma<br />

brucei gambiense and Trypanosoma brucei rhodesiense. These unicellular organisms live freely in the<br />

bloodstream of the human host and invade the brain during the later stage of the disease. Without<br />

treatment the disease is always fatal [1]. The course of the gambiense form may last from months to<br />

years, while T. brucei rhodesiense usually kills within weeks. Sleeping sickness occurs in thirty-six sub-<br />

Saharan African countries, putting fifty million people at risk. Each year 25,000 new cases are reported,<br />

but the actual number of cases is more likely to be about 250,000 [2]. The current state of<br />

antitrypanosomal chemotherapy is dismal; many parasitologists do not want to take the risk of being<br />

infected with T. brucei rhodesiense and prefer to study the parasite T. brucei brucei, which is harmless<br />

to humans [3].<br />

Not more than four drugs are available to treat the disease: pentamidine, suramin, melarsoprol—all from<br />

the first half of this century—and eflornithine, introduced in 1990 (Figure 1). Except for eflornithine,<br />

which is an irreversible ornithine decarboxylase inhibitor [4], the mechanisms of the drugs are poorly<br />

understood [5]. All four drugs require administration <strong>by</strong> injection in a hospital setting, which is a major<br />

drawback in rural Africa [6]. Pentamidine is useful for treating early stage T. brucei gambiense<br />

infection, suramin for both early stage gambiense and rhodesiense infection. The permanent charges on<br />

pentamidine<br />

* Current affiliation: Institute of Microbial Technology, Chandigarh, India<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_365.html [4/5/2004 5:31:12 PM]


Document<br />

Figure 1<br />

Available drugs for the treatment of human African trypanosomiasis.<br />

Page 366<br />

and suramin explain why they show poor oral absorption and do not cross the blood–brain barrier,<br />

making them unsuitable for the treatment of late-stage trypanosomiasis. Melarsoprol, an organoarsenical<br />

compound, was until 1990 the only drug effective in the late stage of both forms of trypanosomiasis.<br />

Unfortunately, it is also highly toxic, causing reactive encephalopathy in up to 10% of the patients, of<br />

which about one half die. This deadly complication is well known to villagers of areas where the disease<br />

is endemic and, ironically, discourages people from participating in diagnostic surveys. Eflornithine,<br />

heralded as the “resurrection drug” upon its introduction [7], cures patients infected with late-stage T.<br />

brucei gambiense but is ineffective against the more virulent rhodesiense form. Additionally, it causes<br />

bone marrow suppression in half of the patients and occasionally convulsions [6]. A serious concern is<br />

that resistance has been reported against each of the four antitrypanosomal drugs [1].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_366.html [4/5/2004 5:31:28 PM]


Document<br />

Page 367<br />

Sleeping sickness has been largely ignored <strong>by</strong> the pharmaceutical industry because the poor<br />

socioeconomic situation in the part of the African continent afflicted <strong>by</strong> this debilitating disease offers<br />

little prospect of reasonable financial returns [8]. It is revealing that eflornithine was originally<br />

developed as an anticancer drug. It was screened for antitrypanosomal properties only when the<br />

biochemistry of the trypanosomal polyamine metabolism was understood [9]. Fortunately, the cell<br />

biology of trypanosomes is so extraordinary that they have been the subject of more fundamental<br />

research than most other protozoan parasites [5]. Each of these unique features of T. brucei may become<br />

a target for new drugs, provided they prove to be essential for the survival of the parasite in the human<br />

host. With such a wealth of biochemical information, structure-<strong>based</strong> drug design provides a tremendous<br />

opportunity to arrive at new drugs to cure sleeping sickness.<br />

B. Targets for Future <strong>Drug</strong>s<br />

Preventing trypanosomiasis would be a nobler goal than curing it. Unfortunately, trypanosomes are<br />

experts in evading our immune system. They achieve this <strong>by</strong> varying their dense surface coat. It is<br />

composed of ten million copies of a single protein, the variant surface glycoprotein (VSG), for which<br />

they have no less than a thousand different genes. In this way trypanosomes can change surface antigens<br />

more rapidly than the host can produce new antibodies [10]. Clearly, such a mechanism leaves little<br />

hope for preventing sleeping sickness <strong>by</strong> vaccination.<br />

In contrast to the small number of drugs available to treat trypanosomiasis, the opportunities for<br />

developing new drugs are ample, as can be seen from Table 1. They range from unique RNA processing<br />

to reduced metabolism, salvage systems, and different rates of protein turnover. All of these features<br />

were the subject of an outstanding review <strong>by</strong> C.C. Wang [5].<br />

An inventory of the structural information waiting to be exploited <strong>by</strong> structure-<strong>based</strong> drug design reveals<br />

eight potential target enzymes (Table 2). Since for most trypanosomatid proteins there is a human<br />

counterpart it is mandatory that designed inhibitors be selective, i.e., have very little affinity for the<br />

equivalent enzymes of the human host. As a consequence, selective design requires pairs of equivalent<br />

structures from the parasite and from the host. A complication for selective design is that for three of the<br />

mammalian enzymes only the structure of one isoenzyme is known. For example, the structure of human<br />

aldolase A has been determined [23] but not the structures of isoenzymes B and C, which share only 69<br />

and 82% sequence identity to isoenzyme A [40–42]. Of course, homology modeling might be a way to<br />

overcome this problem.<br />

From Table 2 it is evident that only for three proteins, TIM, GAPDH, and trypanothione reductase the<br />

structures of the parasite and host enzymes are<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_367.html [4/5/2004 5:31:49 PM]


Document<br />

Table 1 Trypanosomal Targets, Their Human Equivalents, and Current Leads for <strong>Drug</strong> <strong>Design</strong><br />

T. brucei Human host Lead<br />

RNA editing <strong>by</strong> trans-splicing [11] cis-splicing none<br />

All energy from fast glycolysis [12] glycolysis and oxidative<br />

phosphorylation<br />

Glucose transporter [14] human erythrocyte glucose<br />

transporter<br />

Purine P2 transporter [15] none<br />

Purine salvage enzymes, e.g., HGPRT b [16] HGPRT none<br />

Slow rate of enzyme turnover [17], e.g., ornithine<br />

decarboxylase [18]<br />

Polyamine metabolism, e.g., S-adenosyl<br />

methionine decarboxylase [19]<br />

MMBA a [13]<br />

none<br />

Page 368<br />

fast rate of enzyme turnover eflornithine (= drug) [9]<br />

S-adenosyl methionine<br />

decarboxylase<br />

MDL 73811 c<br />

Trypanothione reductase [20] glutathione reductase mepacrine<br />

VSG anchor: a myristate-containing GPI [21] 10-(propoxy)-decanoate d<br />

a MMBA = 2' -deoxy-2' -(m-methoxybenzamido)-adenosine.<br />

b HGPRT = hypoxanthine guanine phosphoribosyltransferase.<br />

c MDL 73811 = 5' -{[(Z)-4-amino-2-butenyl]methylamino}-5' -deoxyadenosine.<br />

dIt has not been established whether the trypanocidal effect of this compound is due to its incorporation in the GPI anchor<br />

[5].<br />

known. For five more targets the crystal structure of only the mammalian enzyme is available. Efforts to<br />

solve the structures of trypanosmatid aldolase, PGK, and PK counterparts are underway in our lab. This<br />

review explains how we attempted to arrive at selective inhibitors of three trypanosomal glycolytic<br />

enzymes.<br />

C. Trypanosomal Glycolysis: Enzyme Inhibition as a Target<br />

In the bloodstream of the human host, trypanosomes are metabolically<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_368.html (1 of 2) [4/5/2004 5:32:08 PM]


Document<br />

“lazy”. Since there is plenty of glucose and oxygen available, they rely solely on glycolysis to the stage of<br />

pyruvate for their energy supply [12]. Their glycolysis proceeds at an amazing rate, which is about fifty<br />

times faster than in the cells of the human host [43]. This fast rate is a necessity because only two<br />

molecules of ATP are generated per molecule of glucose instead of the thirty-six produced <strong>by</strong> complete<br />

oxidation. These findings led to the proposal that inhibitors of trypanosomal glycolysis might be turned<br />

into drugs. Support for this idea comes from in vitro experiments where salicylhydroxamic acid (SHAM)<br />

was used to bring the parasite under anaerobic conditions, after which glycolysis was<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_368.html (2 of 2) [4/5/2004 5:32:08 PM]


Document<br />

Table 2 Three-Dimensional <strong>Structure</strong>s Available for Trypanosomal <strong>Drug</strong> <strong>Design</strong><br />

(a) Glycolytic:<br />

(b) Other:<br />

Trypanosomatid PDB code Mammalian PDB code<br />

PGIa — — Porcine 1PGI[22]<br />

Aldolase — — Human b 1ALD[23]<br />

TIM T. brucei 5TIM[24] Human 1HTI[25]<br />

GAPDH T. brucei<br />

L.mexicana<br />

1GGA[26,27]<br />

[30]<br />

PGK — — Horse<br />

Pig<br />

Human c 3GPD[28,29]<br />

2PGK[31]—[32]<br />

PK — — Cat d 1PYK[33]<br />

TR T. cruzi 1NDA[34] HumanGR 3GRS[35]<br />

C. fasciculata 1TYT[36]<br />

1PPR[37]<br />

HGPRT — — Human 1HMP[38]<br />

VSG T. brucei 1VSG[39] No equivalent<br />

a Abbreviations: PGI = phosphoglucose isomerase, PGK = phosphoglycerate kinase, PK = pyruvate kinase, TR =<br />

trypanothione reductase; GR = glutathione reductase, HGPRT = hypoxanthine-guanine phosphoribosyl<br />

transferase, VSG = variable surface glycoprotein.<br />

b A isoenzyme, from muscle; other mammalian isoenzymes are B in liver and C in brain.<br />

c Muscle isoenzyme; a liver isoenzyme exists.<br />

dM1 isoenzyme, from muscle; mammals also have M2 in kidney, adipose tissue and lung, L in liver, and R in<br />

rood blood cells.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_369.html (1 of 2) [4/5/2004 5:32:27 PM]<br />

Page 369


Document<br />

blocked <strong>by</strong> mass action through the addition of glycerol (Figure 2). As a result trypanosomes were lysed<br />

within five minutes [44,45]. Treatment of infected rodents with the SHAM/glycerol mixture proved to<br />

be effective to clear the blood of the animals from T. brucei [46], although only with sublethal doses was<br />

permanent aparasitemia obtained [47]. If glycolysis could be blocked selectively, i.e., without affecting<br />

the equivalent enzymes of the host, one might have a promising therapy against trypanosomiasis.<br />

D. Beyond Enzyme Inhibition: Protein Routing as a Target<br />

In trypanosomes, seven enzymes involved in glycolysis, from hexokinase to phosphoglycerate kinase,<br />

are sequestered in specialized organelles, called glycosomes [48]. These microbodies are probably<br />

evolutionary relics of an endosymbiont [49] but are devoid of genetic material encoding for the<br />

glycosomal enzymes. Instead, these enzymes are encoded in the nucleus and are post-translationally<br />

imported into the glycosome. Since the import process likely involves unfolding one might envision<br />

blocking import <strong>by</strong> stabilizing the folded<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_369.html (2 of 2) [4/5/2004 5:32:27 PM]


Document<br />

Figure 2<br />

Glycolysis in bloodstream-form trypanosomes. All net energy production takes place<br />

in the cytosol as the result of ATP formation <strong>by</strong> pyruvate kinase. However, the<br />

majority of glycolytic enzymes are sequestered into a specialized organelle, the<br />

glycosome. There the net ATP synthesis is zero irrespective of the presence of oxygen.<br />

Under aerobic conditions the NADH produced <strong>by</strong> glyceraldehyde-3-phosphate dehydrogenase<br />

is reoxidized via the G-3-P/DHAP shuttle, which couples glycolysis to a<br />

mitochondrial glycerophosphate oxidase. Under anaerobic conditions or when the<br />

oxidase is blocked <strong>by</strong> SHAM, equimolar amounts of glycerol and 3-phosphoglycerate<br />

are formed. However, the addition of an excess of glycerol to the cytosol prevents the<br />

reoxidation of NADH. As a result trypanosomes treated with a mixture of SHAM and<br />

glycerol die in a matter of minutes.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_370.html (1 of 2) [4/5/2004 5:33:24 PM]<br />

Page 370


Document<br />

state through the tight binding of ligands. In this way inhibitors might act at two levels, directly <strong>by</strong><br />

blocking catalysis and indirectly <strong>by</strong> preventing proper enzyme routing in the parasite. Thus far, we have<br />

engaged in a collaborative effort to inhibit three of the glycosomal enzymes. (From Ref. 95. Copyright<br />

1988 <strong>by</strong> Elsevier.)<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_370.html (2 of 2) [4/5/2004 5:33:24 PM]


Document<br />

II. Three Glycolytic Enzymes Of T. Brucei: Molecular Biology, Biochemistry, and X-Ray<br />

Crystallography<br />

A. Triosephosphate Isomerase (TIM)<br />

Page 371<br />

TIM is a homodimeric enzyme that interconverts dihydroxyacetone phosphate and glyceraldehyde-3phosphate.<br />

It ensures that both trioses derived from glucose can be used for ATP production in the<br />

glycolytic pathway. Triosephosphate isomerase does not require any cofactor. Both the T. brucei and<br />

human enzymes have been overexpressed in Escherichia coli [50,25] and their crystal structures were<br />

solved in our group [24,25]. In addition, the structures of T. brucei TIM in complex with seven<br />

nonselective competitive inhibitors, with inhibition constants of 300 μM or higher were determined:<br />

monohydrogen phosphate [51], 2-phosphoglycerate [52], 3-phosphoglycerate [53], 3phosphonopropionate<br />

[53], glycerol-3-phosphate [53], 2-(N-formyl-N-hydroxyamino)-ethyl phosphonic<br />

acid [54], and N-hydroxy-4-phosphonobutanamide [55]. These studies gave an excellent picture of<br />

different ligand binding modes and of the conformational flexibility of the enzyme.<br />

All ligands interact with the main features of the catalytic machinery of the enzyme (Figure 3): (1) the<br />

phosphate is sequestered <strong>by</strong> the positive end of a 3 10-helix and Lys13; (2) polar groups on the carbon<br />

framework interact with His95 and Glu167, the catalytic electrophile and the catalytic base of the<br />

enzyme, respectively; (3) the entire inhibitor is shielded from the bulk solvent <strong>by</strong> a flexible loop, which<br />

normally closes over the substrate during catalysis to prevent phosphate elimination [56] (Figure 4). The<br />

only exception to flexible loop closure is N-hydroxy-4-phosphono-butanamide. It binds to the enzyme<br />

with the flexible loop in the open conformation because its size precludes loop closure. Thus, the<br />

crystallographic binding studies point out that it should be possible to design two very different classes<br />

of selective inhibitors: a class that binds to the enzyme in the closed loop conformation and one that<br />

binds to the open loop conformation.<br />

Selective inhibitor design in the case of TIM appears to be a formidable task. All residues within 10 Å<br />

of the active site are conserved [25]. This is also reflected in the similarity of the kinetic characteristics<br />

between trypanosomal and human TIM: for T. brucei TIM, K m (glyceraldehyde-3-phosphate) = 0.25<br />

mM, kcat = 3.7 × 10 5min -1 [57]; for human TIM, K m = 0.49 mM, kcat = 2.7 × 10 5 min -1 [25]. There are<br />

significant differences in the surface protein of the two enzymes about 15 Å away from the substrate<br />

phosphorus atom [58]. In a shallow cleft, T. bruceiTIM has Ala100-Tyr101, while the human<br />

counterpart of these residues is His-Val (Figure 5). The cleft is formed <strong>by</strong> the flexible loop of one<br />

subunit of the enzyme and a different loop originating from another subunit. When the flexible loop<br />

changes its conformation from the closed to the open form the cleft widens substantially. Moreover, the<br />

Ala-Tyr dipeptide becomes then directly accessible from the active site, the distance being about<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_371.html [4/5/2004 5:34:04 PM]


Document<br />

Figure 3<br />

Schematic representation of the structure of a TIM monomer. Helices and strands<br />

are labeled as H and B, respectively. The view is along the axis of the β barrel, into<br />

the active site. Key catalytic residues Lys13, His95, and Glu167 are shown along with<br />

the helix that binds the substrate phosphate and the flexible loop that covers the<br />

substrate during catalysis. Black dots indicate residues in contact with the second<br />

monomer of the enzyme. (From Ref.24. Copyright 1991 <strong>by</strong> Harcourt Brace.)<br />

Page 372<br />

10 Å instead of the 15 Å in the closed loop conformation of the enzyme. In any case, selective inhibitor<br />

design for TIM appears to require de novo design as there are no leads known that interact with the Ala-<br />

Tyr region of the enzyme.<br />

B. Glyceraldehyde-3-Phosphate Dehydrogenase (GAPDH)<br />

Glyceraldehyde-3-phosphate dehydrogenase is a homotetramer that carries out the oxidative<br />

phosphorylation of glyceraldehyde-3-phosphate into 1,3-bisphos- phoglycerate. During this reaction<br />

NADH is formed. Each subunit of the enzyme consists of two domains and has an NAD + binding site.<br />

The N-terminal domain anchors the adenosine portion of the cofactor while the nicotinamide portion is<br />

involved in the catalytic reaction at the C-terminal domain. T. brucei<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_372.html (1 of 2) [4/5/2004 5:34:37 PM]


Document<br />

Figure 4<br />

Stereoview of superimposed TIM monomers, one with an open flexible loop (full black)<br />

and another one with a closed flexible loop (open gray). The loop location is marked<br />

<strong>by</strong> an asterisk. Note the proximity of the active site, indicated <strong>by</strong> the catalytic residues<br />

Lys13, His95, and Glu167 (all sterofigures in this paper were drawn with<br />

MOLSCRIPT [93]).<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_373.html (1 of 2) [4/5/2004 5:36:17 PM]<br />

Page 373


Document<br />

Figure 5<br />

Exploitable structural differences between T.brucei(full) and human (dashed)<br />

TIM. The inhibitor 2-phosphoglycolate as observed in the structure of the human<br />

enzyme indicates the location of the active site. <strong>Drug</strong> design targets are the T.brucei<br />

Ala100-Tyr101, which are considerably different from their human counterpart<br />

His-Val. (From Ref.25. Copyright 1994 <strong>by</strong> Cambridge University Press.)<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_373.html (2 of 2) [4/5/2004 5:36:17 PM]


Document<br />

Figure 6<br />

Stereoview of NAD binding <strong>by</strong> trypanosomal GAPDH (full black). For clarity<br />

only 2 of the 4 subunits are shown. A substantial deviation of the protein backbone<br />

occurs in human GAPDH (open gray, only one subunit shown) near the adenosine<br />

part of the cofactor. The near<strong>by</strong> cleft (marked <strong>by</strong> an asterisk) is important for<br />

introducing selectivity in inhibitor binding and has therefore been termed “selectivity cleft.”<br />

Page 374<br />

glycosomal GAPDH has been overexpressed in E. coli [59] while human erythrocyte GAPDH is<br />

available from commercial suppliers. The sequences of the two enzymes are only 55% identical [60].<br />

The crystal structure of the parasite enzyme was solved from Laue data at 3.2 Å resolution in our group<br />

(Figure 6) and in a second crystal form at 2.8 Å [26]. The structure of the human muscle enzyme was<br />

solved at 3.5 Å resolution in the group of the late Herman Watson [28]. Its resolution was improved to<br />

2.3 Å in our group [26,29]. Both structures were of the holo-enzyme, i.e., the enzyme in presence of the<br />

cofactor.<br />

The active site and the nicotinamide-binding site of the two enzymes are very well conserved. This is<br />

reflected in similar K m (glyceraldehyde-3-phos-phate) values of 0.15 and 0.17 mM for the trypanosomal<br />

and human enzyme, respectively [61]. Surprisingly, the K m (NAD +) values differ <strong>by</strong> a factor ten: 0.45<br />

mM for T. brucei and 0.04 mM for human GAPDH [61]. In view of the conservation of the<br />

nicotinamide and pyrophosphate binding sites, the substantial difference in K m (NAD +) has to be<br />

ascribed to the adenosine binding environment. Indeed, some of the residues embracing the adenine ring<br />

of the cofactor are not identical. In the trypanosomal enzyme the adenine is sandwiched in between a<br />

Thr in the back and a Met in the front. This Met is replaced <strong>by</strong> Pro and Phe in the human enzyme<br />

(Figure 7). A second difference involves the residue flanking the C2 atom of the purine ring. It is a Val<br />

in the trypanosomal enzyme, and Asn in the human enzyme. These differences apparently account for a<br />

ten-fold lower affinity for the cofactor.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_374.html (1 of 2) [4/5/2004 5:38:25 PM]


Document<br />

Figure 7<br />

Comparison of the binding modes of the adenosine moiety of NAD to GAPDH:<br />

(left) in T.brucei,(right) in the human enzyme. Note the identical hydrogen bonds to<br />

the purine N6 and the ribosyl hydroxyls. The purine ring is embraced <strong>by</strong><br />

hydrophobic residues that are not conserved. Also, a unique cleft near O2', which we<br />

called the selectivity cleft, is present in the T.brucei enzyme. (From Ref. 13.Copyright<br />

1994 <strong>by</strong> the American Chemical Society.)<br />

Page 375<br />

Other differences in the vicinity of the adenosine portion of the NAD + cofactor are prime targets for<br />

selective inhibitor design. Close to C8 of the adenine ring, the trypanosomal GAPDH exhibits a Leu<br />

while its human counterpart has a smaller residue, namely Val (Figure 7). Also, the parasite enzyme<br />

possesses a hydrophobic cleft near the 2' -hydroxyl of the adenosine ribose. This cleft, termed the<br />

“selectivity cleft” is almost absent in the human enzyme due to a different local backbone conformation<br />

and the presence of the Ile37 side chain. In conclusion, the adenosine binding region looks like an<br />

excellent target for selective inhibitor design.<br />

It is exciting that the residues responsible for binding adenosine in T. brucei GAPDH are identical to<br />

their counterparts in glycosomal GAPDH of Leishmania mexicana, another trypanosomatid [31]. L.<br />

mexicana is one of the most common species of Leishmania throughout Central and South America and<br />

the southern United States. In humans it hides as amastigotes in the macrophages and causes hideous<br />

skin lesions. Together with about twenty other species of Leishmania these parasites infect about twelve<br />

million people annually [63]. Though they are less dependent on glycolysis than T. brucei there is<br />

evidence that stibogluconate, a well-known drug for treating leishmaniasis, specifically inhibits<br />

glycolysis in these parasites [64]. Therefore, we decided to study GAPDH of L. mexicana in parallel<br />

with the T. brucei enzyme. Its structure<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_375.html (1 of 2) [4/5/2004 5:39:00 PM]


Document<br />

Page 376<br />

was recently solved at 2.8 Å resolution in our lab [30]. The kinetic parameters of the two enzymes are<br />

virtually identical: K m (glyceraldehyde-3-phosphate) = 0.13 mM and K m (NAD +) = 0.41 mM for the L.<br />

mexicana enzyme [62]. As expected, the structures of the two parasite enzymes are very similar: the rms<br />

deviation for all backbone atoms is 0.7 Å and the adenosine binding environment is perfectly<br />

superimposable except for Asn39 of T. brucei GAPDH, which is a Ser in the L. mexicana enzyme. In<br />

this way drug design for one disease may have implications for another one.<br />

C. Phosphoglycerate Kinase (PGK)<br />

A monomeric enzyme, PGK transfers the acylated phosphoryl group from 1,3- bisphosphoglycerate to<br />

ADP, thus forming 3-phosphoglycerate and ATP. The enzyme uses a metal ion as a cofactor, namely<br />

Mg 2+ [65]. The PGK enzyme from T. brucei has been overexpressed in E. coli [66]. Human PGK is not<br />

Figure 8<br />

Steroview of B.stearothermophilus PGK [69] in complex with ADP bound to the<br />

C-terminal domain. To illustrate the sugar binding site, the 3-phosphoglycerate has<br />

been added to the figure <strong>based</strong> on the crystal structure of pig muscle PGK [32].From<br />

the distance between the two substrates it is obvious that during catalysis a<br />

hinge-bending motion between the domains of the protein has to occur to bring the<br />

substrates together.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_376.html [4/5/2004 5:39:23 PM]


Document<br />

Table 3 Residues Involved in ADP Binding in Glycosomal and Human PGK<br />

T.brucei Human_1 a Human_2 b In contact with ADP moiety<br />

Ala 242 Gly 238 Gly 238 Adenine<br />

Tyr 245 Phe 241 Tyr 241 Adenine<br />

Lys 259 Leu 256 Leu 256 Adenine<br />

Ala 314 Gly 309 Gly 309 Adenine<br />

Ser 378 Thr 375 Thr 375 β-phosphate<br />

a Somatic PGK [67].<br />

b PGK in spermatogenic cells [68].<br />

Page 377<br />

commercially available. However, its sequence has been determined [67] and appears to be 97%<br />

identical to horse and pig PGK. For completeness it should be mentioned that there is a second human<br />

PGK in testis tissue that is 87% identical to the somatic enzyme [68].<br />

The crystal structures of the apo-enzyme from horse [31] and of the binary complex between pig PGK<br />

and its substrate [32] (Figure 8) are available from the Protein Databank. The substrate was found to<br />

bind to the N-terminal domain of the enzyme. The binding site for ADP is known from the structure of<br />

its binary complex with PGK from B. stearothermophilus [69]. It resides in the Cterminal domain. Since<br />

the substrate and ADP binding sites are 10 Å apart, a hinge-bending motion between the two domains<br />

has been postulated to occur during catalysis [70].<br />

Kinetically glycosomal PGK from T. brucei and mammalian PGK are very similar: the K m values for<br />

ATP are 0.29 and 0.46 and mM, respectively; the K m values for 3-phosphoglycerate are 1.62 and 0.62<br />

mM, respectively (due to the unavailability of the human enzyme the rabbit muscle enzyme was used as<br />

a substitute) [71]. The residues responsible for binding the substrate [32] are identical between human<br />

and glycosomal PGK [72]. However, five of the residues involved in the binding of ADP differ between<br />

the two enzymes (Table 3). Apparently, the biggest difference between human PGK and the parasite<br />

enzyme is the charged residue Lys259, which has the apolar Leu256 as a human counterpart. Molecules<br />

that bind at the ADP binding site and specifically recognize Lys259 might therefore be good starting<br />

points for drug design.<br />

III. Search For New Leads<br />

A. Triosephosphate Isomerase: The Crystallographic Cocktail Soak Approach<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_377.html (1 of 2) [4/5/2004 5:39:43 PM]


Document<br />

Because there are no known leads that bind to the selectivity region of TIM, the design of selective<br />

inhibitors is an exercise in de novo ligand design. We tried to<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_377.html (2 of 2) [4/5/2004 5:39:43 PM]


Document<br />

Page 378<br />

design such molecules on the basis of the trypanosomal TIM structure <strong>by</strong> a linked-fragment approach<br />

[57]. In that strategy small building blocks are designed to be complementary to the targeted surface of a<br />

protein. Such fragments can then be synthesized or purchased, tested for their effect on enzyme kinetics<br />

and for their binding mode <strong>by</strong> crystallography. Promising fragments are then linked together into larger<br />

molecules. The idea behind this stepwise approach is to obtain early experimental feedback in the drugdesign<br />

cycle.<br />

The crystallographic follow-up of our linked-fragment approach design for trypanosomal TIM was<br />

disappointing. Two designed fragments were soaked into a crystal of trypanosomal TIM, namely 4hydroxy-2-butanone<br />

and D-asparagine. Despite high concentrations of these molecules in the mother<br />

liquor, 220mM and 30 mM respectively, no convincing electron density could be seen in difference<br />

Fourier maps calculated between 10.0-2.8 Å [72]. Common to both molecules is that they are fairly<br />

polar, rather flexible, and were expected to displace crystallographically observed water molecules.<br />

Apparently, de novo design of tightly binding small ligands is far from trivial.<br />

We also tried to find new leads <strong>by</strong> a completely experimental approach. For that purpose the<br />

crystallographic cocktail soak (CCS) approach was developed. In this method cocktails of fine<br />

chemicals are soaked into a crystal in the hope of finding crystallographic evidence of binding for one of<br />

the molecules from the cocktail. The identification of such a molecule might not be clear immediately<br />

because several molecules in the cocktail might be compatible<br />

with the shape of the electron density, especially if the resolution is not very high. An outcome would be<br />

provided <strong>by</strong> a dichotomic approach (Figure 9), in which the crystallographic soaking experiment is<br />

repeated with ever smaller subcocktails of the original one. For example, if a ligand shows up from a<br />

cocktail soak of 32 compounds, a second experiment should be done with only half of the compounds. If<br />

the ligand fails to show up, one knows that it is one of the alternative 16 compounds. After at most six<br />

experiments the identity of the ligand is known. One might like to think of the CCS approach as the<br />

experimental analog of the computational methods in programs like GRID [73] or MCSS [74], but with<br />

thirty-two compounds at a time.<br />

Figure 9<br />

Dichotomic search for unknown ligand from<br />

cocktail 1. The number of compounds in the sub<br />

cocktail is indicated <strong>by</strong> n, and the<br />

interpretation about the presence of compound<br />

X in the subcocktail is given as Y/?/N.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_378.html (1 of 2) [4/5/2004 5:39:46 PM]


Document<br />

Page 379<br />

For trypanosomal TIM we experimented with three different cocktails of 32 compounds (Table 4).<br />

Molecules were chosen in such a way that they would be compatible, soluble, cheap, and as varied as<br />

possible. Each compound was present at a concentration of 1 mM. The final cocktail solutions were<br />

clear and devoid of precipitate. Since this was a pilot experiment both subcocktails were checked at each<br />

stage of the dichotomic strategy. Only the soak with cocktail 1 revealed electron density that could not<br />

be accounted for <strong>by</strong> water molecules, hereafter called peak X. The soaks with cocktails 2 and 3 led to<br />

featureless difference Fourier maps. The quality of the data and refinement can be inspected from Table<br />

5, while Figure 9 illustrates the dichotomic search to identify peak X. An oxidized molecule of DTT,<br />

identified in the high-resolution structure of the native TIM crystals [24], served as an internal reference<br />

to judge the quality of the data and the noise level in the final difference Fourier maps.<br />

Peak X was found near His95 of the second subunit of the enzyme, i.e., the subunit where the flexible<br />

loop adopts the closed conformation in this crystal form. Its signal was somewhat weaker than that of<br />

DTT. The same density showed up when crystals were soaked with subcocktail 1B but not with 1A,<br />

narrowing down the list of potential ligands to sixteen compounds. However, the next round of the<br />

dichotomic search led to a problem that has not been solved thus far. Peaks of roughly the same shape as<br />

the original peak X appeared with both subcocktails 1BA and 1BB. Several strategies were followed to<br />

improved the quality of the maps. First, the model was further refined with all data while a bulk solvent<br />

scattering correction [75] was incorporated. Second, a variety of maps were calculated: (|F o|-|F c|) e iαc,<br />

(2|Fo|-|Fc|) eiαc, (3|Fo|-2|Fc|) eiαc and (|Fo|-|Fo,native|)eiαc. Third, all maps were SIGMAA-weighted [76].<br />

Since the shape of peak X varied substantially between the different maps it can be tentatively<br />

concluded that peak X did not originate from the presence of a compound but was noise. The lesson of<br />

this experiment seems to be that the crystallographic cocktail soaking approach should only be tried<br />

when high- resolution data can be obtained, probably better than 1.8 Å resolution.<br />

B. Glyceraldehyde-3-Phosphate Dehydrogenase: Docking<br />

In order to discover new ligands that would block GAPDH of T. brucei <strong>by</strong> occupying the adenosine<br />

binding region we used the program DOCK [77], version 3.5. This program characterizes a binding site<br />

<strong>by</strong> filling it with a set of overlapping spheres. The centers of these generated spheres constitute an<br />

irregular grid, called a “graph” <strong>by</strong> mathematicians. Docking of a ligand then consists of matching<br />

subsets of ligand interatomic distances onto subsets of the receptor graph. Finally, the quality of the fit<br />

between a docked ligand and the receptor is evaluated. Within DOCK 3.5 three methods are available<br />

for this evaluation: contact scoring, which measures shape complementarity; force-field scoring, which<br />

is an estimate of the enthalpy of the intermolecular interaction; and elec-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_379.html [4/5/2004 5:39:49 PM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_380.html (1 of 2) [4/5/2004 5:40:39 PM]<br />

Page 380


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_380.html (2 of 2) [4/5/2004 5:40:39 PM]


Document<br />

Table 5 Crystallographic Cocktail Soaking Experiments of Trypanosomal TIM Crystals a<br />

res DTT X<br />

C a(Å) b(Å) c(Å) m(°) (Å) R-sym Compl R (σ) (σ)<br />

1 112.7(1) 97.7(1) 46.7(1) 0.7 2.85 0.044 0.97 0.136 5.0 3.5<br />

1A 112.6(1) 97.6(1) 46.8(1) 0.9 2.50 0.099 0.96 0.173 7.0 2.0<br />

1B 112.7(2) 97.5(3) 46.8(2) 0.7 2.30 0.044 0.88 0.172 5.0 4.0<br />

1BA 112.6(2) 97.9(2) 46.7(2) 0.7 2.40 0.051 0.96 0.202 7.8 3.0<br />

1BB 112.6(1) 97.8(1) 46.7(1) 0.7 2.30 0.034 0.90 0.203 5.0 3.0<br />

2 112.9(1) 97.8(1) 46.7(1) 0.9 2.40 0.060 0.92 0.172 5.0 -<br />

3 112.9(1) 97.7(3) 46.7(2) 0.7 2.40 0.057 0.93 0.170 4.0 -<br />

a The following data are tabulated column-wise: C = cocktail; a, b, c = cell parameters of the P212 12 1 crystals; m =<br />

mosaicity; res = resolution; R-sym = agreement between symmetry-related reflections; Compl = completeness; R =<br />

agreement between data and model; DTT = signal of oxidized DTT in the final difference Fourier map. DTT is<br />

present in the mother liquor but not incorporated in the model. X = signal of peak X in the final difference Fourier<br />

map. Maps were calculated with data between 10.0 Å and the high resolution limit.<br />

trostatic scoring, where the linearized Poisson-Boltzmann equation is solved [78]. We report here on<br />

docking experiments to identify GAPDH inhibitors from the Available Chemicals Directory-3D 93<br />

(ACD) [80].<br />

Force-field scoring and electrostatic scoring require the assignment of partial atomic charges.<br />

Unfortunately, such charges are not available in the ACD, mainly because there is no consensus on a<br />

method to calculate them. Since the number of molecules in the ACD is very large, about 73,000 in<br />

1993, we opted for the charge-equilibration algorithm developed <strong>by</strong> Rappé and Goddard [81] as<br />

implemented in the BIOGRAF program [82]. This method leads to charges that are in excellent<br />

agreement with experimental dipole moments and with atomic charges obtained from electrostatic<br />

potentials of accurate ab initio calculations. A script was written that processed the entire ACD<br />

automatically on an R4000 processor in about two days. Charges of the protein atoms were assigned<br />

from the table of AMBER-derived charges provided with DOCK 3.5.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_381.html (1 of 2) [4/5/2004 5:40:43 PM]<br />

Page 381


Document<br />

Because there are tens of parameters that can be varied in the program there is no such a thing as the<br />

DOCK run for a given protein target. We chose to perform two parallel runs that differ in receptor<br />

description. In run 1 the DOCK sphere center description was used while in run 2 the atomic coordinates<br />

of 2'- deoxy-2'-(3-methoxybenzamido)adenosine, a designed selective inhibitor of T. brucei GAPDH<br />

(see Section IV), were picked to describe the receptor site. For each run the same program parameters<br />

were chosen (Table 6) and the ACD database was split up in batches of 10,000 molecules. The<br />

computation was done on an Indigo2 workstation with an R4400 processor operating at 175 MHz. It<br />

took 3 h 19 min of CPU time for run 1 and 34 h 48 min for run 2. The ten-fold time difference between<br />

the two runs originates from the different number of<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_381.html (2 of 2) [4/5/2004 5:40:43 PM]


Document<br />

Table 6 Parameters Chosen for the DOCK Runs with T. brucei GAPDH a<br />

Program Variable Value Program Variable Value<br />

SPHGEN dentag X DOCK distance_tolerance 1.0<br />

dotlim 0.0 nodes_maximum 4<br />

radmax 5.0 nodes_minimum 4<br />

radmin 1.0 ligand_binsize 0.4<br />

DISTMAP polcon 2.3 ligand_overlap 0.1<br />

ccon 2.8 receptor_binsize 0.8<br />

discut 4.5 receptor_overlap 0.2<br />

perang 3 atom_minimum 5<br />

a Variables as defined in the DOCK 3.5 manual and discussed in References 79 and 83.<br />

centers to describe the receptor, namely 20 for run 1 and 34 for run 2. According to theory [83], the<br />

difference should scale as 34 4/20 4 = 8.3, which fits our observation. Since it is well known that scores<br />

exhibit poor correlation with real affinities [84] we decided to subject the 200 best-scoring ligands of<br />

each batch to inspection on the graphics. Scanning through the 3200 compounds required about ten<br />

days. Eventually, sixteen compounds were selected for purchase on<br />

Table 7 Parasite GAPDH Inhibitors Discovered with DOCK<br />

Contact Delphi<br />

Inhibitor IC 50 (mM) Score Rank Score Rank<br />

Run 1<br />

2-Guanidinobezimidazole 1.2 97 1432 1.01 1427<br />

2-Benzimidazoylurea 1.8 102 958 0.27 1241<br />

4-Nitrophenyl sulfone 2.8 106 608 -0.46 370<br />

Tryptophan 6.0 128 32 -0.48 357<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_382.html (1 of 2) [4/5/2004 5:40:52 PM]<br />

Page 382


Document<br />

5-Methoxytryptamine 19.0 92 1495 -0.32 488<br />

Ephedrine 25.0 99 1315 -0.22 590<br />

Epinephrine >7 a 102 989 -3.09 27<br />

Aspartic acid dimethyl ester >10 87 1551 -1.66 146<br />

3-Amino-L-tyrosine >11 a 102 969 -0.14 703<br />

1,3-Diphenylguanidine >11 a 112 301 -1.28 209<br />

Octopamine >50 a 106 660 -2.03 83<br />

Run 2<br />

2-Nitrobenzoic acid hydrazide 4.4 132 900 0.11 1112<br />

Dopamine 12.0 143 241 -2.02 156<br />

Norepinephrine >5.4 a 151 84 -2.09 137<br />

4'-Amino-N-methyl acetanilide >7.8 a 131 1027 0.04 1021<br />

L-histidinol >9.3 142 300 0.13 1156<br />

a Could not be tested at higher concentrations due to solubility problems.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_382.html (2 of 2) [4/5/2004 5:40:52 PM]


Document<br />

Figure 10<br />

Binding mode of 2-guanidinobenzimidazole to T.brucei GAPDH as predicted <strong>by</strong><br />

DOCK. The benzimidazole moiety occupies roughly the position occupied <strong>by</strong> adenine<br />

in the holo-enzyme, whereas the guanidino group a salt bridge to Asp37.<br />

the basis of structural rigidity, chemical stability, solubility, and electrostatic complementarity. The<br />

latter property was evaluated with the program DELPHI [85].<br />

Page 383<br />

Each of the sixteen compounds was tested for GAPDH inhibition (Table 7). Half of them were inactive<br />

while the other ones showed inhibition in a range between 1.2 and 25 mM. Unfortunately, there appears<br />

to be no correlation between the DOCK scores and the IC 50 values. For examples, norepinephrine and<br />

1,3-diphenylguanidine are inactive while they have a better score than 2- guanidinobenzimidazole<br />

(Figure 10), the compound with the best IC 50. Also, it appeared that the two different receptor<br />

descriptions used led to almost completely different lists of compounds. Only 156 molecules occurred in<br />

both lists of top-scoring molecules. In the modeled binding mode all of the inhibitors occupy roughly the<br />

same position as the purine ring of NAD in the crystal structure of GAPDH. While the values obtained<br />

for IC 50 are indicative of poor inhibition, one has to keep in mind that adenosine exhibits an IC 50 of 50<br />

mM [13]. By using the program DOCK we were able to discover ligands that have a substantially higher<br />

affinity for GAPDH than the natural ligand.<br />

C. Phosphoglycerate Kinase: Leads from the Past<br />

The starting point for drug design in the case of PGK is quite different from TIM or GAPDH because a<br />

number of nonsubstrate-like inhibitors have been<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_383.html (1 of 2) [4/5/2004 5:41:28 PM]


Document<br />

Figure 11<br />

Two-dimensional structure of SPADNS, a<br />

micromolar inhibitor of T.brucei PGK.<br />

Page 384<br />

discussed in the literature. Suramin inhibits glycosomal PGK of T. brucei with a K i of 8.0 μM [69]. In<br />

addition a number of yeast PGK inhibitors are known: gallic acid with a K i = 0.4 mM [86],<br />

hydroxyethylidene biphosphate with a K i = 24 mM [87]. 1,3,6-naphthalenetrisulphonic acid with a K i =<br />

5.5 mM, and 2-(p- sulphophenylazo)-1,8-dihydroxy-3,6-disulphonic acid, also known as SPADNS<br />

(Figure 11), with a K i = 126 μM [87]. None of the four yeast PGK inhibitors are potent, but, for<br />

SPADNS, the binding mode has been further characterized. Studies <strong>by</strong> Williams et al. [87] have<br />

demonstrated that SPADNS is directly competitive with both enzyme substrates, 3-phosphoglycerate<br />

and ATP. Moreover, <strong>by</strong> 600 MHz 1H-NMR it was shown that SPADNS interacts with the nucleotide<br />

binding site while the conformation of the enzyme changes substantially [87].<br />

Since the four yeast PGK inhibitors are commercially available it was logical to test them for T. brucei<br />

PGK inhibition. The first three compounds were active in the millimolar range. However, SPADNS<br />

exhibited a K i of 10.0 μM in these in preliminary tests [88]. Moreover, when assayed against a<br />

commercially available rabbit muscle PGK, SPADNS had no influence on the enzyme kinetics up to a<br />

concentration of 250 μM [88]. In conclusion, SPADNS appears to be an excellent lead because of its<br />

potency and selectivity. Crystallographic experiments to determine its binding mode to T. brucei PGK<br />

are underway.<br />

IV. Lead Optimization: Glycosomal Gapdh<br />

From the selectivity point of view the adenosine binding site of GAPDH is attractive for drug design, as<br />

we explained in Section II.B. Unfortunately, inhibition studies on T. brucei and L. mexicana GAPDH<br />

revealed the poor affinity of our natural lead adenosine with IC 50 values of 100 mM and 50 mM,<br />

respectively. Moreover, adenosine is an “antiselective” lead because the IC 50<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_384.html [4/5/2004 5:41:34 PM]


Document<br />

for human GAPDH is better than for the parasite enzymes, namely 35 mM [13]. Despite the fact that<br />

these IC 50 values are about ten thousand times higher than what would be considered a lead in the<br />

pharmaceutical industry we decided to optimize the affinity and selectivity of adenosine.<br />

Page 385<br />

Each of the three areas where differences occur between the parasite and the human enzyme are<br />

hydrophobic. Therefore, we modeled hydrophobic substituents at positions C2, C8, and O2' of adenosine<br />

under the constraint that they were conformationally compatible with the C2'-endo pucker of the ribose<br />

sugar. <strong>Design</strong>ing derivatives at O2' was a problem, however. Each of the two ribosyl hydroxyls forms a<br />

hydrogen bond with the carboxylate of Asp37. Since making direct derivatives of the hydroxyl, such as<br />

ethers or esters, would deprive the Asp of a hydrogen-bond partner while burying the carboxylate,<br />

resulting molecules would have a dramatically reduced affinity. Moreover, an alignment of 47 GAPDH<br />

sequences made it clear that the Asp is highly conserved [89]. An elegant way to overcome this problem<br />

was to replace the 2' -hydroxyl <strong>by</strong> a 2'- amino function. Moreover, coupling with carboxylic acids was<br />

appealing from a synthetic point of view while the conformational properties of the amido-substituted<br />

system would ensure the correct orientation of substituents into the selectivity cleft. The modeled<br />

inhibitors were evaluated for the quality of their fit to the protein surface and subsequently synthesized.<br />

From Table 8 it can be seen that our predictions were successful. The addition of a methyl group at C2<br />

of the adenine ring, which is close to Val36, increased the affinity for parasite GAPDH <strong>by</strong> an order of<br />

magnitude. The effect of a thienyl substitution on C8, targeted to Leu112, was even bigger, namely two<br />

orders of magnitude. However, both substitutions are only mildly selective (Table 8). As expected, the<br />

greatest gain in selectivity was obtained <strong>by</strong> modifying the 2'-position of the ribose, so that the selectivity<br />

cleft is filled up (Figure 12). The 2'-deoxy-2'-(3-methoxybenzamido) adenosine compound (Figure 13)<br />

bound at least 48 times better to L. mexicana GAPDH than to the human enzyme. The selectivity versus<br />

T. brucei GAPDH appeared to be smaller. This has to be ascribed to a difference in residues contacting<br />

the 3-methoxy moeity. The residue Asn39 of T. brucei GAPDH has a Ser equivalent in the L. mexicana<br />

Table 8 Inhibition Gains of <strong>Design</strong>ed GAPDH Inhibitors with Respect to Adenosine<br />

C2-subst C8-subst C2'-subst T. brucei L. mexicana human<br />

CH 3 H OH 12.5 6.25


Document<br />

Figure 12<br />

Predicted binding mode of 2' -deoxy-2'-(3-methoxybenzamido)adenosine to T.brucei<br />

GAPDH. (From Ref. 13. Copyright 1994 <strong>by</strong> the American Chemical Society.)<br />

enzyme. In conclusion, the strategy of burying hydrophobic residues with lipophilic substituents paid<br />

off.<br />

Page 386<br />

Despite the rather poor IC 50 values of our optimized inhibitors, an evaluation of their effect on live<br />

trypanosomes appeared to be useful. Enzyme inhibitors that are not taken up <strong>by</strong> the parasites would be<br />

of no use as a drug. Therefore, the effect of 2-methyl-adenosine, 8-(thien-2-yl)-adenosine and 2'- deoxy-<br />

2'-(3-methoxybenzamido)adenosine on the growth of T. brucei in cultures, as described <strong>by</strong> Baltz et al.<br />

[90], was monitored. At 0.1 mM all compounds inhibited the growth completely, unlike adenosine<br />

derivatives that were without inhibitory effect against T. brucei GAPDH [91]. Experiments are<br />

underway to confirm that the growth inhibition is due to blockage of the glycolytic pathway. Also, te<br />

mechanism of uptake of te inhibitors will be examined because it is now well established that<br />

trypanosomes possess a unique P 2 purine transporter that they use for uptake of purines from the host<br />

[15]. The experimental antitrypanosomal drug 5'-{[(Z)-4-amino-2-butenyl}methylamino}-5'deoxyadenosine(MDL<br />

73811) Figure 13), which is an irreversible S- adenosyl-L-methionine<br />

decarboxylase inhibitor, is actively taken up through the P2 transporter. Moreover, MDL 73811 is not<br />

actively transported in the human host, which presumably contributes to the drug's selectivity [19]. It is<br />

not unthinkable that our inhibitors might use the same transporter because of the nature of their scaffold,<br />

adenosine.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_386.html (1 of 2) [4/5/2004 5:42:07 PM]


Document<br />

Figure 13<br />

Two-dimensional structures of 2'-deoxy-2'-(3-methoxybenzamido)<br />

adenosine (MMBA), a selective T.brucei GAPDH inhibitor and<br />

MDL 73811, an irreversible<br />

inhibitor of trypanosomal S-adenosyl<br />

methionine decarboxylase.<br />

Page 387<br />

Finally, we want to point out that pessimism about adenosine derivatives as drugs is not necessarily<br />

warranted. This doubt stems from the argument that many proteins recognize NAD(P), adenosine, or<br />

ATP. Cross-reactivity of adenosine with these different proteins and, therefore, toxicity may be<br />

expected. That this is not necessarily so is evident from the use of adenosine derivatives as antileukemia<br />

agents. For example, fludarabine, a C2'-epimer of adenosine, exhibits relatively low toxicity [92]. The<br />

much bigger changes to the adenosine scaffold in our inhibitors may hence lead to a surprisingly high<br />

overall selectivity.<br />

V. Conclusions<br />

Our goal is to discover and design selective inhibitors of trypanosomal glycolsysis. Thusfar, three<br />

enzymes have been targeted. Whereas little success was obtained with TIM, substantial progress is being<br />

made with GAPDH and PGK.<br />

Obstacles encountered during this project were the need for selective inhibitors and the absence of<br />

potent inhibitors as lead compounds. From our studies to inhibit TIM it appears that it is difficult to<br />

come up with inhibitors for areas on the protein surface for which no known inhibitors exist. On the<br />

other hand, lead optimization for GAPDH <strong>by</strong> over two orders of magnitude in just one cycle of drug<br />

design was straightforward. Selectivity was obtained <strong>by</strong> using part of the cofactor as a lead and<br />

exploiting the hydrophobic patches at the surface of the parasite enzyme. In particular, 2'-deoxy-2' (3methoxy-benzamido)ade-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_387.html (1 of 2) [4/5/2004 5:42:15 PM]


Document<br />

Page 388<br />

nosine proved to inhibit the parasite more than fifty times better than the human enzyme. Additionally,<br />

<strong>by</strong> means of the program DOCK, eight new leads for GAPDH inhibition were found. None of them<br />

were micromolar inhibitors but all of them were more potent than the natural lead, adenosine. For<br />

trypanosomal PGK a potent lead compound, SPADNS, was discovered <strong>by</strong> testing inhibitors that had<br />

been described as weak inhibitors of yeast PGK. Moreover, this lead had no effect on mammalian PGK<br />

at concentrations up to twenty-five times higher than that needed for T. brucei inhibition. At present,<br />

none of our inhibitors is potent enough to consider clinical tests. There is hope, however, since our<br />

GAPDH inhibitors inhibit the growth of trypanosomes in cultures.<br />

Acknowledgments<br />

It is a pleasure to thank the many colleagues and collaborators who have contributed to this project: Paul<br />

Michels, Veronique Hannaert, Sylvie Allert, and Linda Kohl (Institute for Cellular Pathology in<br />

Brussels) for cloning and overexpressing trypanosomatid enzymes; Phil Petra (University of<br />

Washington, Seattle) for helping us out with protein purification protocols; Mia Callens and Fred<br />

Opperdoes (Institute for Cellular Pathology in Brussels) for enzymology and parasitology; Rik<br />

Wierenga, Martin Noble, Fred Vellieux, Randy Read, Risto Lapatto, Hillie Groendijk, Tjaard Pijning,<br />

and Kor Kalk for laying the structural foundation of the project in Groningen and Heidelberg; Cees<br />

Witmans and the late Alan Horn (University of Groningen), Michèle Willson and Jacques Perié<br />

(University of Toulouse), Serge Van Calenbergh, Arthur Van Aerschot, and Piet Herdewijn (University<br />

of Leuven) for synthesizing TIM and GAPDH inhibitors; Kim Simons (University of Washington) for<br />

going after completely new GAPDH inhibitors; Véronique Mainfroid and Joseph Martial (University of<br />

Liège) for providing human TIM; Klaus Muml;uller and Klaus Gubernator (Hoffmann-La Roche, Basel)<br />

for valued modeling advice; and Mike Gelb (University of Washington) for valued discussions.<br />

Financial support for these investigations has been provided <strong>by</strong> the World Health Organization,<br />

Hoffmann-LaRoche in Basel, the Dutch Organization for the Advancement of Science (NWO), the STD<br />

program of the European Community, the School of Medicine of the University of Washington, and the<br />

Murdock Charitable Trust.<br />

Note Added in Proof<br />

We just solved the ternary structure of PGK from Trypanosoma brucei in complex with ADP and 3phosphoglycerate.<br />

The enzyme is in the closed conformation that has eluded crystallographers for 20<br />

years.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_388.html [4/5/2004 5:42:18 PM]


Document<br />

References<br />

1. Kuzoe FAS. Current situation of African trypanosomiasis. Acta Trop 1993; 54:153–162.<br />

2. Kuzoe F. African trypanosmiasis. In: Tropical Disease Research Eleventh Programme Report.<br />

Geneva: World Health Organization, 1993:57–66.<br />

3. Herwaldt BL, Juranek DD. Laboratory-acquired malaria, leishmaniasis, trypanosomiasis, and<br />

toxoplasmosis. Am J Trop Med Hyg 1993; 48:313–323.<br />

Page 389<br />

4. Mc Cann PP, Bacchi CJ, Bitonti AJ, Kierszenbaum F, Sjoerdsma A. Inhibition of ornithine or<br />

arginine decarboxylase as an experimental approach to African or American trypanosomiasis. Adv Exp<br />

Med Biol 1988;250:727–735.<br />

5. Wang CC. Molecular mechanisms and therapeutic approaches to the treatment of African<br />

trypanosomiasis. Annu Rev Pharmacol Toxicol 1995; 35:93–127.<br />

6. Pépin J, Milord F. The treatment of human African trypanosomiasis. Adv Parasitol 1994; 33:1–47.<br />

7. TDR News. 1990; No. 34:1–2.<br />

8. Aldhous P. Fighting parasites on a shoestring. Science 1994; 264:1857–1859.<br />

9. Schechter PJ, Barlow JLR, Sjoerdsma A. Clinical aspects of inhibition of ornithine decarboxylase<br />

with emphasis on therapeutic trials of eflornithine (DFMO) in cancer and protozoan diseases. In: Mc<br />

Cann PP, Pegg AE, Sjoerdsma A, eds. Inhibition of Polyamine Metabolism. New York: Acad Press,<br />

Inc., 1987: 345–364.<br />

10. Borst P, Rudenko G. Antigenic variation in African trypanosomes. Science 1994; 264:1872–1873.<br />

11. Nilsen TW. Trans-splicing in protozoa and helminths. Infect Agents Dis 1992; 1:212–218.<br />

12. Opperdoes FR. Compartmentation of carbohydrate metabolism in trypanosomes. Annu Rev<br />

Microbiol 1987; 41:127–151.<br />

13. Verlinde CLMJ, Callens M, Van Calenbergh S, Van Aerschot A, Herdewijn P, Hannaert V, Michels<br />

PAM, Opperdoes FR, Hol WGJ. Selective inhibition of trypanosomal glyceraldehyde-3-phosphate<br />

dehydrogenase <strong>by</strong> protein structure<strong>based</strong> design: toward new drugs for the treatment of sleeping<br />

sickness. J Med Chem 1994; 37:3605–3613.<br />

14. Seyfang A, Duszenko M. Specificity of glucose transport in Trypanosoma brucei. Effective<br />

inhibition <strong>by</strong> phloretin and cytochalasin B. Eur J Biochem 1991; 202:191–196.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_389.html (1 of 2) [4/5/2004 5:42:54 PM]


Document<br />

15. Carter NS, Fairlamb AH. Arsenical-resistant trypanosomes lack an unusual adenosine transporter.<br />

Nature 1993; 361:173–176.<br />

16. Ullman B, Allen TE. Hypoxanthine-guanine phosphoribosyltransferase in trypanosomatids: a<br />

rational target for antiparasitic chemotherapy. In: Boothroyd J, ed. Molecular Approaches to<br />

Parasitology. New York: Wiley-Liss, Inc., 1995:123–141.<br />

17. Wang CC. A novel suicide inhibitor strategy for antiparasitic drug design. J Cell Biochem 1991;<br />

45:49–53.<br />

18. Phillips MA, Coffino P, Wang CC. Cloning and sequencing of the ornithine decarboxylase gene<br />

from Trypanosoma brucei. Implications for enzyme turnover and selective difluoromethylornithine<br />

inhibition. J Biol Chem 1987; 262:8721–8727.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_389.html (2 of 2) [4/5/2004 5:42:54 PM]


Document<br />

Page 390<br />

19. Byers TL, Castara P, Bitonti AJ. Uptake of the antitrypanosomal drug 5'-{[(Z)-4- amino-2butenyl]methylamino}-5'-deoxyadenosine<br />

(MDL 73811) <strong>by</strong> the purine transport system of Trypanosoma<br />

brucei brucei. Biochem J 1992; 283:755–758.<br />

20. Schirmer RH, Müller JG, Krauth-Siegel RL. Disulfide-reductase inhibitors as chemotherapeutic<br />

agents: the design of drugs for trypanosomias is and malaria. Angew Chem Int Ed Engl 1995;<br />

34:141–154.<br />

21. Doering TL, Raper J, Buxbaum LU, Adams SP, Gordon JI, Hart GW, Englund PT. An analog of<br />

myristic acid with selective toxicity for African trypanosomes. Science 1991; 252:1851–1854.<br />

22. Shaw PJ, Muirhead H. Crystallographic structure analysis of glucose-6-phosphate isomerase at 3.5 Å<br />

resolution. J Mol Biol 1977; 109:475–485.<br />

23. Gamblin SJ, Muirhead H. Crystallographic structure analysis of glucose-6-phosphate isomerase at<br />

3.5 Å resolution. J Mol Biol 1977; 109:475–485.<br />

23. Gamblin SJ, Cooper B, Millar JR, Davies GJ, Littlechild JA, Watson HC. The crystal structure of<br />

human muscle aldolase at 3.0 Å resolution. FEBS Lett 1990; 262:282–286.<br />

24. Wierenga RK, Noble MEM, Vriend G, Nauche S, Hol WGJ. Refined 1.83 Å structure of<br />

triosephosphate isomerase crystallized in the presence of 2.4 M ammonium sulphate. A comparison with<br />

the structure of the trypanosomal triosephosphate isomerase—glycerol-3-phosphate complex. J Mol Biol<br />

1991; 220:995–1015.<br />

25. Mande SC, Mainfroid V, Kalk KH, Goraj K, Martial JA, Hol WGJ. Crystal structure of recombinant<br />

human triosephosphate isomerase at 2.8 Å resolution. Triosephosphate isomerase-related human genetic<br />

disorders and comparison with the trypanosomal enzyme. Protein Sci 1994; 3:810–821.<br />

26. Vellieux FMD, Hajdu J, Verlinde CLMJ, Groendijk H, Read RJ, Greenhough TJ, Campbell JW,<br />

Kalk KH, Littlechild JA, Watson HC, Hol WGJ. <strong>Structure</strong> of glycosomal glyceraldehyde-3-phosphate<br />

dehydrogenase from Trypanosoma brucei determined from Laue data. Proc Natl Acad Sci USA 1993;<br />

90:2355–2359.<br />

27. Vellieux FMD, Hajdu J, Hol WGJ. Refined 3.2 Å structure of glycosomal holo glyceraldehyde<br />

phosphate dehydrogenase from Trypanosoma brucei brucei. Acta Cryst 1995; D 51:575–589.<br />

28. Mercer WD, Winn SI, Watson HC. Twinning in crystals of human skeletal muscle D-glyceraldehyde-<br />

3-phosphate dehydrogenase. J Mol Biol 1976; 104:277–283.<br />

29. Read RJ, et al. Unpublished results.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_390.html (1 of 2) [4/5/2004 5:43:00 PM]


Document<br />

30. Kim H, Feil I, Verlinde CLMJ, Petra PH, Hol WGJ. Crystal structure of glycosomal glyceraldehyde-<br />

3-phosphate dehydrogenase from Leishmania mexicana: Implications for structure-<strong>based</strong> drug design<br />

and a new position for the inorganic phosphate binding site. Biochem 1995; 34:14975–14986.<br />

31. Rice DW. The use of phase combination in the refinement of phosphoglycerate kinase at 2.5 Å<br />

resolution. Acta Cryst A 1981; 37:491–500.<br />

32. Harlos K, Vas M, Blake CF. Crystal structure of the binary complex of pig muscle phosphogycerate<br />

kinase and its substrate 3-phospho-D-glycerate. Proteins: Struct Funct Genet 1992; 12:133–144.<br />

33. Stuart DI, Levine M, Muirhead H, Stammers DK. Crystal structure of cat muscle pyruvate kinase at<br />

a resolution of 2.6 Å J Mol Biol 1979; 134:109–142.<br />

34. Lantwin CB, Schlichting I, Kabsch W, Pai EF, Krauth-Siegel RL. The structure of Trypanosoma<br />

cruzi trypanothione reductase in the oxidized and NADH reduced state. Proteins: Struct Funct Genet<br />

1994; 18:161–173.<br />

35. Karplus PA, Schulz GE. Refined structure of glutathione reductase at 1.54 Å resolution. J Mol Biol<br />

1987; 195:701–729.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_390.html (2 of 2) [4/5/2004 5:43:00 PM]


Document<br />

Page 391<br />

36. Kuriyan J, Kong XP, Krishan TS, Sweet RM, Murgolo NJ, Field H, Cerami A, Henderson GB. Xray<br />

structure of trypanothione reductase from Crithidia fasciculata at 2.4 Å resolution. Proc Natl Acad<br />

Sci USA 1991; 88:8764–8768.<br />

37. Hunter WN, Bailey S, Habash J, Harrop SJ, Helliwell JR, Aboagye-Kwarteng T, Smith K, Fairlamb<br />

AH. Active site of trypanothione reductase. A target for rational drug design. J Mol Biol 1992;<br />

227:322–333.<br />

38. Eads JC, Scapin G, Xu Y, Grubmeyer C, Sacchettini JC. The crystal structure of human<br />

hypoxanthine-guanine phosphoribosyltransferase with bound GMP. Cell 1994; 78:325–334.<br />

39. Freymann D, Down J, Carrington M, Roditi I, Turner M, Wiley D. 2.9 Å resolution structure of the<br />

N-terminal domain of a Variant Surface Glycoprotein from Trypanosoma brucei. J Mol Biol 1990;<br />

216:141–60.<br />

40. Izzo P, Constanzo P, Luppo A, Rippa E, Paolella G, Salvatore F. Human aldolase A gene. Structural<br />

organization and tissue-specific expression <strong>by</strong> multiple promotors and alternate mRNA processing. Eur<br />

J Biochem 1988; 174:569–578.<br />

41. Paolella G, Santamaria R, Izzo P, Constanzo P, Salvatore F. Isolation and nucleotide sequence of a<br />

full-length c-DNA coding for aldolase B from human liver. Nucl Acids Res 1984; 12:7401–7410.<br />

42. Rottmann WH, Deselms KR, Niclas J, Camerato T, Holman PS, Green CJ, Tolan DR. The complete<br />

amino acid sequence of the human aldolase C isozyme drived from genomic clones. Biochim 1987;<br />

69:137–145.<br />

43. Grant PT, Sargent JR. Properties of L-α-glycerophosphate oxidase and its role in the respiration of<br />

Trypanosoma rhodesiense. Biochem J 1960; 76:229–237.<br />

44. Clarkson AB, Brohn FH. Trypanosomiasis: an approach to chemotherapy <strong>by</strong> the inhibition of<br />

carbohydrate metabolism. Science 1976; 194:204–206.<br />

45. Fairlamb AH, Opperdoes FR, Borst P. New approach to screening drugs for activity against African<br />

trypanosomes. Nature 1977; 265:270–271.<br />

46. Evans DA, Brightman CJ, Holland MF. Salicylhydroxamic acid/glycerol in experimental<br />

trypanosomiasis. Lancet 1977; 769.<br />

47. Van Der Meer C, Versluijs-Broers JAM, Opperdoes FR. Trypanosoma brucei: trypanocidal effect of<br />

salicylhydroxamic acid plus glycerol in infected rats. Exp Parasitol 1979; 48:126–134.<br />

48. Opperdoes FR, Borst P. Localization of nine glycolytic enzymes in a microbody-like organelle in<br />

Trypanosoma brucei: the glycosome. FEBS Létt 1977; 80:360–364.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_391.html (1 of 2) [4/5/2004 5:43:14 PM]


Document<br />

49. Michels PAM, Hannaert V. The evolution of kinetoplastid glycosomes. J Bioenerg Biomembr 1994;<br />

26:213–219.<br />

50. Borchert TV, Pratt K, Zeelen JP, Callens M, Noble MEM, Opperdoes FR, Michels PAM, Wierenga<br />

RK. Overexpression of trypanosomal triosephosphate isomerase in Escherichia coli and characterization<br />

of a dimer-interface mutant. Eur J Biochem 1993; 211:703–710.<br />

51. Verlinde CLMJ, Noble MEM, Kalk KH, Groendijk H, Wierenga RK, Hol WGJ. Anion binding at<br />

the active site of trypanosomal triosephosphate isomerase: monohydrogen phosphate does not mimic<br />

sulphate. Eur J Biochem 1991; 198:53– 57.<br />

52. Noble MEM, Verlinde CLMJ, Groendijk H, Kalk KH, Wierenga RK, Hol WGJ. Crystallographic<br />

and molecular modeling studies on trypanosomal triosephosphate isomerase: a critical assessment of the<br />

predicted and observed structures of the complex with 2-phospho-glycerate. J Med Chem 1991;<br />

34:2709–2718.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_391.html (2 of 2) [4/5/2004 5:43:14 PM]


Document<br />

Page 392<br />

53. Noble MEM, Wierenga RK, Lambeir AM, Opperdoes FR, Thunnissen AMWH, Groendijk H, Hol<br />

WGJ. The adaptability of the active site of trypanosomal triosephosphate isomerase as observed in the<br />

crystal structures of three different complexes. Proteins: Struct Funct Genet 1991; 10:50–69.<br />

54. Noble MEM, Witmans CJ, Verlinde CLMJ, Wierenga RK, Hol WGJ. Unpublished results.<br />

55. Verlinde CLMJ, Witmans CJ, Pijning T, Kalk KH, Hol WGJ, Callens M, Opperdoes FR. <strong>Structure</strong><br />

of the complex between trypanosomal triosephosphate isomerase and N-hydroxy-4-phosphonobutanamide:<br />

binding at the active site despite an “open” flexible loop. Protein Sci 1992; 1:1578–1584.<br />

56. Knowles JR. Enzyme catalysis: not different, just better. Nature 1991; 350:121–124.<br />

57. Lambeir AM, Opperdoes FR, Wierenga RK. Kinetic properties of triose-phosphate isomerase from<br />

Trypanosoma brucei brucei. A comparison with the rabbit muscle and yeast enzymes. Eur J Biochem<br />

1987; 168:69–74.<br />

58. Verlinde CLMJ, Rudenko G, Hol WGJ. In search of new lead compounds for trypanosomiasis drug<br />

design: a protein structure-<strong>based</strong> linked-fragment approach. J Comput-Aided <strong>Drug</strong> Des 1992;<br />

69:131–147.<br />

59. Hannaert V, Opperdoes FR, Michels PAM. Glycosomal glyceraldehyde-3- phosphate dehydrogenase<br />

of Trypanosoma brucei and Trypanosoma cruzi: expression in Escherichia coli, purification, and<br />

characterization of the enzymes. Protein Expr Purif 1995; 6:244–250.<br />

60. Michels PAM, Poliszczak A, Osinga KA, Misset O, Van Beeumen J, Wierenga RK, Borst P,<br />

Opperdoes FR. Two tandemly linked identical genes code for the glycosomal glyceraldehyde-phosphate<br />

dehydrogenase in Trypanosoma brucei. EMBO J 1986; 5:1049–1056.<br />

61. Lambeir AM, Loiseau AM, Kuntz DA, Vellieux FM, Michels PAM, Opperdoes FR. The cytosolic<br />

and glycosomal glyceraldehyde-3-phosphate dehydrogenase from Trypanosoma brucei. Kinetic<br />

properties and comparison with homologous enzymes. Eur J Biochem 1991; 198:429–435.<br />

62. Hannaert V, Callens M, Opperdoes FR, Michels PAM. Purification and characterization of the<br />

native and recombinant Leishmania mexicana glycosomal glyceraldehyde-3-phosphate dehydrogenase.<br />

Eur J Biochem 1994; 225:143–149.<br />

63. Desjeux P. Human leishmaniases: epidemiology and health aspects. World Health Stat Q 1992;<br />

45:267–275.<br />

64. Berman JD, Gallalee JV, Best JM. Sodium stibogluconate (Penstostam) inhibition of glucose<br />

catabolism via the glycolytic pathway, and fatty acid β-oxidation in Leishmania mexicana amastigotes.<br />

Biochem Pharmacol 1987; 36:197–201.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_392.html (1 of 2) [4/5/2004 5:43:36 PM]


Document<br />

65. João HC, Williams RJP. The anatomy of a kinase and the control of phosphate transfer. Eur J<br />

Biochem 1993; 216:1–18.<br />

66. Michels PAM, personal communication.<br />

67. Huang IY, Welch CD, Yoshida A. Complete amino acid sequence of human phosphoglycerate<br />

kinase. Cyanogen bromide peptides and complete amino acid sequence. J Biol Chem 1980;<br />

255:6412–6420.<br />

68. McCarrey JR, Thomas K. Human testis-specific PGK gene lacks introns and possesses<br />

characteristics of a processed gene. Nature 1987; 326:501–504.<br />

69. Davies GJ, Gamblin SJ, Littlechild JA, Dauter Z, Wilson KS, Watson HC. <strong>Structure</strong> of the ADP<br />

complex of the 3-phosphoglycerate kinase from Bacillus stearothermophilus at 1.65 Å. Acta Cryst D<br />

1994; 50:202–209.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_392.html (2 of 2) [4/5/2004 5:43:36 PM]


Document<br />

Page 393<br />

70. Banks RD, Blake CC, Evans PR, Haser R, Rice DW, Hardy GW, Merrett M, Phillips AW.<br />

Sequence, structure, and activity of phosphoglycerate kinase: a possible hinge-bending enzyme. Nature<br />

1979; 279:773–777.<br />

71. Misset O, Opperdoes FR. The phosphoglycerate kinases from Trypanosoma brucei. A comparison of<br />

the glycosomal and the cyosolic isoenzymes and their sensitivity towards suramin. Eur J Biochem 1987;<br />

162:493–500.<br />

72. Osinga KA, Swinkels BW, Gibson WC, Borst P, Veeneman GH, Van Boom JH, Michels PAM,<br />

Opperdoes FR. Topogenesis of microbody enzymes: a sequence comparison of the genes for the<br />

glycosomal (microbody) and cytosolic phosphoglycerate kinases of Trypanosoma brucei. EMBO J<br />

1985; 4:3811–3817.<br />

73. Verlinde CLMJ, Pijning T, Kalk KH, Van Calenbergh S, Van Aerschot A, Herdewijn P, Callens M,<br />

Michels P, Opperdoes FR, Wierenga RK, Hol WGJ. Protein structure-<strong>based</strong> inhibitor design: towards<br />

new drugs for sleeping sickness. In: Testa B, Kyburz E, Fuhrer W, Giger R, eds. Perspectives in<br />

Medicinal Chemistry. Basel: Verlag Helvetica Chimica Acta, 1993:135–148.<br />

74. Goodford PJ. A computational procedure for determining energetically favorable binding sites on<br />

biologically important macromolecules. J Med Chem 1985; 28:849–857.<br />

75. Miranker A, Karplus M. Functionality maps of binding sites: a multiple copy simultaneous search<br />

method. Proteins: Struct, Funct, and Genetics 1991; 11:29–34.<br />

76. Tronrud DE, Ten Eyck LF, Matthews BA. An efficient general-purpose least- squares refinement<br />

program for macromolecules. Acta Cryst A 1987; 43:489–501.<br />

77. Read RJ. Improved coefficients for maps using phases from partial structures with errors. Acta Cryst<br />

A 1986; 42:140–149.<br />

78. Kuntz ID. <strong>Structure</strong>-<strong>based</strong> strategies for drug design and discovery. Science 1992; 257:1078–1082.<br />

79. Meng EC, Shoichet BK, Kuntz ID. Automated docking with grid-<strong>based</strong> energy evaluation. J Comput<br />

Chem 1992; 13:505–524.<br />

80. ACD-3D. MDL Information Systems Inc, San Leandro, California.<br />

81. Rappé AK, Goddard WA III. Charge equilibration for molecular dynamics simulations. J Phys Chem<br />

1991; 95:3358–3363.<br />

82. BIOGRAF. Molecular Simulations Inc, Burlington, Massachusetts.<br />

83. Shoichet BK, Bodian DL, Kuntz ID. Molecular docking using shape descriptors. J Comput Chem<br />

1992; 13:380–397.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_393.html (1 of 2) [4/5/2004 5:43:57 PM]


Document<br />

84. Verlinde CLMJ, Hol WGJ. <strong>Structure</strong>-<strong>based</strong> drug design: progress, results and challenges. Struct<br />

1994; 2:577–587.<br />

85. Gilson MK, Sharp KA, Honig BH. Calculating the electrostatic potential of molecules in solution:<br />

method and error assessment. J Comput Chem 1987; 9:327–335.<br />

86. João HC, Williams RJP, Littlechild JA, Nagasuma R, Watson HC. An investigation of large<br />

inhibitors binding to phosphoglycerate kinase and their effect on anion activation. Eur J Biochem 1992;<br />

205:1077–1088.<br />

88. Bernstein B et al., unpublished results.<br />

87. Boyle HA,<br />

Fairbrother WJ,<br />

Williams RJP.<br />

An NMR<br />

analysis of<br />

inhibitors to<br />

yeast<br />

phosphoglycerate<br />

kinase. Eur J<br />

Biochem 1989;<br />

184:535–543.<br />

89. Fothergill-Gilmore LA, Michels PAM. Evolution of glycolysis.<br />

Prog Biophys Mol Biol 1993; 59:105–235.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_393.html (2 of 2) [4/5/2004 5:43:57 PM]


Document<br />

90. Baltz T, Baltz D, Giroud C, Crockett J. Cultivation in a semi-defined medium of animal infective<br />

forms of Trypanosoma brucei, T. equiperdum, T. evansi, T. rhodesiense and T. gambiense. EMBO J<br />

1985; 4:1273–1277.<br />

Page 394<br />

91. Opperdoes FR, personal communication.<br />

92. Whelan JS, Davis CL, Rule S, Ransom M, Smith OP, Metha AB, Catovsky D, Rohatiner AZ, Lister<br />

TA. Fludarabine phosphate for the treatment of low grade lymphoid malignancy. Br J Cancer 1991;<br />

64:120–123.<br />

93. Kraulis PJ. MOLSCRIPT: a program to produce both detailed and schematic plots of protein<br />

structures. J Appl Crystallography 1991; 24:946–950.<br />

94. Bernstein BE, Michels PAM, Hol WGJ. Synergistic effects of substrate-induced conformational<br />

changes in phosphoglycerate kinase activation. Nature 1997; 385:275–278.<br />

95. Michels PAM. Biology of the Cell 1988; 64:157–164.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_394.html [4/5/2004 5:43:59 PM]


Document<br />

16<br />

Progress in the <strong>Design</strong> of Immunomodulators Based on the <strong>Structure</strong> of<br />

Interleukin-1<br />

Glen Spraggon<br />

University of California, San Diego, La Jolla, California<br />

Pandi Veerapandian<br />

Axiom Biotechnologies, Inc., San Diego, California, and<br />

La Jolla Institute for Experimental Medicine, La Jolla, California<br />

I. Introduction<br />

Page 395<br />

The interleukin-1 (IL-1) family of cytokines exhibit both normal and pathological effects in almost<br />

every tissue and organ system and, as such, have been associated with cells engaged in the immune<br />

response, inflammatory cells, and cells engaged in development, differentiation, and repair processes<br />

[1–4]. Interleukin-1 can produce either a direct response on one specific target cell or act as an indirect<br />

effector molecule, inducing the expression of a variety of genes and synthesis of several proteins such as<br />

IL-2–IL-8, tumor necrosis factor (TNF), colony-stimulating factors (CSF), platelet-derived factors<br />

(PDFs), and other cytokines. Pathologically uncontrolled, IL-1 activity can induce disease states and has<br />

been linked to septic shock, the growth of acute and chronic myelogenous leukemia cells, inflammation<br />

associated with arthritis and colitis, development of atherosclerotic plaques, insulin-dependent diabetes,<br />

osteoporosis, parsitemia, and cancer [1,5,6]. Strategies to treat such diseases are being developed and<br />

usually involve the inhibition of IL-1 synthesis or the blocking of its activity [7]. The therapeutic<br />

advantage of reducing the activity of IL-1 resides in preventing its deleterious biological effects without<br />

interfering with homeostasis. In order to achieve such a goal, an understanding of the structure-function<br />

relationship of the IL-1 family of proteins at the molecular level is important. Such a knowledge could<br />

lead to the design and development of<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_395.html [4/5/2004 5:44:01 PM]


Document<br />

Page 396<br />

synthetic agents with the ability to modulate IL-1 responses. Here we review the present knowledge of<br />

the IL-1 system and the attempts toward the development of its modulators.<br />

A. Cytokines and Physiological Responses<br />

Cytokines are multifunctional hormones produced <strong>by</strong> a variety of cells to carry out a wide range of<br />

overlapping biological actions including communication within the immune system and between other<br />

cell types. They are produced <strong>by</strong> macrophages, endothelial cells, fibroblasts, keratinocytes, T cells, B<br />

cells, and natural killer cells in response to injury and infection [7]. Upon production they act either<br />

locally or as systemic intercellular signaling factors and induce the production of other cytokines for<br />

continuing their communications via cell surface receptors. The cytokines network also has associated<br />

with it a complementary set of soluble or membrane-bound antagonist or mediator molecules that are<br />

capable of shutting down these effects [8]. During infection and injury, the leukocytes adhere<br />

themselves to the endothelium and migrate into tissues. Cells like T-lymphocytes, monocytes,<br />

macrophages, and neutrophils collect themselves as inflammatory infiltrate. Once within the tissues,<br />

these cells neutralize the microorganisms or infected cells, secreting other cytokines that can modulate<br />

the adhesion molecule expression on the endothelium. In addition, chemotactic signals are delivered to<br />

cells passing in the circulation. Cytokines are also involved in tissue repair, i.e., remodeling of<br />

connective tissues and revascularization of damaged areas. Soluble mediators are released from the site<br />

of damage into the circulation to act in an endocrine fashion. The cytokines then control hepatic<br />

responses to tissue damage. During the regulation of hematopoiesis, cytokines induce the production and<br />

release of a number of colony-stimulating factors and other interleukins. Thus induced factors are<br />

responsible for the replacement of leukocytes and erythrocytes that were lost after trauma.<br />

A large number of cytokine molecules have been characterized [7]. Structurally it appears that these<br />

cytokines can be classified into the following main groups <strong>based</strong> on their folding pattern: 4-helix<br />

bundles, beta-trefoil, beta sandwich, EGF-like, beta cysteine knot, and alpha/beta. Thus far the structures<br />

of many representatives of each family have been solved <strong>by</strong> both x-ray crystal- lography and NMR<br />

(Table 1). Extensive experimentation has indicated that despite a large degree of structural diversity,<br />

there is a large degeneracy in the cytokine network and the molecules have a wide range of overlapping<br />

biological functions. For example, IL-1, TNF, PDGF, and TGFB all exhibit similar biological behavior.<br />

The structures of the receptors for these molecules have not been forthcoming and only three examples<br />

of the extracellular domains of the receptors have been reported. They are growth hormone receptor in<br />

complex<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_396.html [4/5/2004 5:44:02 PM]


Document<br />

Table 1 Known Three-Dimensional <strong>Structure</strong>s of Cytokines<br />

Brookhaven <strong>Structure</strong><br />

Molecule Type Code Determined <strong>by</strong><br />

Interleukin-1 alpha (IL-1α) Beta trefoil 2ILA X-ray crystallography<br />

Interleukin-1 beta (IL-1β) Beta trefoil 1ILB, 2ILB, 4ILB,<br />

5ILB, 6ILB<br />

Interleukin-1 receptor antagonist<br />

(IL-1ra)<br />

Fibroblast growth factor (acidic)<br />

(aFGF)<br />

Fibroblast growth factor (basic)<br />

(bFGF)<br />

Granulocyte-colony stimulating<br />

factor (G-CSF)<br />

X-ray crystallography and<br />

NMR<br />

Beta trefoil 1ILR, 1IRP X-ray crystallography and<br />

NMR<br />

Beta trefoil 1AFC X-ray crystallography<br />

Beta trefoil 1BFG X-ray crystallography<br />

Long-Chain 4 helix<br />

bundle<br />

Growth hormone Long-Chain 4 helix<br />

bundle<br />

Leukemia inhibitory factor (LIF) Long-Chain 4 helix<br />

bundle<br />

Granulocyte-macrophage colony<br />

stimulating factor (GM-CSF)<br />

Short-Chain 4 helix<br />

bundle<br />

Interleukin-2 (IL-2) Short-Chain 4 helix<br />

bundle<br />

Interleukin 4 (IL-4) Short-Chain 4 helix<br />

bundle<br />

1RHG, 1GNC X-ray crystallography<br />

1HGU X-ray crystallography<br />

1LKI X-ray crystallography<br />

1CSG, 1GMF X-ray crystallography<br />

3INK X-ray crystallography<br />

1BBN, 1ITM,<br />

2CYK<br />

NMR<br />

Interferon gamma (IFN-γ) Dimeric 4 helix bundle 1IKI X-ray crystallography<br />

Interleukin 10 (IL-10) Dimeric 4 helix bundle 1ILK X-ray crystallography<br />

Nerve growth factor (NGF) Beta cysteine knot 1BET, 1BTG X-ray crystallography<br />

Platelet derived growth factor<br />

(PDGF)<br />

Beta cysteine knot 1PDG X-ray crystallography<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_397.html (1 of 2) [4/5/2004 5:44:05 PM]<br />

Page 397


Document<br />

Transforming growth factor beta2<br />

(TGFB2)<br />

Transforming<br />

growth factor<br />

alpha<br />

(TGFA)<br />

Tumour<br />

necrosis<br />

factor<br />

(TNF)<br />

Macrophage inflammatory protein<br />

1-beta<br />

Beta cysteine knot 1TFG, 2TGI X-ray crystallography<br />

Beta EGF-like 2TGF NMR<br />

Beta sandwich 1TNF X-ray crystallography<br />

Alpha/Beta 1HUM NMR<br />

Interleukin 8 (IL-8) Alpha/Beta 1IL8 NMR<br />

Melanoma growth stimulating<br />

activator (MGSA)<br />

Alpha/Beta 1MGG NMR<br />

Platelet factor 4 Alpha/Beta 1RHP X-ray crystallography<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_397.html (2 of 2) [4/5/2004 5:44:05 PM]


Document<br />

Page 398<br />

with growth hormone, prolactin receptor (PRLR), and tumor necrosis factor with a ligand. In the case of<br />

the human growth hormone-receptor complex and tumor necrosis factor-receptor complex the<br />

interaction between the ligand and the receptor is intricate and takes place over a large area. It is likely<br />

that this will be a characteristic feature of all cytokine-receptor complexes.<br />

B. Cytokine-Based Therapy<br />

Certain disease states can occur due to a disregulation of the cytokine network. This can lead to chronic<br />

stimulation of the immune and inflammatory response and ultimately to disease. Many cytokines have<br />

been implicated in autoimmune diseases like myasthenia gravis, insulin-dependent diabetes,<br />

atherosclerosis, systemic lupus erythematosus, and rheumatoid arthritis [1,7,9]. The intimate relationship<br />

between cytokines and the pathology of disease development and progress can be exploited to provide<br />

therapeutic benefit. By subtly manipulating the cytokine communication network one can modulate the<br />

disease process. Understanding of the functional roles of cytokines that mediate the communication<br />

between them and the factors involved in the immune system's cell-cell interactions forms the<br />

foundation for cytokine therapies. Such therapeutic agents are now a possibility due to modern<br />

techniques such as molecular biology, structural biology, high-power computation, combinatorial<br />

chemistry, and functional screening. The findings from biological techniques in conjunction with the<br />

structural insights as obtained through protein crystallography and nuclear magnetic resonance form the<br />

foundation for rational drug design [10–13]. Discovery of IL-1-<strong>based</strong> therapeutic agents have led to<br />

promising applications in the treatment of the above mentioned diseases. <strong>Design</strong> of synthetic adjuvants,<br />

for exogenous immunomodulation, is also an important factor to be considered in the field of vaccine<br />

development. In this review we will discuss the available literature on interleukin-1 and the recent<br />

attempts toward the design of immunomodulators.<br />

II. The Interleukin-1 Family<br />

The three molecules of the IL-1 family, interleukin-1α (IL-1α), interleukin-1β (IL-1β), and interleukin-1<br />

receptor antagonist (IL-1Ra) map to the long arm of chromosome two in the human genome. It appears<br />

that the family arose via a gene duplication event some 350 million years ago, and the molecules possess<br />

between 27.5 and 36% sequence identity with each other (Table 2) [1,14,15]. In addition, the genes for<br />

the two IL-1 receptors IL-1R1 and IL-1RII [16,17], and an IL-1R accessory protein (IL-1RacP), which<br />

binds to the IL-1, IL-1 receptor complex [18], have been identified. Together, these molecules via their<br />

differential activity serve primarily to modulate the host defense mechanism.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_398.html [4/5/2004 5:44:07 PM]


Document<br />

Table 2 RMS Distance of Aligned Cα Moieties of the Three Interleukin-1 <strong>Structure</strong>s<br />

IL-1β IL-1α IL-1Ra<br />

IL-1β 0.0 1.8 Å (36.3%) 1.42 Å (30.7%)<br />

IL-1α 1.8 Å (36.3%) 0.0 1.8 Å (27.5%)<br />

IL-1Ra 1.42 Å (30.7%) 1.8 Å (27.5%) 0.0<br />

Number in parentheses referred to sequence identities between the aligned structures.<br />

A. Expression and Processing of IL-1<br />

Page 399<br />

All three IL-1 molecules, IL-1α, IL-1β, and IL-1Ra are synthesized as 31 kDa precursor molecules<br />

produced primarily <strong>by</strong> mononuclear phagocytes. These precursor proteins can be subsequently<br />

processed to mature 17 kDa molecules. The means where<strong>by</strong> this is achieved appears to be different for<br />

each type of molecule and may help to explain the roles of the three. Both agonist molecules IL-1α and<br />

IL-1β lack a classical hydrophobic leader sequence and thus must be processed in an alternative way.<br />

The two can be separated <strong>by</strong> their biological activity in the precursor form, IL-1α being active in both<br />

precursor and mature forms while IL-1β produces a biological response only in its processed form. In<br />

addition, a specific enzyme, Interleukin-1 beta converting enzyme (ICE), cleaves IL-1β to its mature<br />

form. This enzyme is a cysteine protease whose only known substrate is proIL-1β, which it cleaves at<br />

Ala 117 [19,20]. The means where<strong>by</strong> pro IL-1α is processed is still largely a mystery although it has<br />

been postulated that a calpain or related protein may perform the task.<br />

Naturally occurring IL-1Ra is a 22 kDa glycosylated protein [14,15,21– 24] that possesses a signal<br />

sequence. It is likely that it is processed in the conventional manner as can be concluded <strong>by</strong> the<br />

glycosylation states of the molecule not present in the agonists [25]. The LPS-stimulated human blood<br />

monocytes initially express the gene for IL-1Ra [25]. An alternatively spliced form of IL-1Ra also exists<br />

(intracellular IL-1Ra), which remains inside the cell presumably to block intracellular IL-1 action [26].<br />

Both soluble and intracellular forms of IL-1Ra block IL-1R but do not trigger any biological response.<br />

B. Receptors and Responses<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_399.html (1 of 2) [4/5/2004 5:44:09 PM]


Document<br />

All effects of the IL-1 family are exerted via their receptors, which are present on a variety of different<br />

cell types. There are two known IL-1 receptors. The first, type I (IL-1RI) is an 80-kDa formation found<br />

mainly on T cells and fibroblasts; the second, type II (IL-1RII) is present on B cells, monocytes,<br />

neutrophils, and heptoma cells and is approximately 60 kDa in size. From sequence analysis these<br />

molecules are believed to belong to the Ig-superfamily, the extracellular portion of the receptors<br />

consisting of three immunoglobulin-like domains of approximately 100 residues each [27,28]. The IL-<br />

1RI receptor<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_399.html (2 of 2) [4/5/2004 5:44:09 PM]


Document<br />

Figure 1<br />

Schematic representations of the members of the IL-1 family.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_400.html [4/5/2004 5:44:15 PM]<br />

Page 400


Document<br />

has a single membrane-spanning region and a cytoplasmic region of 213 amino acids. The IL-1RII<br />

receptor has a single membrane-spanning region and a cytoplasmic region of 29 amino acids.<br />

Page 401<br />

All three IL-1 molecules bind with a high affinity to the receptor IL-1RI. However only IL-1α and IL-<br />

1β produce a detectable response [29]. The IL-1Ra molecule completely blocks the binding of the<br />

former two molecules without inducing any signal transduction events. The second IL-1 receptor, IL-<br />

1RII, similarly binds IL-1 tightly but does not produce any signal [30]. The soluble portion of IL-1RI<br />

binds to IL-1Ra, IL-1α, and IL-1β in decreasing order of affinity whereas the soluble portion of IL-1RII<br />

binds IL-1β, proIL-1β, IL-1α, and IL-1Ra, in that order. It therefore appears that both IL-1Ra and IL-<br />

1RII take part in modulation of the agonist molecules, an argument supported <strong>by</strong> the finding that soluble<br />

IL-1RII binds IL-1β with a similar affinity to the cell-associated receptor; whereas, the affinity of IL-<br />

1Ra to such a soluble receptor is some 2000 times less [31,32,33]. Greenfeder, et al., [18] have<br />

identified a new molecule, the IL-1 accessory protein IL-1RacP. The complex of IL-1 and IL-1 receptor<br />

binds to IL-1RacP and together they induce the signal [18]. Greenfeder also observed that antibodies to<br />

IL-1RI and to IL-1RacP block IL-1 binding and activity. From sequence analysis it appears that the<br />

cytoplasmic domains of IL-1R and IL-1RacP contain the same amino acid domains commonly found in<br />

the members of the GTPase family of proteins [34]. It has been proposed that such a complexation may<br />

lead to a closer proximity of these cytoplasmic domains and thus facilitate signal transduction. A scheme<br />

showing functional relationships between molecules that make up the IL-1 system—IL-1α, IL-1β, IL-<br />

1Ra, IL-1RI, IL-1RII, and related receptors—is shown in Figure 1.<br />

C. Autoantibodies of IL-1<br />

In addition to these molecules, naturally occurring neutralizing autoantibodies of IgG type to IL-1α have<br />

been identified. These have been detected in serum isolated from human donors. [35,36]. These<br />

antibodies bind to both proIL-1α and 17-kDa IL-1α [37] and completely prevent the binding of IL-1α to<br />

type-I cell surface receptors [38]. Patients with autoimmune diseases have higher populations of these<br />

antibodies [39].<br />

III. Three-Dimensional Structural Information<br />

In an effort to understand the natural and interactions of the IL-1 family, various laboratories have<br />

undertaken the task of elucidating the three-dimensional structure of the molecules. To the present this<br />

has resulted in the structures of all three members of the IL-1 family being solved independently <strong>by</strong> both<br />

x-ray<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_401.html [4/5/2004 5:44:16 PM]


Document<br />

Page 402<br />

crystallography and heteronuclear NMR spectroscopy [40–49]. In addition to these, the structure of IL-<br />

1β converting enzyme has been determined [50]. The molecule is an oligomeric cysteine protease<br />

consisting of 10- and 20-kDa subunits. To date proIL-1β is its only physiological substrate. As<br />

processing is essential for IL-1β extracellular transport and activity, inhibitors of ICE could provide a<br />

method to block the production of the active form of IL-1β. Although the elucidation of these structures<br />

have been important in determining various differences among the individual proteins, perhaps a more<br />

important result would be the elucidation of the structures of the two interleukin-1 receptor molecules,<br />

either individually or in complex with their substrates. This information could provide a means to draw<br />

together the present structural and biochemical knowledge into a coherent picture of the agonistic<br />

activity of IL-1α and IL-1β as opposed to the antagonist effect of IL-1Ra. The structures should also<br />

provide a foundation upon which structure-<strong>based</strong> drug design could proceed. To date no crystals or<br />

structural reports for the receptors have been published.<br />

A. The β-Trefoil Fold<br />

Structurally IL-1 exhibits a unique fold, known as a β-trefoil fold. Each IL-1 molecule show the<br />

characteristic β-trefoil fold and small deviations of the back-bone despite the relatively low sequence<br />

identity (Table 2). Figure 2 shows the structural alignment of the IL-1 molecules. The β-trefoil fold was<br />

first observed in Kunitz-type soybean trypsin inhibitor [51]. The fold has since been identified numerous<br />

times in the Kunitz family of protease inhibitors, the interleukin-1 system molecules, and the acidic and<br />

basic fibroblast growth factors [52]. Although these proteins have diverse biological function and low<br />

sequence identity to each other, all appear to have a common structural core. Each one, however, binds<br />

to its specific receptor with high affinity. The present known examples of proteins with such a fold are<br />

composed of between 125 and 170 amino acids. The overall fold itself is an antiparallel β barrel<br />

consisting of six two-stranded hairpins. Three of these form a barrel structure (strands denoted β1, β4,<br />

β5, β8, β9, and β12) while the other three are in a triangular array that caps the barrel. The arrangement<br />

of these moieties is such as to give the molecule a pseudo-three-fold axis [52]. Each fragment<br />

contributes one pair of antiparallel beta strands to the barrel and one pair to the cap of the barrel, thus<br />

forming a so called “open” and “closed” end to the structure [42].<br />

B. Three-Dimensional <strong>Structure</strong> of IL-1β<br />

The structure of IL-1β was the first of the IL-1 family to be solved and has been solved independently<br />

five times: four times <strong>by</strong> x-ray crystallography [40–<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_402.html [4/5/2004 5:44:18 PM]


Document<br />

Figure 2<br />

Structural alignment of IL-1α, IL-1β, and IL-1Ra. Residue numbering is taken from IL-1β. Residues bordered in black are conserved<br />

over the three molecules while those in gray constitute a conservative substitution. Arrows indicate sheet region, and the cylinders<br />

indicate helix. Figure produced <strong>by</strong> Stamp [109] and Alscript [110].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_403.html [4/5/2004 5:44:25 PM]<br />

Page 403


Document<br />

Figure3<br />

(a) Stereo diagram of the secondary structural elements of IL-1β. Produced <strong>by</strong> Molscript [106].<br />

(b) Stereo plot of all the atoms of IL-1β viewed parallel to the axis of the barrel.<br />

Page 404<br />

42,44], and once <strong>by</strong> NMR [43]. The four crystal structures, each to 2.0 Å, were all solved in the same<br />

space group, P4 3, each structure being in relatively good agreement with the others [53]. As pointed out<br />

previously, the molecule adopts a β-trefoil fold with about 65% of the molecule in beta sheet and 35% in<br />

random coil/turn (Figure 3a). The IL-1β molecule resembles a conical barrel with a shallow open face<br />

on one end and a closed face on the other. The length of the long tubular core of the molecule is about<br />

23 Å. Twelve antiparallel β<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_404.html (1 of 2) [4/5/2004 5:44:35 PM]


Document<br />

Page 405<br />

strands constitute the secondary structural elements of IL-1β. Three pairs of β strands (one pair from<br />

each of the fragments) form the six-stranded barrel and the other three pairs cover one end of the barrel,<br />

referred to as the “closed end”. The amino and carboxy termini are close to each other at the “open end”<br />

of the barrel (Figure 3b). The overall structure of the molecule consists of three similar fragments (F1,<br />

F2, F3) related <strong>by</strong> a pseudo-3-fold symmetry, with each fragment having a βββLβ motif. Residues 1 to<br />

52 (F1), 53 to 107 (F2), and 108 to 153 (F3) form these fragments. These βββLβ motifs can be<br />

superposed to display structural similarities.<br />

The core of the molecule consists entirely of hydrophobic residues, two-thirds of which are leucines and<br />

phenylalanines—a feature that seems to be essential to maintain the structural integrity of the molecule.<br />

Most of the aromatic groups have their planes aligned along the barrel axis and both ends of the barrel<br />

have concentrations of exposed polar residues that may be involved in binding interactions with the<br />

receptor (see below). No alpha-helical structure is observed in IL-1β although one short region of 3 10<br />

helix is observed between residues Gln34 and Gln38. Two of the β hairpins are in the open end and<br />

three are at the closed end. Residues 18 to 28, 69 to 82, and 122 to 135 form the three β hairpins at the<br />

closed end where three strands (residues 24–28, 78–82, and 129–135) are in close proximity and form<br />

three sides of a triangle, covering this end of the barrel.<br />

C. Open End of the Barrel<br />

Analysis of the surface of IL-1 reveals an epitope consisting of many polar residues widely spaced,<br />

approximately in an annular fashion, around the open end of the barrel (Figure 3b). It has a broad<br />

surface area containing many polar residues [42]. Residues from six β-turns [type I, β turn 1 (T1=11-<br />

17); type III, β turn 3 (T3=33-36); type I', β turn 6 (T6=52-55); type I, β turn 7 (T7=62-65); type I, β<br />

turn 9 (T9=86-89); type I', β turn 10 (T10=106-109)] are present in this open end. There is a β bulge<br />

between strands β4 and β5 in one of the hairpins; this bulge may play a key role in the formation of the<br />

putative binding surface. The β bulge in IL-1 is of the “wide” type stabilized <strong>by</strong> multiple interactions,<br />

whose conformation is such as to place the side chains of polar residues (Glu51, Asn53, Asp54) in the<br />

proposed binding surface, fanning out in the direction of the open end of the barrel. Multiple interactions<br />

serve to stabilize the strained conformation of the loop, between residue 86 and 94, presenting the side<br />

chains of Asp86, Lys88, Asn89, and Lys93 at the open end of the barrel. In well-defined residues having<br />

abnormal φ, ψ values, strained conformations often may be related to functional properties and are<br />

observed either in the active site or in the region responsible for its activity. The reason for having such<br />

strains in the loop at this open end may be to place these charged side chains in the putative receptor<br />

binding surface. There is a short 3 10-helix in the region between β3 and<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_405.html [4/5/2004 5:44:42 PM]


Document<br />

Page 406<br />

β4 (residue Gly33 and Gln39). By such a conformation, residues Gln32, Gln34, and Glu37 have their<br />

side chains in the proposed binding surface. A β turn between 62 and 65 is folded in such a way that the<br />

charged groups of the residues Lys63, Glu64, and Lys65 present themselves on the proposed binding<br />

surface. Conformational arrangement via saltbridge scaffolds and other multiple hydrogen-bonding<br />

interactions stabilize the residue 102 to 113 loop, and orient the side chains of Lys103, Glu105, Asn107,<br />

and Asn108 in the proposed surface at the open end of the barrel. Similar arrangement places the side<br />

chains Arg11 and Gln15 in the proposed epitope. The amino and carboxy termini form a part of the<br />

proposed epitope, contributing the polar side chains of Arg4, Ser5, Asn7, Gln149, Ser152, and Ser153.<br />

Based on extensive analysis, we found that the structural elements like electrostatic and hydrogenbonding<br />

interactions in the loops between strands allow the polypeptide to adopt a conformation that<br />

enables an unusual concentration of polar and charged groups to be presented at the open end of the<br />

barrel. Such a cluster of charged residues around an area that is almost perpendicular to the barrel axis<br />

forms a hydrophilic patch with which IL-1 might bind to the receptor [42]. The core of the barrel,<br />

though probably not involved in binding, must nevertheless be important to the function of IL-1 because<br />

mutations within it reduce activity but not binding. We hypothesized that binding and cell proliferation<br />

through signal transduction involve separate regions of the IL-1 molecule; the surface polar loops are<br />

required for the binding and the core of the barrel is required for the physiological response.<br />

D. Three-Dimensional <strong>Structure</strong> of IL-1α<br />

The structure of IL-1α has been determined <strong>by</strong> x-ray crystallography [45] to a resolution of 2.7 Å in<br />

space group P2 1. Its general fold is very similar to that of IL-1β, having the same central β barrel along<br />

with the adjoining loops (Figure 4a, b). The overall rms distance between 133 aligned Cα positions of IL-<br />

1β and IL-1α is 1.8 Å (Figure 5). The major difference between the two molecules is an N-terminal<br />

extension of 8 residues beyond the N-terminus of IL-1β. This projection forms a short β strand (residues<br />

6-10), the presence of which positions the N-terminus of IL-1α in an alternative conformation. It has<br />

been postulated that this conformation is responsible for the differences in binding of precursor IL-1α<br />

and β: while precursor and mature IL-1α bind to IL-1RI and elicit response, only processed IL-1β binds<br />

with the receptor. This suggests that the N-terminal region of IL-1 probably plays a role in receptor<br />

binding. Extra residues in the alternate conformation of immature IL-1β serve to inhibit the receptor<br />

binding and there<strong>by</strong> the biological activity. In contrast to IL-1β, IL-1α incorporates two other secondary<br />

structural elements: a short strand (residues 97–99) and about two turns of a 3 10 helix (residues<br />

101–105), neither have been shown to be important in the function of the molecule.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_406.html [4/5/2004 5:44:45 PM]


Document<br />

Figure 4<br />

(a) Stereo diagram of the secondary structural elements of IL-1α. Produced <strong>by</strong> Molscript [106].<br />

(b) Stereo plot of all the atoms of IL-1α viewed parallel to the axis of the barrel.<br />

E. Three-Dimensional <strong>Structure</strong> of IL-1Ra<br />

Page 407<br />

The x-ray structure of IL-1Ra was first reported in 1994 [47] and has since been solved <strong>by</strong> x-ray<br />

crystallography and NMR <strong>by</strong> several different laboratories [46–49]. Again, this molecule possesses the<br />

characteristic trefoil fold (Figure 6a) and similar tertiary structures (Figure 6b). The structure is similar<br />

to that of IL-1α and IL-1β with an rms deviation of 1.8 and 1.42 Å, respectively, when considering their<br />

Cα positions (Figure 7a). Structural superposition of these three molecules shows a common trefoil core<br />

(Figure 7b). In the absence of receptor substrate complexes, the comparison of all three structures<br />

combined with mutational studies (see below) is invaluable in obtaining a model of the regions<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_407.html (1 of 2) [4/5/2004 5:45:02 PM]


Document<br />

Figure5<br />

Superposition of the Cα atoms of IL-1α and β. Thinner line represents IL-1α<br />

and the thicker line is that of IL-1β.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_408.html (1 of 2) [4/5/2004 5:45:11 PM]<br />

Page 408


Document<br />

Figure 6<br />

(a) Stereo diagram of the secondary structural elements of IL-1Ra. Produced<br />

<strong>by</strong> Molscript [106]. (b) Stereo plot of all the atoms of IL-1Ra viewed parallel to the axis<br />

of the barrel.<br />

involved in binding and exertion of biological effect in the IL-1 system. Similarity between the three<br />

molecules has been observed mainly in the β strands. One notable region of<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_409.html (1 of 2) [4/5/2004 5:45:22 PM]<br />

Page 409


Document<br />

difference between IL-1Ra and IL-1 agonists has been observed on the side of the beta barrel on site B<br />

of the receptor binding site (see below). The Cα positions in this region (residues 84–94 in loop β7–β8)<br />

differ <strong>by</strong> 9.1 Å [26] in comparison with IL-1β. This region also contains Asn84, which is the site for at<br />

least two forms of N-linked glycosylation that occur in IL-1Ra in addition to the nonglycosylated form.<br />

Mutagenesis experiments have also<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_409.html (2 of 2) [4/5/2004 5:45:22 PM]


Document<br />

Figure 7<br />

(a) Superposition of the Cα atoms of IL-1α, IL-1β, and IL-1Ra. (b) A<br />

common trefoil core <strong>based</strong> on the structural superposition of all the three IL-1<br />

molecules.<br />

Page 410<br />

identified this region and the surrounding area in IL-1β with a large receptor binding epitope<br />

encompassing the N and C termini of the protein as well as loops β4–β5 and β9–β10. The other<br />

postulated binding epitope (epitope A) is structurally conserved in IL-1Ra. Other differences in structure<br />

not related to any implicated biological structure are an extended 3 10 helix in residues 92–99 in IL-1Ra<br />

as opposed to a type-1 β turn found in IL-1β.<br />

F. Three-Dimensional <strong>Structure</strong> of Interleukin-1 Beta Converting Enzyme (ICE)<br />

As mentioned above, the processing of the different members of the IL-1 family is important to<br />

discovering their function and may also provide clues in the<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_410.html (1 of 2) [4/5/2004 5:45:50 PM]


Document<br />

Page 411<br />

design of inhibitors. The structure of ICE has been solved <strong>by</strong> x-ray crystallography to 2.6 Å in complex<br />

with an acetyl-Tyr-Val-Asp-H tetra-peptide [50]. Figure 8a illustrates the tertiary structure of the<br />

heterodimer with a dimension of about 45 Å × 35 Å × 25 Å. The molecule itself is oligomeric and<br />

contains two subunits, p20 (residues 120–297) and p10 (317–404) of relative molecular weight 20 kDa<br />

and 10 kDa, respectively. The two subunits form an intimately connected heterodimer. The core of ICE<br />

is a six-stranded β sheet containing 5 parallel strands and one antiparallel strand. The core is bounded <strong>by</strong><br />

six alpha helices that lie parallel to the beta sheets. Physiologically ICE occurs as a (p20) 2(p10) 2<br />

tetramer. The molecule is a cysteine protease, which has a unique preference among the mammalian<br />

proteases for cleaving bonds. It has only one known with aspartic acid adjacent and N-terminal to the P 1<br />

scissile bond. It has only one known physiological substrate, proIL-1β, which it cleaves to active IL-1β.<br />

The struc-<br />

Figure 8<br />

(a) Stereo diagrams of the heterodimer structure of Interleukin-1 Converting<br />

Enzyme (ICE). (b) The tetrapeptide inhibitor (Asp-Ala-Val-Tyr) covalently<br />

bound to Cys285 in the active site. Tetrapeptied shown in black, p20 subunit in<br />

dark gray, and p10 subunit in light gray. Produced <strong>by</strong> Molscript [106].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_411.html (1 of 2) [4/5/2004 5:46:08 PM]


Document<br />

Page 412<br />

ture is unlike other cysteine proteases such as papain and appears to have only one known possible<br />

homologue, CED-3 protein [54]. The crystal structure with the tetrapeptide (acetyl-Tyr-Val-Ala-Asp-H)<br />

substrate in the binding pocket (Figure 8b), coupled with the results obtained <strong>by</strong> the mutational<br />

experiments, has provided a detailed understanding of the enzyme mechanism and substrate bonding of<br />

the molecule. Structural analysis has pinpointed a catalytic diad of residues Cys285 and His237<br />

responsible for the catalytic activity of the molecule: Cys285 acts as an active-site nucleophile while the<br />

imidazole ring of His237 participates in destabilizing the cysteine Oγ hydrogen. Although both of above<br />

these residues are contained in the p20 subunit, the p10 subunit is essential for the maintenance of a<br />

binding pocket (and thus the specificity of the molecule), providing binding sites S2 to S4 (residues<br />

338–341) and jointly contributing, with subunit p20, the S1 (Arg179–Arg341) site. The uniqueness and<br />

substrate specificity of ICE make it an ideal target for small molecules to block the production of<br />

mature, active IL-1β. Since the substrate peptide sequence is known (Tyr-Val-Ala-Asp), it has been a<br />

starting point to develop peptidomimetics to inhibit ICE.<br />

IV. Therapeutic Strategies<br />

A. Manipulating the IL-1 System<br />

Since IL-1 is an important mediator of human disease processes, modulating its activity or completely<br />

inhibiting its synthesis may be of therapeutic benefit. Blake and Henderson [7] reviewed and portrayed a<br />

general strategy to interfere with the cytokine at any one of the following stages: induction or initiation<br />

of gene expression–transcription, RNA processing and translation in IL-1 production, folding, release<br />

and secretion of extracellular protein, IL-1 in circulation, IL-1 binding to its cell surface receptors, signal<br />

transduction and resulting activities (Figure 9). All of the above approaches hold promise for the future.<br />

The wealth of structural information on IL-1 combined with extensive site-directed mutagenesis studies<br />

on the molecules have helped to build up a picture of the regions involved in biological function. A<br />

schematic representation of the present-day efforts to design efficient antagonists of IL-1 its shown in<br />

Figure 10.<br />

B. Site-Directed Mutants<br />

Extensive mutagenesis studies have provided information related to the structural integrity, receptor<br />

binding region, and residues that are important for IL-1 function. Site-directed mutagenesis (SDM) can<br />

create a single-site mutant and its receptor binding and bioactivity values can be calculated. The results<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_412.html [4/5/2004 5:46:10 PM]


Document<br />

Figure 9<br />

Points of intervention for IL-1-<strong>based</strong> immunomodulation.<br />

Page 413<br />

obtained from such studies can be imported into the available 3-dimensional structures. A combination<br />

of such structural insights coupled with the bioassay results provide clues leading to IL-1 functional<br />

information. In a previous structural report on IL-1β, we summarized the SDM results and identified a<br />

plausible receptor-binding epitope of interleukin-1 [42]. As mentioned before, electrostatic and<br />

hydrogen-bonding interactions in the loops between strands allow the polypeptide to adopt a<br />

conformation that enables an unusual concentration of polar and charged groups to be presented at the<br />

open end of the barrel. This cluster of charged residues forms an epitope with which IL-1 might bind to<br />

the receptor. Following our proposal, many groups have supported this hypothesis <strong>by</strong> employing<br />

mutagenesis studies. For a list of mutants and their activity results refer to References 42, 56, 59, and 61.<br />

Ju and coworkers are employing SDM to characterize interleukin-1 [55,56,58,61]. Substitution of Lys<br />

for the Asp145 of IL-β (D145K) greatly reduced agonist activity, while retaining 100% binding to the IL-<br />

1RI [56]. Based on the sequence alignment of IL-1β with IL-1Ra, they selected Lys145 of IL-1Ra for<br />

mutagenesis and converted it to an aspartic acid. This mutant analog (IL-1Ra K145D) maintained<br />

receptor binding and gained partial agonist activity [56]. Following this study, Ju and coworkers selected<br />

five other amino acid residues in IL-1Ra for further analysis because the side chains of these residues<br />

appear to be in close proximity to Lys145 in IL-1Ra [61]. Mutations were made at Val18, Thr108,<br />

Cys116, Cys122, and Tyr147, usually <strong>by</strong> a replacement with the corresponding amino acid of IL-1β at<br />

each position. None of these muta-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_413.html [4/5/2004 5:46:20 PM]


Document<br />

Figure 10<br />

Schematic diagram showing the IL-1-<strong>based</strong> therapeutic molecules.<br />

Page 414<br />

tions provided enough information related to the structure-activity relationship. The mutant IL-1Ra<br />

K145D was mutated further to V18S, T108K, C116F, C122S, C122A, Y147T, Y147G, H54P, and a<br />

H54I. Of these, K145D + T108K showed a 2-fold decrease in IL-1RI binding and a 3-fold decrease in<br />

bioactivity compared to the IL-1Ra K145D analog. The K145D + C116F combination resulted in the<br />

complete loss of bioactivity, whereas full receptor-binding activity was maintained. The observation that<br />

receptor-binding activity is preserved indicates that the binding site is not altered that much. Structural<br />

alignment<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_414.html (1 of 2) [4/5/2004 5:46:59 PM]


Document<br />

indicates that Cys116 in IL-1Ra is in a homologous position with Phe117 in IL- 1β. The K145D +<br />

Y147T analog lost all detectabled activity (both binding and bioactivity), whereas the Y147G analog<br />

lost all bioactivity but retained 100% binding. These data suggest that Tyr147 is important for<br />

bioactivity of IL-1Ra K145D.<br />

C. Insertion of β Bulge<br />

Page 415<br />

A region of charged amino acids (Gln48 to Asn53, a β bulge) positioned between β strands 4 and 5 has<br />

been implicated in IL-1β binding to its receptor and its immunostimulatory properties [42, 64–67]. The<br />

β-bulge residues form a protrusion on the edge of the open end of the β barrel. Evidence for this patch of<br />

amino acids involved in the function of this molecule comes from mutagenesis studies in which<br />

deletions or substitutions of residues in this region reduced IL- 1β agonist activity without affecting<br />

receptor binding [62,63]. Simoncsits et al. have shown that deletion of amino acids 52–54 (SND) in IL-<br />

1β reduces IL-1RI binding <strong>by</strong> 10 fold and biological activity <strong>by</strong> 1000 fold [63]. Also, studies indicate<br />

that a synthetic peptide derived from IL-1β (VQGEESNDK), which contains these six β bulge amino<br />

acids, has immunostimulatory but no inflammatory effects normally associated with IL-1 [64–66]. The<br />

insertion of VQGEESNDK into recombinant human ferritin H chain and recombinant flagellin from<br />

Salmonella muenchen increased the immunogenicity of these antigens in mice [67].<br />

Greenfeder et al. inserted this β-bulge region into IL-1Ra K145D either after Ile51 of after Pro53 [61].<br />

The insertion of the β bulge (QGEESN) after either position 51 or 53 of IL-1Ra K145D resulted in<br />

analogs that retained full IL-1RI binding and increased bioactivity <strong>by</strong> 3–4 fold. Aslo they tried to obtain<br />

the analogs of the QGEESN insertion in the absence of the K145D mutation either after amino acid 51<br />

or 53 of IL-1Ra. None of the plasmid clones with the insertion at position 51 of IL-1Ra produced the<br />

appropriate protein, whereas they were able to isolate clones with the insertion at position 53. Based on<br />

these results, Greenfeder et al. suggest that in the first case the insertion interfered in the proper folding<br />

of the protein, whereas the second mutant folded properly and exhibited only 10 to 20% of the IL-1RI<br />

binding activity. The mutants with K145D + QGEESN insertion after Ile51 of IL-1Ra showed an<br />

increase in bioactivity in the range of 3–8 fold. The triple mutant IL-1Ra K145D/H54P/QGEESN<br />

showed a higher bioactivity, and <strong>based</strong> on their mutagenesis study, Greenfeder et al. suggest that this<br />

increase in activity may be due to the introduction of Pro and not the removal of His at position 54 in the<br />

K145D mutant of IL-1Ra. The cumulative effects of these three mutations are also interesting since their<br />

positions on the IL-1Ra protein appear to be spatially separated. The residues Ile51 and His54 are<br />

located on the open face of the β barrel of IL-1Ra, whereas Lys145 is located away from the open barrel<br />

end. In IL-1β, the same relative<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_415.html [4/5/2004 5:47:38 PM]


Document<br />

Page 416<br />

positions of the β bulge and Asp145 are observed [42] with the two regions separated <strong>by</strong> the known IL-<br />

1RI binding site.<br />

We have compiled the available site-directed mutants of IL-1β and sorted them according to the<br />

following four categories: (1) mutants that shows significantly higher agonistic activity, A1T, P2M,<br />

S5R, N89G, K92R, and K93R; (2) mutants that show significantly higher antagonistic activity, R4E,<br />

C8S, T9G, T9Q, T9E, L10T,C,S,A, C71X, R11G, K93M, M95R, E96Q, K103Q, and D145K;, (3)<br />

mutants that show significantly higher binding, A1T + P2M, S5R, T9L, T9W, P87S, P87H, K88V,G,L,<br />

N89R, E96Q, K103Q, G, C and M148A; and (4) mutants that show significantly lower binding,<br />

R4A,K,D, L6A, T9E, L10N,T,C,S,A, H30R, M44S, F46D, A, 156A, V58A, K92E, K93L,A,F,S,E,Q,L,<br />

K103S, and E105S,K. These four types of mutants are shown in Figure 11a–d.<br />

Taken together, this information supports our earlier proposal of the receptor-binding epitope. This was<br />

further characterized as functional sites A and B of interleukin-1. The first site, Area A is structurally<br />

conserved in all three molecules and contains residues Arg11, His30, and Asp145 in IL-1β Asn17,<br />

Ala36, and Asp147 in IL-1α; and Trp16, Tyr34, and Lys145 in IL-1Ra. Present in both active IL-1<br />

molecules, Asp145 has been recognized as an important residue in IL-1 binding. Area B has also been<br />

identified in both IL-1α and IL-1β it contains a large hydrophilic ridge around solvent-accessible<br />

hydrophobic residues. In-IL-1β, this region contains the 7-residue hydrophilic ridge (Arg4, Gln48,<br />

Glu51, Asn53, Lys93, Glu105, Asn108) around 5 hydrophobic solvent-accessible residues (Leu6,<br />

Val47, Ile56, Leu110, Val151). Figure 12 shows a surface presentation of the proposed receptor-binding<br />

epitope. In IL-1α, the 7-residue hydrophilic ridge has been identified with residues (Arg12, Ile14,<br />

Asp60, Asp61, Ile64, Lys96, and Trp109, which when mutated resulted in significant loss of binding to<br />

the receptor. Area B is structurally conserved in IL-1β and in IL-1α, but is lacking in IL-1Ra. For this<br />

reason Area A has been proposed as the binding region while Area B has been proposed as the<br />

triggering region.<br />

D. Peptide Fragments<br />

Peptide-<strong>based</strong> IL-1 antagonists have been derived from the primary sequence of IL-1α or IL-1β.<br />

Stepwise synthesis of a series of peptides from amino-terminal to carboxy-terminal regions did not<br />

provide any satisfactory results [68]. Polypeptide fragments of IL-1β, termed somnogeneic peptides,<br />

induced sleep in mammals [69], 61.5% NREMS at a dose of 20 ng. The numbering of these peptides in<br />

proIL-1 are 178–207, 199–225, 208–240 and that of mature IL-1β are 62–91, 83–109, and 92–124 (seq<br />

92–124=KKKMEKRFVFNKIENNKLEFESAQFPNWYIST). Monsanto company has identified a<br />

peptide fragment of IL-1β (41–70) exhibiting inhibitory effects for IL-1β and IL-1β, but not TNF-α, at 5<br />

nM (5 ng/mL). Peptide 56–70 is a<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_416.html [4/5/2004 5:48:12 PM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_417.html (1 of 2) [4/5/2004 5:48:41 PM]<br />

Page 417


Document<br />

Figure 11<br />

Functional residues of IL-1β, identified <strong>by</strong> the<br />

site-directed mutagenesis results. Produced <strong>by</strong><br />

Molscript [106].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_417.html (2 of 2) [4/5/2004 5:48:41 PM]


Document<br />

Figure 12<br />

The open end of IL-1β, shown as a surface. The positions of some of the<br />

residues that have been subjected to site-directed mutagenesis studies have<br />

been marked. Produced using the program GRASP [108].<br />

Page 418<br />

weak agonist. Peptide 47–55 (VQGEESNDK) has been identified as an activator of T cells. It also<br />

stimulates glycosaminoglycan synthesis, excites antitumor activity in vivo, and lacks proinflammatory<br />

and pyrogenic activities [70]. Later a shorter segment (49–53 = GEESN) with higher activity than 47–55<br />

has been identified. Peptides <strong>based</strong> on IL-1α and IL-1β sequences were claimed to induce production of<br />

prostaglandin E2 but actually maintain other biological activities. Few other peptides or peptidecontaining<br />

epitopes were identified <strong>by</strong> using neutralizing antibodies. Labriola-Tompkins and his<br />

colleagues have reported obtaining epitopes for neutralizing antibodies <strong>by</strong> fractionation of a goat<br />

polyclonal antiserum over columns containing individual immobilized synthetic<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_418.html (1 of 2) [4/5/2004 5:49:05 PM]


Document<br />

Figure 13<br />

The identified peptide fragments. Peptide fragments are shown as striped<br />

coil, with residues at the start and end of the peptides are numbered. Produced <strong>by</strong><br />

Molscript [106], modified <strong>by</strong> R. Esneuf.<br />

peptides derived from an IL-1α sequence [58]. Their work resulted in 4 peptide regions with residue<br />

numbers 4–12, 44–63, 64–88, and 89–105. Peptides 47–55, 81–99, 92–124, 121–153 in the 3dimensional<br />

structure of IL-1β are shown in Figure 13.<br />

E. Other Strategies<br />

Page 419<br />

Biochemical and structural knowledge has opened many pathways for the development of novel<br />

therapeutics. Such strategies include inducer blockers, nucleotide intercalators, antisense RNAs, and<br />

other novel molecular mimics. It is known that potent inducers such as lipopolysaccharides, c5a, and<br />

integrins bind to IL-1-producing cells and induce the over expression of IL-1. Such an induction can be<br />

interrupted <strong>by</strong> raising the level of neutralizing monoclonal antibodies against the inducers. Alternatively,<br />

one can design ligands that can bind to the inducer receptors, which leads to the inhibition of the IL-1<br />

synthesis process. Transcription factors bind to specific DNA sequences and stimulate gene<br />

transcription. Controlling such a specific gene transcription can be achieved <strong>by</strong> a number of means.<br />

Intercalators or fragments of the same DNA sequences as those bound <strong>by</strong> the transcription factors can<br />

selectively inhibit transcription. Double-stranded oligonucleotides, having the same consensus sequence,<br />

could complete with the transcription factor for binding to the promoter region. Antisense RNA, which<br />

when introduced into eukaryotic cells induces sequence-specific inhibition of target gene expression,<br />

can be used. The<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_419.html [4/5/2004 5:49:18 PM]


Document<br />

Page 420<br />

antisense strand hybridizes the complementary mRNA to form a double-stranded helix there<strong>by</strong><br />

achieving the inhibition of the gene products. Antisense oligo-deoxynucleotide derivatives have been<br />

shown to inhibit species-specific fibroblast PGE 2 synthesis stimulated <strong>by</strong> IL-1. These approaches<br />

generated much interest and many laboratories are pursuing the design of compounds <strong>based</strong> on antisense<br />

nucleotides, triple-helical transcription inhibitors, aptamers, and other novel nucleic acid derivatives.<br />

V. Neutralizing Soluble IL-1 in Circulation<br />

Once IL-1 is released into the extracellular fluid, the following molecules can be employed to<br />

manipulate its activity: soluble receptors, IL-1Ra, IL-1 neutralizing antibodies, IL-1-specific binding<br />

proteins, high-affinity small molecules (Figure 10).<br />

A. Soluble IL-1 Receptors<br />

The most effective neutralizer of a cytokine is likely to be its receptor. Some viruses have been reported<br />

to use an IL-1 receptor mimic to evade the immune system [72]. Natural shedding of cytokine receptors<br />

is a common occurrence and may form part of a normal homeostatic regulatory system, and there is a<br />

high potential for the use of such soluble forms as therapeutic agents. The binding affinity of the mature<br />

forms of the interleukin-1 molecules and their receptors have been reported [8,31–33,71]. These studies<br />

show that the affinity of both the membrane-bound and soluble forms of human IL-1RI for the mature<br />

forms of human IL-1α, IL-1β, and IL-1Ra are approximately the same. In contrast to IL-1RI, IL-1RII<br />

binds IL-1βpreferentially. Based on the binding-affinity studies one can infer that soluble IL-1RI is a<br />

better inhibitor of IL-1α than IL-1β and soluble IL-1RII is a better inhibitor of IL-1β. Both soluble<br />

receptors, at sufficiently high concentrations, will completely block the binding of both IL-1 forms to<br />

cells. Dower and coworkers [8] have studied the real-time binding of human IL-1α, IL-1β, and IL-1Ra<br />

to human soluble IL-1RI and IL- 1RII. It seems that the binding of IL-1Ra to IL-1RI is essentially<br />

irreversible, whereas its binding to IL-1RII is rapidly reversible. In contrast, IL-1RII binds IL-1β<br />

irreversibly [8,1]. This indicates that the IL-1RII can be used as a high-affinity antagonist of IL-1β.<br />

Dower further suggest in his studies that the soluble receptors have several potential advantages over<br />

anticytokine antibodies due to the fact that they have much higher affinities (100 to 1000 fold) and<br />

should not be recognized <strong>by</strong> the immune system. Even though soluble receptors have high-binding<br />

affinity they are difficult to synthesize in large quantities and they have a low half-life.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_420.html [4/5/2004 5:49:21 PM]


Document<br />

B. Interleukin-1 Receptor Antagonist<br />

Page 421<br />

Over the past 5–10 years a number of inhibitors of IL-1 and TNF have been found in biological fluids<br />

and cell-culture supernatants. Interleukin-1 receptor antagonist was the first protein receptor antagonist<br />

to be described. Recombinant IL-1Ra has been shown to block the activity of IL-1α and IL-1β both in<br />

vitro and in vivo in animal models <strong>by</strong> binding to both type I and type II IL-1 receptors without<br />

demonstrable agonist activity. These effects have been extensively studied. However clinical trails<br />

carries out <strong>by</strong> Synergen on human subjects provided a negative support for the antagonist behavior of IL-<br />

1Ra for sepsis. It may be that the system took an alternative route <strong>by</strong> inducing TNF in excessively high<br />

levels thus achieving signal transduction. Therefore neutralizing TNF in combination with IL-1<br />

inhibition could be an alternate procedure. Even though IL-1Ra exhibits very high binding affinity it is<br />

rather a poor inhibitor of IL-1 action in vivo. It must be present at greater than a 100-fold molar excess<br />

over either agonist form to block action, and large doses are required to block IL-1-mediated effects in<br />

vivo [1]. Burger and Dayer suggest that the simultaneous use of IL-1Ra and IL-1RII might be beneficial,<br />

since this mixture—contrary to the use of IL-1Ra alone—completely abolished the production of<br />

interstitial collagenase in the inflammatory pathway [73].<br />

C. Monoclonal Antibodies<br />

Chimarized or humanized neutralizing monoclonal antibodies for IL-1 can be used as IL-1 can be used<br />

as IL-1 antagonists. Otsuka Pharmaceuticals has developed a monoclonal antibody (IgG1 kappa) against<br />

IL-1β. It can be used in the immunoassay method for the selective detection of human IL-1β and also to<br />

determine the biological activity of IL-1β. Their patent (EP 0-364-778) also covers the use of such an<br />

antibody against IL-1β. when it is abnormally produced in disease states. Another patent of Otsuka<br />

Pharmaceuticals (EP 0- 408-859-A2) relates to an antibody and its application to inflammatory<br />

processes. This antibody is directed to a specific antigen on activated human endothelial cells (1E7/2G7)<br />

and blocks the binding of white blood cells causing inflammatory responses. Based on the antibody<br />

complementarity-determining region low-molecular-weight nonpeptide mimetics can be developed. This<br />

attempt might result in low-molecular-weight, orally active compounds.<br />

D. Low-Molecular-Weight Antagonists<br />

Low-molecular-weight antagonists are attractive due to their low cost and bioavailability. Available<br />

literature indicates that many laboratories are attempting to create immunomodulators of small synthetic<br />

molecules, peptidomimetics, bacterial cell-wall components, macrocycles, corticosteriods, and others<br />

[74].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_421.html [4/5/2004 5:49:24 PM]


Document<br />

Figure 14<br />

2-Dimensional structures of small molecules that exhibit IL-1 modulating<br />

activities. They are tenidap, ciprofloxacin, 3-Deazaadenosine, (SK&F 86002),<br />

E5110, DMARDs (Chloroquine, Auranofin, Sodium aurothiomalate and<br />

Dexamethasone) tiaprofenic acid, dexamethasone, tricyclic-ylidene-acetic<br />

acid and its derivative, Probucol, eicosapentenoic acid + docosahexenoic,<br />

pentoxifylline, Denbufylline, and Romazarit (Ro-31-3948).<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_422.html (1 of 2) [4/5/2004 5:49:57 PM]<br />

Page 422


Document<br />

Page 423<br />

American Home Products has patented molecules ranging from substituted quinoline, piperidine, and<br />

naphthpyridine compounds as immunomodulators. Ciprofloxacin, a quinolone antibiotic, reduced the<br />

extracellular IL-1 activity in human monocytes and delayed the peak production of IL-1α IL-β <strong>by</strong> 24 h<br />

and decreased total IL-1β production, but did not change total IL-1α production [75,76].<br />

Tenidap, an antiarthritic drug, has shown efficacy both in rheumatoid arthritis and osteoarthritis [77,78].<br />

It is a known inhibitor of 5-lipoxygenase and cyclooxygenase (5-LO/CO) and in vitro studies indicate<br />

that it also inhibits the synthesis of mature IL-1 and pro IL-1 [77,79]. Kadin reports analogs of Tenidap<br />

as antiinflammatory agents and analgesics [80]. An antiarthritic molecule, 3- Deazaadenosine, has been<br />

demonstrated to inhibit IL-1 production <strong>by</strong> LPS- stimulated human PBMCs acting at the level of RNA<br />

synthesis and also <strong>by</strong> blocking the effects of IL-1α on EL4 cells and induction of PGE 2 release <strong>by</strong><br />

human fibroblast [81,82]. Another 5-LO/CO inhibitor (SK&F 86002) has been reported to inhibit the<br />

synthesis of IL-1 in human monocytes and human synovial cells in a dose-dependent manner [83,84].<br />

Analogs of this compound<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_423.html [4/5/2004 5:50:16 PM]


Document<br />

Figure 14<br />

(Continued)<br />

have been patented as IL-1 inhibitors [85–87]. In human monocytes, E-5110 is a dual 5-LO/CO<br />

inhibitor found to reduce extra- and intracellular IL-1 activity induced <strong>by</strong> LPS in a dose-dependent<br />

manner. This compound also inhibits the IL-1 generation induced <strong>by</strong> antigen-antibody complexes,<br />

zymosan, and silica particles [88].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_424.html (1 of 2) [4/5/2004 5:50:24 PM]<br />

Page 424


Document<br />

Page 425<br />

Antinflammatory DMARDs such as chloroquine, auranofin, sodium aurothiomalate, and dexamethasone<br />

have been shown to inhibit IL-1 synthesis [89]. Analogs of these compounds have exhibited potent<br />

inhibition of IL-1α- induced cartilage resorption [90]. Elevated collagenase and proteoglycanase levels<br />

caused <strong>by</strong> IL-1 in human cartilage were found to be reduced <strong>by</strong> tiapro-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_425.html (1 of 2) [4/5/2004 5:50:33 PM]


Document<br />

Figure 14<br />

(Continued)<br />

Page 426<br />

fenic acid and dexamethasone [91]. A patent report from the National Institutes of Health describes a<br />

method of treating diseases associated with elevated levels of interleukin-1. Rosenthal, as the inventor of<br />

this patent, describes a method for inhibiting the release of IL-1 from IL-1-producing cells <strong>by</strong><br />

administering a therapeutically effective amount of an aromatic diamidine (WO9115201-A).<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_426.html (1 of 2) [4/5/2004 5:50:41 PM]


Document<br />

Page 427<br />

These aromatic diamidines inhibit IL-1 production and also block IL-6 and TNF. Imidazoline blocks IL-<br />

1 and TNF and is less toxic to the cell with an in vitro LD 50 >>10 -4. Examples of these compounds are<br />

1,5-bis(4-amidophenoxy) pentane (pentamidine), in the form of pentamidine isothionate, and an<br />

imidazoline in the form of 1,5-di(4-imidazolinophenoxy)pentane. Tricycli- cylidene-acetic acid [92] and<br />

its 2-chloro derivative [93] were found to be inhibitors of IL-1 release, claiming clinical improvement in<br />

patients with psoriasis, periodontal disease, and Alzheimer's disease. In vitro this compound blocks the<br />

synthesis of prostaglandins and inhibits the release of IL-1α and IL-1β from human monocytes and<br />

murine macrophages.<br />

Probucol, a hypocholesterolemic drug that possesses antioxidant activity, inhibits the ex vivo release of<br />

IL-1 from LPS-stimulated macrophages of mice pretreated orally with 100 mg/kg/day of this compound<br />

[94,95]. This compound has been shown to inhibit LPS-induced zinc-lowering effect, is cited as direct<br />

evidence for the inhibition of IL-1 release, and may be useful candidate for the treatment of<br />

atherosclerosis [95,96]. An amino-dithiol-one derivative (RP 54745) blocked the proliferative action of<br />

IL-1β on murine thymocytes in vitro and also inhibited the production of IL-1 in mouse peritoneal<br />

macrophages in vitro and in vivo. The compound RP 54745 selectively inhibited the expression of IL-<br />

1α and IL-1β mRNA while TNFα mRNA was unaffected [97, 98].<br />

Administration of a cocktail containing eicosapentenoic acid and docosahexenoic acid to volunteers for<br />

up to 6 weeks, resulted in a significant depression in IL-1β (61%), IL-1α (39%), and TNF (40%)<br />

synthesis. These levels returned to normal after a few weeks [99]. In vitro studies indicate that<br />

Pentoxifylline can block the effects of IL-1 and TNF on neutrophils [100]. It is a phosphodiesterase<br />

(PDE) inhibitor that causes increased capillary blood flow <strong>by</strong> decreasing blood viscocity and is used<br />

clinically in chronic occlusive arterial disease of the limbs with intermittent claudication. Denbufylline,<br />

a closely related xanthine, has been patented as a functional inhibitor of cytokines and exhibits a similar<br />

profile to Pentoxifylline [101]. Romazarit (Ro-31-3948) derived from oxazole and isoxazole propionic<br />

acids has been shown to block IL- 1-induced activation of human fibroblasts in vitro and in animal<br />

models reduces inflammation [102,103,104]. By using a spontaneous autoimmune MRL/lpr mouse<br />

model, a significant efficacy was shown [105]. Two-dimensional structures of some of these molecules<br />

are shown in Figure 14.<br />

Even though the above mentioned small molecules exhibit IL-1 inhibition none of them were discovered<br />

<strong>based</strong> on defined functional or structural aspects. An understanding of the three-dimensional structure of<br />

IL-1s and their receptors, <strong>by</strong> themselves or in complexes, will form a very strong foundation for<br />

structure-<strong>based</strong> design of more specific and potent IL-1-<strong>based</strong> immunomodulators.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_427.html [4/5/2004 5:50:49 PM]


Document<br />

VI. Conclusion<br />

Page 428<br />

The design of novel compounds to inhibit or manipulate the IL-1 system remains a daunting task. At this<br />

time, the design of immunomodulators for the IL-1 system is still in its infancy and has largely been<br />

confined to the use of whole or fragmented proteins or the identification of nonspecific small molecules.<br />

In addition, newer approaches have also been initiated and these include the use of antisense<br />

oligonucleotides, small molecules designed to compete with IL-1's binding to its receptor, ICE<br />

inhibitors, and intracellular signaling inhibitors. All such strategies show promise.<br />

<strong>Structure</strong>-<strong>based</strong> design has not been explicitly used in the design of agonists and antagonists of IL-1. But<br />

as of now we have the structures of IL-1α, IL-1β, and IL-1Ra. A new insight may be forthcoming once<br />

the complex crystallographic structure of one of the interleukin-1 molecules and its corresponding<br />

receptor molecule is available. This structural information, coupled with the anticipated IL-1 + IL-1R<br />

complex structure, will form the foundation for rational design of inhibitors with improved selectivity<br />

for the treatment of various IL-1-mediated diseases.<br />

Acknowledgements<br />

Our special thanks to Professor Russell Doolittle for his encouragement and support, and also to Dr.<br />

Mitch Lewis for providing us with the very high-resolution coordinates of IL-1α. We gratefully<br />

acknowledge San Diego Super Computing Center for their assistance and support in providing valuable<br />

software and high-power computing time. We thank Dr. Donald Kyle for his valuable comments. We<br />

also thank Dr. Per Kraulis for providing us the latest version of MOLSCRIPT, Dr. Anthony Nicholls for<br />

the program GRASP, Dr. Rob Russell for the structural alignment, and Professor Lynn Ten Eyck and<br />

Dr. Jerry Greenberg for their help.<br />

References<br />

1. Dinarello CA. Biological basis for interleukin-1 in disease. Blood 1996; 87- 6:2095–2147.<br />

2. Dinarello CA. Interleukin-1 is produced in response to infection and injury. Rev infect Dis<br />

1984;6:51–56.<br />

3. Oppenheim JJ, Kovacs EJ, Matsushima K, Durum SK. There is more than one interleukin 1. Immunol<br />

Today 1986;7:45–56.<br />

4. Durum KS, Oppenheim JJ, Neta R. Immunophysiologic role of interleukin-1. In: Oppenheim JJ,<br />

Shevach EM, eds. Immunophysiology. New York: Oxford University Press, 1990:210–225.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_428.html [4/9/2004 12:10:04 AM]


Document<br />

5. Dinarello CA. On the Biology of Interleukin-1. Adv Immunology 1989; 44:153– 205.<br />

Page 429<br />

6. Meager A. Cytokines. 1990; Open University Press.<br />

7. Henderson B, Blake S. Therapeutic Advantage of Cytokine Manipulation. TIPS 1992; 13:145–152.<br />

8. Dower SK, Fanslow W, Jacobs C, Waugh S, Sims JE, Widmer MB. Interleukin-1 antagonists.<br />

Therapeutic Immunol 1994; 1:113.<br />

9. Burger DD, J. Inhibitory Cytokines and Cytokine Inhibitors. Neurology 1995; 45(supp 16):s39–43.<br />

10. Abraham DJ. X-ray crystallography and drug design. In: Perun TJ, Propst CL, eds. Computer Aided<br />

<strong>Drug</strong> <strong>Design</strong>. New York: Marcel Dekker, Inc., 1989:93–132.<br />

11. Fesik SW. Approaches to drug design using nuclear magnetic resonance spectroscopy. In: Perun TJ,<br />

Propst CL, eds. Computer-Aided <strong>Drug</strong> <strong>Design</strong>. New York: Marcel Dekker, Inc., 1989:133–184.<br />

12. Veerapandian B. <strong>Structure</strong> aided drug design. In: Wolf M, ed. Buerger's <strong>Drug</strong> Discovery and<br />

Medicinal Chemistry. New York: John Wiley and Sons, 1995:303– 348.<br />

13. Whittle PJ, Blundell TL. Protein <strong>Structure</strong> <strong>based</strong> drug design. Ann Rev Biophy and Biomol<br />

<strong>Structure</strong> 1994; 23:349–375.<br />

14. Carter DB, Deibel MRJ, Dunn CJ, Tomich C-SC, Laborde AL, Slightom JL. Berger AE,<br />

Bienkowski MJ, Sun FF, McEwan RN, Hams PKW, Yem AW, Waszak GA, Chosay JG, Sieu LC,<br />

Hardee MM, Zucher-Neely HA, Reardon IM, Heinnckson RL, Truesdell SE, Shelly JA, Eessalu TE,<br />

Taylor BM, Tracey DE. Purification, cloning, expression, and biological characterization of an<br />

interleukin-1 receptor antagonist protein. Nature 1990; 344:633–638.<br />

15. Eisenberg SP, Evans RJ, Arend WP, Verderber E, Brewer MT, Hannum CH, Thompson RC.<br />

Primary structure and functional expression from complementary DNA of a human interleukin-1<br />

receptor antagonist. Nature, 1990; 343:341–346.<br />

16. Sims JE, March CJ, Cosman D, Widmer MB, MacDonald HR, McMahan CJ, Grubin CE, Wignall<br />

JM, Jackson JL, Call SM, et al. cDNA expression cloning of the IL-1 receptor, a member of the<br />

immunoglobulin superfamily. Science 1988; 241:585.<br />

17. McMahon CJ, Slack JL, Mosley B, Cosman D, Lupton SD, Brunton LL, Grubin CE, Wignall JM,<br />

Jenkins NA, Brannan C1, Copeland NG, Huebner K, Croce CM, Cannizzaro LA, Benjamin D, Dower S,<br />

Spriggs MK, Sims JE. A novel IL-1 receptor cloned from B cells <strong>by</strong> mammalian expression is expressed<br />

in many cell types. EMBO J 1991; 10:2821.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_429.html (1 of 2) [4/9/2004 12:10:07 AM]


Document<br />

18. Greenfeder SA, Nunes P, Kwee L, Labow M, Chizzonite RA, Ju G, Molecular cloning and<br />

characterization of a second subunit of the Interleukin-1 receptor complex. J Biol Chem<br />

1995;270:13757–13765.<br />

19. Thornberry NA, Bull HG, Calaycay JR, Chapman KT, Howard AD, Kostura MJ, Miller DK,<br />

Molineaux SM, Weidner JR, Aunins J, Schmidt JA, Tocci M.A novel heterodimeric cysteine protease is<br />

required for interleukin-1β processing in monocytes. Nature 1992; 356:768.<br />

20. Cerretti DP, Kozlosky CJ, Mosley B, Nelson N, Van Ness K, Greenstreet TA, March CJ, Kronheim<br />

SR, Druck T, Cannizzaro LA, Huebner K, Black RA. Molecular cloning of the IL-1β processing<br />

enzyme. Science 1992; 256:97.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_429.html (2 of 2) [4/9/2004 12:10:07 AM]


Document<br />

21. Seckinger P, Lowenthal JW, Williamson K, Dayeri M, MacDonald HR. A urine inhibitor of<br />

interleukin-1 activity that blocks ligand binding. J Immunol 1987; 139:1546.<br />

22. Hannum CH, Wilcox CJ, Arend WP, Joslin FG, Dripps DJ, Heimdal PL, Armes LG, Sommer A,<br />

Eisenberg SP, Thompson RC. Interleukin-1 receptor antagonist activity of a human interleukin-1<br />

inhibitor. Nature 1990; 343:336.<br />

23. Mazzei GJ, Seckinger PL, Dayer JM, Shaw AR. Purification and characterization of a 26-kDa<br />

competitive inhibitor of interleukin 1. Eur J Immunol, 1990; 20:683.<br />

24. Haskill S, Manin M, VanLe L, Morris J, Peace A, Bigler CF, Jaffe GJ, Sporn SA, Fong S, Arend<br />

WP, Ralph P. cDNA cloning of a novel form of the interleukin-1 receptor antagonist associated with<br />

epithelium. Proc Natl Acad Sci USA 1991; 88:3681.<br />

Page 430<br />

25. Arend WP. Interleukin-1 receptor antagonist. Adv Immunol 1993; 54:167.<br />

26. Muzio M, Polentarutti N, Sironi M, Pli G, DeGioia L, Introna M, Mantovani A, Colotta F.<br />

Characterization of intracellular interleukin-1 receptor antagonist II. Cytokine 1995; 7:632.<br />

27. Kroggel R, Martin M, Pingoud V, Dayer J-M, Resch K. Two Chain structure of the interleukin-1<br />

receptor. FEBS Lett 1988; 229:59.<br />

28. Sims JE, Painter SL, Gow IR. Genomic organization of the type I and type II IL-1 receptors.<br />

Cytokine 1995, 7:483–490.<br />

29. Sims JE, GMA, Slack JL, Alderson MR, Bird TA, Gin JG, Colotta F, Re F, Mantovani A,<br />

Shanebeck K, Grabstein KH Dower SK. Interleukin-1 signaling occurs exclusively via the type I<br />

receptor. Proc Natl Acad Sci USA 1993; 90:6155–6159.<br />

30. Colotta F, DSK, Sims JE, Mantovani A. The type II “decoy” receptor: A novel regulatory pathway<br />

for interleukin-1. Immunol Today 1994; 15:562.<br />

31. Symons JA, Young PA, Duff GW. The soluble interleukin I receptor: Ligand binding properties and<br />

mechanisms of release. Lymphokine Cytokine Res 1993; 12:381.<br />

32. Arend WP, Malyak M, Smith MF, Whisenand TD, Slack JL, Sims JE, Giri JG, Dower SK. Binding<br />

of IL-1α, IL-1β, and IL-1 receptor antagonist <strong>by</strong> soluble IL-1 receptors and levels of soluble IL-1<br />

receptors in synovial fluids. J Immunol 1994; 153:4766.<br />

33. Symons JA, Young PA, Duff GW. Differential release and ligand binding of type II IL-1 receptors.<br />

Cytokine 1994; 6:555.<br />

34. Hopp TP. Evidence from sequence information that the interleukin-I receptor is a transmembrane<br />

GTPase. Protein Sci 1995; 4:1851.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_430.html (1 of 2) [4/9/2004 12:10:13 AM]


Document<br />

35. Bendtzen K, Svenson M, Jonsson V, Hippe E. Autoantibodies to cytokines-Friends or foes?<br />

Immunol Today 1990; 11:167.<br />

36. Svenson M, Hensen MB, Bendtzen K. Distribution and characterization of autoantibodies to<br />

interleukin-1α in normal human sera. Scand J Immunol 1990; 32:695.<br />

37. Svenson M, Hensen MB, Kayser L, Rasmussen AK, Reimert CM, Bendtzen K. Effects of human<br />

anti-IL-1α autoantibodies on receptor binding and biological activities of IL-1. Cytokine 1992; 4:125.<br />

38. Satoh H, Chizzonite R, Ostrowski C, Ni-Wu G, Kim H, Fayer B, Mae N, Nadeau R, Liberato DJ.<br />

Characterization of antiIL-1α autoantibodies in the sera from healthy humans. Immunopharmacology<br />

1994; 27:107.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_430.html (2 of 2) [4/9/2004 12:10:13 AM]


Document<br />

39. Saurat JH, Schfferli J, Steiger G, Dayer J-M, Didierjean L. Anti-interleukin-1α auto antibodies in<br />

humans. J Allergy Clin Immunol 1991; 88:244.<br />

40. Preistle JP, Schar H-P, Grutter MG. Crystallographic Refinement of Interleukin 1β at 2.0 Å<br />

Resolution. Proc Natl Acad Sci USA 1989; 86:9667–9671.<br />

Page 431<br />

41. Finzel BC, Clancy LL, Holland DR, Muchmore SW, Watenpaugh K, Einspahr HM. Crystal structure<br />

of recombinant human interleukin-1β at 2.0 Å resolution. J Mol Biol, 1989; 209:779–791.<br />

42. Veerapandian B, Gilliland GL, Raag R, Svensson AL, Masui Y, Hirai Y, Poulos TL. Functional<br />

implications of interleukin-1β bases on the three-dimensional structure. Proteins: Struct, Funct Genet<br />

1992; 12:10–23.<br />

43. Clore GM, Wingfield PT, Gronenborn AM. High resolution three dimensional structure of<br />

interleukin-1β in solution <strong>by</strong> three and 4 dimensional nuclear magnetic spectroscopy. Biochemistry<br />

1991; 30:2315–2319.<br />

44. Ohlendorf DHT, A, Weber PC, Wondolski JJ, Salemme FR, Lischwe M, Newton RC. A comparison<br />

of the high resolution structures of human and marine interleukin-1β to be published.<br />

45. Graves BJ, Hatada MH, Hendnckson WA, Miller JK, Madison VS, Satow Y. <strong>Structure</strong> of interleukin-<br />

1α at 2.7 A resolution. Biochemistry 1990; 29:2679– 2684.<br />

46. Stockman BJS, TA, Strakalaitis NA, Brunner DP, Yem AW, Deibel MR Jr. Solution structure of<br />

human interleukin-1 receptor antagonist protein. FEBS Letters 1994; 349:79–83.<br />

47. Vigers GPA, Caffes P, Evans RJ, Thompson RC, Eisenberg SP, Brandhuber BJ. X-ray structure of<br />

interleukin-1 receptor antagonist at 2.0 Å resolution. J of Biol Chem 1994; 269:12874–12879.<br />

48. Schreuder HAR, J-M, Tardif C, Soffientini A, Sarubbi E, Akeson A, Bowlin TL, Yanofsky S,<br />

Barrett RW. Crystal structure of the interleukin-1 receptor antagonist. To be published.<br />

49. Spraggon G, Singh O, Stuart DI, Jones EY. The crystal structure of intact interleukin-1 receptor<br />

antagonist. To be published.<br />

50. Wilson KPB, JF, Thomson JA, Kim EE, Griffith JP, NAvia MA, Murcko MA, Chambers SP,<br />

Aldape RA, Raybuck SA, Livingston DJ. <strong>Structure</strong> and mechanism of interleukin-1β converting<br />

enzyme. Nature 1994; 370:270–275.<br />

51. McLachlan AD. Three-fold structural pattern in the soybean trypsin inhibitor (Kunitz). J Mol Biol<br />

1979; 133:557–563.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_431.html (1 of 2) [4/9/2004 12:10:17 AM]


Document<br />

52. Murzin AG, Lesk AM, Chothia C. b- Trefoil Fold: patterns of structure and sequence in the kunitz<br />

inhibitors interleukins-1β and 1α and fibroblast growth factors. J Mol Biol 1992; 223:531–543.<br />

53. Ohelndorf DH. Accuracy of refined proteins structures II. Comparison of four<br />

independantly refined models of interleukin-1β. Acta Cryst 1994; D50:808–812.<br />

54. Yuan J, Shaam S, Ledoux S, Ellis HM, Horvitz HR. The C. elegans cell death gene ced-3 encodes a<br />

protein similar to mammalian interleukin-1β converting enzyme. Cell 1993; 75:641–652.<br />

55. Labriola-Tompkins E, Chandran C, Kaffka K L, Biondi D, Graves BJ, Hatada M, Madison VS,<br />

Karas J, Klian PL, Ju G. Identification of the discontinuous binding site in human interleukin-1β for the<br />

type I interleukin-1 receptor. Proc Natl Acad Sci USA 1991; 88:11182–11186.<br />

56. Ju G, Labriola-Tompkins E, Campen CA, Benjamin WR, Karas J, Plocinski J, Biondi D, Kaffka KL,<br />

Klian PL, Eisenberg SP, Evans RJ. Conversion of the IL-1<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_431.html (2 of 2) [4/9/2004 12:10:17 AM]


Document<br />

receptor antagonist into an agonist <strong>by</strong> a single amino acid substitution. Proc Natl Acad Sci USA<br />

1991; 88:2658–2662.<br />

Page 432<br />

57. Kawashima H, Yamagishi J, Yamayoshi M, Ohue M, Fukwi T, Kotani H, Yamada M. <strong>Structure</strong>activity<br />

relationships in human interleukin-1α: identification of key residues for expression of biological<br />

activities. Protein Eng 1992; 5–2:171–176.<br />

58. Labriola-Tompkins E, Chandran C, Varnell TA, Madison VS, Ju G. <strong>Structure</strong>-function analysis of<br />

human IL-1α: Identification of residues required for binding to the human type I IL-1 receptor. Protein<br />

Eng 1993; 6–5:535–539.<br />

59. Guinet F, Guitton J, Gault N, Folliard F, Touchet N, Cherel J, Crespo A, Destourbe A, Bertrand P,<br />

Denefle P, Mayaux J, Bousseau A, Duchesne M, Terlain B, Cartwright T. Interleukin-1β specific partial<br />

agonists defined <strong>by</strong> site-directed mutagenesis studies. Eur J Biochem 1993; 211:583–590.<br />

60. Grutter MG, van Oostrum J, Pnestle JP, Edelmann E, Joss U, Feige U, Vosbeck K, Schmitz A.<br />

Protein Eng, 1994; 7:663–671.<br />

61. Greenfeder SA, Varnell T, Powers G, Lombard-Gillooly K, Shuster D, McIntyre KW, Ryan DE,<br />

Leven W, Madison V, Ju G. Insertion of a structural domain of interleukin-1β confers agonist activity to<br />

the IL-1 receptor antagonist. J Biol Chem 1995; 270:22460–22466.<br />

62. Wolfson AJ, Kanaoka M, Lau F, Ringe D, Young P, Lee J, and Blumenthal J. Biochemistry 1993;<br />

32:5327–5331.<br />

63. Simoncsits A, Bnstulf J, Tjornhammar ML, Cserzo M, Pongor S, Rybakina E, Gatti S, Bartfai T.<br />

Cytokine 1994; 6:206–214.<br />

64. Antoni G, Presentini R, Penn F, Tagliabue A, Ghiara P, Censini S, Volpini G, Villa L, Boraschu D. J<br />

Immunol 1986; 137:3201.<br />

65. Boraschi D, Nencioni L, Villa L, Censini S, Bossu P, Ghiara P, Presentini R, Perin F, Frasca D,<br />

Dona G, Forni G, Musso T, Giovarelli M, Ghezzi R, Bertini R, Besedovsky H, Del Rey A, Sipe J,<br />

Anotoni G, Silvestn S, Tagliabue A. J Exp Med 1988; 168:675–686.<br />

66. Frasca D, Boraschi D, Baschien S, Bossu P, Tagliabue A, Adorini L, Dona GJ Immunol 1988;<br />

141:2651–2655.<br />

67. Beckers W, Villa L, Gonfloni S, Castagnoli L, Newton SMC, Cesareni, Ghiara P. J Immunol 1993;<br />

151:1757–1764.<br />

68. Joss UR, Schmidli I, Vosbeck K. Mapping the receptor binding domain of interleukin-1β <strong>by</strong> means<br />

of binding studies using overlapping fragments: Why did it fail? J Recept Res 1991; 11:275–282.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_432.html (1 of 2) [4/9/2004 12:10:20 AM]


Document<br />

69. Obal F, Opp M, Cady AB, Johannsen L, Postlethwaite AE, Poppleton HM, Seyer JM, Krueger JM.<br />

Interleukin-1β and an interleukin-1β fragment are somnogenic. Am J Physiol, 1990; 259:R439.<br />

70. Antoni G, Presentini R, Perin F, Tagliabue A, Ghiara P, Censini S, Volpini G, Villa L, Boraschi D.<br />

Peptide analogues of IL-1 and biochemical assay of their binding to its receptors. J Immunol 1986;<br />

137:3201–3204.<br />

71. Slack J, McMahan CJ, Waugh S, Schooley K, Spriggs MK, Sims JE, Dower SK. Independent<br />

binding of interleukin-1α and interleukin-1β to type I and type II interleukin-1 receptors. J Biol Chem<br />

1993; 268:2513.<br />

72. Alcami AS, Smith GL. A soluable receptor for interleukin-1β encoded <strong>by</strong> Vaccinia virus: A novel<br />

mechanism of virus modulation of the host response to infection. Cell 1992; 71:153–167.<br />

73. Burger D, Dayer JM. Inhibitory cytokines and cytokine inhibitors. Neurology 1995; 45 (suppl<br />

6):S39–S43.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_432.html (2 of 2) [4/9/2004 12:10:20 AM]


Document<br />

74. Bender PE, Lee JC., eds. Pharmacological modulation of interleukin-1 (1989). Annual reports in<br />

medicinal Chemistry-25. Johns. Section IV-Metabolic diseases and endocrine function. Chapter 20.<br />

75. Roche Y, Fay M, Gougerot-Pocidalo MA. Antimicrob Chemother 1988; 21:597.<br />

Page 433<br />

76. Bailly S, Mahe Y, Ferrua B, Fay M, Wakasugi H, Tursz T, Gougerot-Pocidalo MA. Cytokine 1989;<br />

1:303.<br />

77. Otterness IG. Abstracts, 3rd Interscience World Conference on Inflammation. Monte-Carlo,<br />

1989:371.<br />

78. Otterness IG, Bliven ML, Downs JT, Manson DC. Arthritis Rheum. Abstracts, 1988; 31–4:S90,<br />

C55.<br />

79. McDonald B, Loose L, Rosenwasser LJ. Arthritis Rheum. Abstracts, 1988; 31– 4:S52, A88.<br />

81. Jurgensen CH, Wolberg G, Zimmerman TP. Agents Actions 1989; 27:398.<br />

80. Kadin, U.S. Patent 4,730,004 (1988).<br />

82. Schmidt JA, Bomford R, Gao XM, Rhodes J. Int J Immunopharmacol 1990; 12:89.<br />

83. Lee JC, Griswold DE, Votta B, Hanna N. Int J Immunopharmacol, 1988; 10:835.<br />

84. Lee JC, Votta B, Griswold DE, Hanna N. Agents Actions 1989; 27:280.<br />

85. Bender PE, Griswold DE, Hanna N, Lee JC. 1988; U.S. Patent 4,794,114.<br />

86. Bender PE, Griswold DE, Hanna N, Lee JC. 1988; U.S. Patent 4,780,470.<br />

87. Bender PE, Griswold DE, Hanna N, Lee JC. 1988; U.S. Patent 4,778,806.<br />

88. Shirota H, Goto M, Hashida R, Yamatsu I, Katayama D. Agents Actions 1989; 27:322.<br />

89. Goodacre J, Carson WD. Allison in Immunopathogenetic Mechanisms of Arthritis. Boston: MTP<br />

Press, 1988:211.<br />

90. Rainford KD. J Pharm Pharmacology 1989; 41:112.<br />

91. Shinmei M, Kikuchi T, Masuda K, Shimomura Y. <strong>Drug</strong>s, 1988; 35 (Suppl. 1):33.<br />

92. Seibel MJ, Bruckle W, Respondek M, Beveridge T. Schnyder J, Muller W, Rheumatol Z. 1989;<br />

48:147.<br />

93. Bollinger P, Gubler HU, Schnyder J. 1989; Derwent 89–138880–B2; DE 38 36 329 Al.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_433.html (1 of 3) [4/9/2004 12:10:25 AM]


Document<br />

94. Ku G, Doherty N. 1988; Derwent 88-314770; AU-A-13160/88.<br />

95. Ku G, Doherty NS, Wobs JA, Jackson RL. A, J Cardiol 1988; 62:778.<br />

96. Marx JL. Science 1988; 239:257.<br />

97. Folliard<br />

F, Terlain B.<br />

Abstracts, 3rd<br />

Inter-science<br />

World<br />

Conference<br />

on<br />

Inflammation.<br />

MonteCarlo;<br />

1989; 415.<br />

99. Endres S, Ghorbani R, Keliey VE, Georgilis K, Lonnemann G, van der Meer JWM, Cannon JG,<br />

Rogers TS, Klempner MS, Weber PC, Schaefer EJ, Woldf SM, Dinarello CA. N Engl J Med 1989;<br />

320:265.<br />

98.<br />

Folliard<br />

F,<br />

Bousseau<br />

A,<br />

Terlain<br />

B.<br />

Cytokine<br />

1989;<br />

1:108.<br />

100. Sullivan GW, Carper HT, Novick Jr. WJ, Mandell GL. Infect Immun 1988; 56:1722.<br />

101. Mandell GL, Sullivan GW, Novick Jr. WJ, 1989; Derwent 89–191 551–B2; WO 89 05145.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_433.html (2 of 3) [4/9/2004 12:10:25 AM]


Document<br />

102.<br />

Machin<br />

PJ,<br />

Osbond<br />

JM, Sqix<br />

CR,<br />

Smithen<br />

CE, Tong<br />

BP. U.S.<br />

Patent<br />

4,774,253<br />

(1988).<br />

103. Bloxham DP, Bradshaw D, Cashin CH, Dodge BB, Lewis EJ, Westmacott D, Barber W.E, Machin<br />

PJ, Osbond JM, Self CR, Smithen CE, Tong BP. Brit J Rheumatol 1987; 26 (Suppl. 2):2.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_433.html (3 of 3) [4/9/2004 12:10:25 AM]


Document<br />

104. Bradshaw D, Dodge BB, Franz PH, Lee SC, Wilson SE. Abstracts, 3rd Interscience World<br />

Conference on Inflammation. Monte-Carlo, 1989, 183.<br />

105. Sedgwick AD. Abstracts, 3rd Interscience World Conference on Inflammation. Monte-Carlo;<br />

1989:183.<br />

106. Kraulis PJ, MOLSCRIPT: a program to produce both detailed and schematic plots of protein<br />

<strong>Structure</strong>s. J Appl Cryst 1991; 24:946–950.<br />

107. Merrit EM, M RASTER 3D version 2.0: a program for photorealistic molecular graphics. Acta<br />

Crystallogr 1994; D50:869–873.<br />

Page 434<br />

108. Nicholls A, Sharp KA, Honig B. Protein folding and association: insights from the interfacial and<br />

thermodynamic properties of hydrocarbons. Proteins 1991; 11:281–296.<br />

109. Russell RB, Barton CJ, Proteins, 1992; 14:309–323.<br />

110. Barton CJ, Protein Engineering, 1989; 6:37–40.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_434.html [4/9/2004 12:10:26 AM]


Document<br />

17<br />

<strong>Structure</strong> and Functional Studies of Interferon: A Solid Foundation for<br />

Rational <strong>Drug</strong> <strong>Design</strong><br />

Michael A. Jarpe<br />

Cambridge NeuroScience, Inc., Cambridge, Massachusetts<br />

Carol H. Pontzer<br />

University of Maryland, College Park, Maryland<br />

Brian E. Szente * and Howard M. Johnson<br />

University of Florida, Gainesville, Florida<br />

I. Introduction<br />

The interferons (IFNs) were discovered in 1957 <strong>by</strong> Isaacs and Lindenman when they observed that a<br />

substance secreted <strong>by</strong> virally infected cells could protect other cells from viral infection [1a]. They<br />

called this substance interferon and found that it was a protein that caused uninfected cells to produce<br />

other proteins that made them resistant.<br />

Page 435<br />

Researchers since then have been finding a growing family of structurally related molecules: the<br />

interferons. Through the years, the interferons have been given many different names including immune,<br />

fibroblast, leukocyte, Type I, and Type II interferons. The recognized nomenclature includes alpha, beta,<br />

omega, tau, and gamma (α, β, ω, τ, and γ) interferons. Alpha, beta, omega, and tau all belong to the<br />

similar Type I subclass. Gamma is the sole member of the Type II or immune interferon class. The Type<br />

I interferons all share a greater sequence homology to each other than they do to IFN-γ (for a recent<br />

general review of the IFNs, see Reference 1b).<br />

The IFNs exert their actions on cells via cell surface receptors. Type I IFNs share the IFN Type I<br />

receptor (IFN-R1) while IFN-γ has its own unique Type II receptor. The signal transduction pathways of<br />

Type I and Type II<br />

* Current affiliation: Brigham and Women's Hospital, Boston, Massachusetts.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_435.html [4/9/2004 12:10:29 AM]


Document<br />

Figure 1<br />

Interferon activity.<br />

Page 436<br />

receptor activation are similar. These pathways involve ligand and receptor binding followed <strong>by</strong> the<br />

activation of tyrosine kinases and the phosphorylation of various proteins and their subsequent<br />

interaction with transcription elements on DNA. The two receptor activation pathways differ at the level<br />

of ligand binding. Type I IFNs bind their receptor as a complex of a single ligand, ligand binding<br />

element, and an accessory molecule. The Type II binding event occurs as a dimer of IFN-γ binding to<br />

two identical receptor molecules leading to receptor dimerization and activation.<br />

The IFNs of all subclasses posses antiviral activity. Additionally, they produce cellular responses that<br />

are distinct from antiviral activity, including antiproliferative and immunomodulatory activities (Figure<br />

1). These activities have led to an interest in their use as potential therapeutics to combat viral disease,<br />

cancer, and autoimmune disease. Currently, the IFNs have a worldwide market in excess of 2 billion<br />

dollars annually. There are six FDA-approved indications in the United States with several more in<br />

clinical trials (Table 1). In fact, two of the top-ten grossing biotechnology-<strong>based</strong> drugs on the U.S.<br />

market are IFNs. Intron A is an IFN-α used for immune protection and has an annual U.S. market of<br />

$570 million. Roferon-A is another IFN-α used for hairy-cell leukemia and Kaposi's sarcoma with an<br />

annual market of $170 million. These sales are despite profound negative side effects associated with<br />

IFN treatment. High doses are required to achieve positive clinical results and can lead to severe flu-like<br />

symptoms including nausea, vomiting, and fever. These side effects can cause patients to drop out of<br />

treatment before beneficial effects are seen. Another drawback of IFNs as drugs is that they require<br />

parenteral delivery. The IFNs are protein drugs that must be administered <strong>by</strong> injection and<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_436.html [4/9/2004 12:10:34 AM]


Document<br />

Table 1 Approval Indications for IFNs<br />

FDA Approved Clinical Trials<br />

IFN-α Chronic hepatitis HIV infection<br />

Kaposi's sarcoma Colon tumors<br />

Genital warts (papillomavirus) Kidney tumors<br />

Bladder cancer<br />

Hairy cell leukemia Malignant melanoma<br />

Non-Hodgkin's lymphoma<br />

Chronic myelogenous leukemia<br />

Throat wartz (papillomavirus)<br />

IFN-β Relapsing remitting multiple sclerosis Basal cell carcinoma<br />

IFN-γ Chronic granulomatous disease Kidney tumors<br />

Leishmaniasis<br />

cannot be given orally. For many of the clinical indications, treatments of many months are needed<br />

requiring repeat injections.<br />

Page 437<br />

These drawbacks, coupled with the market value of IFN-related treatments, now and in the future, have<br />

created an interest in producing second-generation molecules that can mimic IFN activity. These<br />

“mimetics” could potentially have greater specificity with fewer side effects. They may also have the<br />

advantages of reduced manufacturing costs and more versatile delivery. The design of mimetics can be<br />

achieved through structure-<strong>based</strong> drug design methodologies that are currently being developed.<br />

However, in order to apply structure-<strong>based</strong> drug design to a protein, a solid understanding of the<br />

structure/function relationship is needed. A three-dimensional structure, taken alone, gives little insight<br />

into the activity of a protein. <strong>Structure</strong>/function studies must be done for the full potential of structure<strong>based</strong><br />

drug design to be realized.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_437.html (1 of 2) [4/9/2004 12:10:37 AM]


Document<br />

<strong>Structure</strong>/function studies can take a variety of forms and use a number of techniques including the use<br />

of molecular biology, synthetic peptides, and antibodies, or combinations of these methods. Molecular<br />

biology is a powerful tool for structure/function analysis. Mutagenesis of cDNAs to produce mutant<br />

proteins with point mutations, truncations, or deletions can identify functional sites. One drawback to<br />

this approach, especially with large proteins, is the proverbial “needle in a haystack” problem. One has<br />

difficulty determining where to begin placing mutations. The synthetic peptide approach can be equally<br />

as powerful. One can synthesize individual domains or segments of proteins and test them for agonist or<br />

antagonist activities there<strong>by</strong> identifying functional domains. Synthetic peptides can be used to map the<br />

epitope specificity of antibodies that block the activity of a protein. Peptides can also be used to produce<br />

monospecific antisera to a defined region of a protein. The antibody approach has also proved quite<br />

useful in determining functional sites of<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_437.html (2 of 2) [4/9/2004 12:10:37 AM]


Document<br />

Page 438<br />

proteins. One potential disadvantage to using antibodies is the possibility of over interpreting the<br />

blocking results because of steric hindrance. A large antibody molecule may inhibit function through<br />

binding to a distant site and covering up a functional site. These approaches have all been used with<br />

success on a variety of proteins, but are best used in combination. For example, the information obtained<br />

from synthetic peptides and antibodies can significantly narrow down the region for site-directed<br />

mutagenesis studies. The entire sequence is narrowed to a segment, which reduces the size of the<br />

“haystack” in which the needle is hidden.<br />

Even though these approaches are powerful methods for determining functional sites on proteins, they<br />

are limited if not coupled with some form of structural determination. As Figure 2 illustrates, molecular<br />

biology and synthetic peptide/antibody approaches are not only interdependent, they are tied in with<br />

structural determination. Structural determination methods can take many forms, from the classic x-ray<br />

crystallography and NMR for three-dimensional determination, to two-dimensional methods such as<br />

circular dichroism and Fourier Transformed Infrared Spectroscopy, to predictive methods and modeling.<br />

A structural analysis is crucial to the interpretation of experimental results obtained from mutational and<br />

synthetic peptide/antibody techniques.<br />

Figure 2<br />

Flow diagram of structure/function studies.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_438.html [4/9/2004 12:10:42 AM]


Document<br />

Table 2 Three-Dimensional <strong>Structure</strong> Studies of the IFNs<br />

Type of Study Reference<br />

X-ray studies<br />

IFN-β<br />

IFN-γ human<br />

IFN-γ bovine<br />

IFN-γ rabbit<br />

IFN-γ human + receptor<br />

NMR studies<br />

Models<br />

IFN-γ human<br />

IFN-γ mouse N-terminal peptide (1–39)<br />

IFN-α2a<br />

IFN-α8<br />

IFN-τ sheep<br />

2<br />

3<br />

C.T. Samudzi, J.R. Rubin, unpublished data<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

T. Senda, S.I. Saitoh, Y. Mitsui, J.Li, and R.M. Roberts,<br />

unpublished data<br />

Note: This is not meant to be an exhaustive list of all structural studies of the IFNs. It only highlights some of the<br />

three-dimensional studies that have been conducted.<br />

Page 439<br />

While there are no hard-and-fast rules for conducting structure/function studies, the approaches taken for<br />

studying the IFNs can be used to illustrate some of the methods that have been successful. Over the<br />

years, a large body of work has accumulated on the IFNs, including a number of structural studies. Table<br />

2 summarizes some of the studies exploring the three-dimensional structure of the IFNs. The following<br />

sections review some of the structure/function studies that have begun to elucidate important features of<br />

IFN activity and form a basis for future rational drug design.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_439.html (1 of 2) [4/9/2004 12:10:45 AM]


Document<br />

II. Type IFNs<br />

A great deal of structure/function analysis has been done on the Type I IFNs. One Type I IFN in<br />

particular, IFN-τ has received attention recently because of its lower cytotoxicity compared to the IFNαs.<br />

<strong>Structure</strong>/function studies have concentrated on comparing IFN-τ with IFN-α. Therefore, IFN-τ<br />

provides an excellent example of structure/function studies of the Type I IFNs.<br />

First isolated from the conceptuses of sheep, IFN-τ is the major conceptus secretory protein responsible<br />

for signaling maternal recognition of pregnancy in ruminants [11]. it is produced in large quantities (200<br />

μg in 30 h from a day 16 conceptus culture). The protein was purified using a combination of anion<br />

exchange and molecular sieve chromatography.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_439.html (2 of 2) [4/9/2004 12:10:45 AM]


Document<br />

A. IFN-τ Synthetic Peptide Studies<br />

Page 440<br />

A sheep blastocyst library was screened with a probe <strong>based</strong> on the N-terminal sequence of the IFN-τ<br />

protein and the cDNA obtained (Table 3). Surprisingly, it exhibited 45–55% homology with various<br />

IFNs from human, mouse, rat, and pig and 70% homology with bovine IFN-ω [12]. It shared both<br />

molecular weight (19 kDa) and pI (5.4–5.6) with IFN-αs, while its length, 172 amino acids, was<br />

equivalent to the IFN-ωs. In competition studies, IFN-τ was found to compete with IFNs α, β, and ω for<br />

binding to the Type I IFN receptor [13]. In contrast, IFN-τ exhibited several unique properties such as<br />

its reproductive function, its poor inducibility <strong>by</strong> virus, and its apparent reduced cytotoxicity. Thus, IFNτ<br />

conceptus protein appears to be a novel IFN.<br />

Structural studies began with production of overlapping synthetic peptides, each 30–35 amino acids in<br />

length, corresponding to the entire<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_440.html (1 of 2) [4/9/2004 12:11:03 AM]


Document<br />

Page 441<br />

sequence of the molecule [14]. The peptides were used in competition assays with the native molecule.<br />

Peptide inhibition of a particular function would implicate the region of the molecule that it represented<br />

in the elicitation of that function. The effect of the IFN-τ peptides on the antiviral activity of ovine IFNτ<br />

was examined in a dose/response assay using Madin Dar<strong>by</strong> bovine kidney (MDBK) cells challenged<br />

with vesicular stomatitis virus. The carboxy-terminal peptide oIFN-τ(139–172) was found to be the most<br />

effective inhibitor of antiviral activity. Three additional peptides, oIFN-τ(1–37), (62–92), and<br />

(119–150), also reduced IFN-τ antiviral activity. This suggested that multiple regions of the IFN-τ<br />

molecule interact with the Type I IFN receptor and elicit antiviral activity. These regions are underlined<br />

in Table 3. The data were consistent with studies of antiviral activity and receptor binding with IFN-α<br />

analogs demonstrating that 3 distinct sites, located in the amino-terminal, internal, and carboxy-terminal<br />

regions of the molecule, influenced human IFN-α activity [15].<br />

To verify functional results using synthetic peptides, antipeptide antisera were produced [14]. All<br />

antipeptide antisera were reactive with the native molecule. Interestingly, antisera titers correlated with<br />

the hydropathic index of the peptide, rather than with the predicted surface accessibility of the specific<br />

region in the 3-D configuration. Consistent with the peptide studies, antisera against the same four<br />

regions of the molecule inhibited IFN-τ activity while antisera to other regions did not.<br />

Since IFN-τ and IFN-α bind to the same receptor, the ability of the IFN-τ synthetic peptides to block<br />

both bovine and human IFN-α was examined. Interestingly, only three of the four inhibitory peptides<br />

were effective competitors of IFN-α. Cross-inhibition of IFN-α <strong>by</strong> the internal and carboxy-terminal<br />

peptides was observed and suggested that these residues may adopt a similar conformation in both<br />

molecules and bind to a common site on the receptor. The aminoterminal peptide failed to reduce IFN-α<br />

function entirely. Thus, either the IFN-α amino-terminus has a much higher affinity for receptor or the<br />

IFN-τ aminoterminus binds a unique site on the receptor complex that may be associated with its unique<br />

properties. As expected, none of the peptides blocked the antiviral activity of IFN-τ, which interacts<br />

with a different receptor.<br />

Next, it was determined whether the same active regions of IFN-τ were involved in additional systems.<br />

The Type I IFN receptor on cells has been reported to be somewhat more promiscuous than on other<br />

cell types [16]; therefore, vesicular stomatitis challenge of Fc-9 cells was performed. Only the carboxyterminal<br />

peptide inhibited IFN-τ activity in this system [17]. This suggested that it was the carboxyterminus<br />

that was crucial to receptor interaction. In studies examining IFN-τ-treated feline<br />

immunodeficiency virus infected FeT-1 cells and human immunodeficiency virus-infected peripheral<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_441.html [4/9/2004 12:11:10 AM]


Document<br />

blood lymphocytes, peptide inhibition of IFN-τ antiretroviral activity implicated both the amino- and<br />

carboxy-termini as functionally important [17].<br />

Page 442<br />

The structural basis of the antiproliferative activity of IFN-τ was also investigated. While multiple<br />

regions were again involved in IFN-τ antiproliferative activity, it was the area adjacent to the carboxy<br />

terminus, rather than the carboxy-terminus itself, which was the most crucial for antiproliferative<br />

activity, inhibiting cell division <strong>by</strong> blocking entry into the S phase of the cell cycle [18]. Since, for a<br />

particular IFN-α subtype, antiviral potency does not necessarily correlate with antiproliferative potency,<br />

localization of these functions in different domains of the molecule is not unexpected [19]. Within all<br />

known IFN-αs, the 8 amino acids from 139 to 147 are highly conserved. These residues are contained in<br />

both carboxy-terminal peptides, but while they may be involved in antiviral activity, they do not appear<br />

to be solely responsible for antiproliferative activity since the two peptides are not equivalent inhibitors<br />

of IFN-τ antiproliferative activity. This observation is consistent with inhibition of antiviral activity but<br />

not antiproliferative activity <strong>by</strong> a monoclonal antibody in this conserved region in human IFN-α and<br />

with the requirement for tyrosine at position 123 for human IFN-α 1 antiproliferative activity [20,21]. It<br />

has also been reported that mutations around Arg33 affected both antiviral and antiproliferative activity<br />

of human IFN-α 4 on human cells [22], while the amino-terminus did not appear to be as important in<br />

IFN-τ antiproliferative activity on bovine cells.<br />

B. IFN-τ Monoclonal Antibodies<br />

Another approach to structure/function analysis of IFN-τ involved generation of anti-IFN-τ monoclonal<br />

antibodies. Four monoclonal antibodies were produced that reacted with the native IFN-τ protein. They<br />

were epitope mapped using the available IFN-τ peptides. Two of the antibodies were directed against the<br />

carboxy-terminus of the molecule, one against a region adjacent to the aminoterminus, and the final one<br />

appeared to react with a conformational, rather than a linear determinant (C. Pontzer, unpublished data).<br />

When these antibodies were used as competitors in binding assays, all four inhibited IFN-τ binding to<br />

the Type I IFN receptor on MDBK cells. That anti-IFN-τ carboxy-terminal antibodies would inhibit<br />

binding is not unexpected, but the inhibitory activity of the monoclonal antibodies directed against the<br />

more amino-terminal region was not anticipated. There is the caveat that warns that results using<br />

monoclonal antibodies to delineate function sites must be interpreted with caution since their size may<br />

cause significant steric hindrance.<br />

To proceed further with structural studies of IFN-τ, access to larger quantities of pure protein was<br />

required. The obvious route to this end entailed production of recombinant protein. A synthetic gene for<br />

IFN-τ was designed<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_442.html [4/9/2004 12:11:12 AM]


Document<br />

Page 443<br />

that allowed for optimal expression in both bacterial and yeast systems [23]. In addition, restriction sites<br />

were incorporated at intervals throughout the length of the sequence to allow for cassette mutagenesis.<br />

Using the Pichia pastoris expression system, 50 mg of purified IFN-τ were obtained from a one-liter<br />

culture.<br />

C. IFN-τ Binding and Signal Transduction<br />

Detailed receptor-binding studies were performed comparing recombinant human IFN-α and IFN-τ<br />

[24]. The K d of 125I-IFN-τ and 125I-IFN-αA for receptor on MDBK cells was 3.9 × 10 -10 M and 4.45 ×<br />

10 -11 M, respectively. Consistent with the higher binding affinity, IFN-αA was several fold more<br />

effective than IFN-τ as a competitive inhibitor. Functionally, the two IFNs had similar specific antiviral<br />

activities, but IFN-τ was 30 fold less toxic to MDBK cells at high concentrations. Phosphorylation of the<br />

signal transduction proteins, Tyk2, Stat1a, and Stat2 did not appear to be involved in the cellular toxicity<br />

associated with IFN-α relative to IFN-τ. Excess IFN-τ did not block the cytotoxicity of IFN-αA,<br />

suggesting that they recognize the receptor differently. While maximal IFN antiviral activity required<br />

only fractional receptor occupancy, toxicity was associated with maximal occupancy. Thus, “spare”<br />

receptors may exist with respect to certain biological properties, and IFNs may induce a concentrationdependent<br />

selective association of receptor subunits.<br />

D. Structural Biology of IFN-τ<br />

In order to better interpret the information derived from the above studies, an understanding of the 3-D<br />

structure of IFN-τ is required. Prior to resolution of the crystal structure, modeling techniques were<br />

employed for structural predictions [10]. For IFN-τ, the homology it shares with the other IFNs can be<br />

exploited. Since the x-ray coordinates for IFN-β are known [2] (Figure 3), it was used as a template for<br />

predicting the topology of IFN-τ. When the sequences of IFN-τ and IFN-β are aligned, the overall<br />

homology is approximately 30%. When residues are compared on the basis of conservative<br />

substitutions, the similarity rises to about 50%, and if only the location of hydrophobic residues is<br />

compared, the similarity is approximately 75%. This is important because hydrophobicity is thought to<br />

be a critical factor in driving protein folding. The interferons IFN-β, IFN-α-2, and several other<br />

cytokines including IL-2, IL-4, growth hormone, and GM-CSF belong to a family in which all share a<br />

four-helix bundle structural motif. Four-helix bundles exhibit a characteristic apolar periodicity in the α<br />

helices where every third or fourth residue is apolar, forming a hydrophobic strip down one side of the<br />

helix, which facilitates packing. The aligned helical regions of IFN-τ show the same apolar periodicity,<br />

suggesting a four-helix bundle motif.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_443.html [4/9/2004 12:11:16 AM]


Document<br />

Figure 3<br />

Stereo view of IFN-β crystal structure [2].<br />

Page 444<br />

The structure of IFN-τ was also examined <strong>by</strong> CD [10]. Analysis of the IFN-τ spectra predicts that the<br />

secondary structural elements derived from CD spectra indicate approximately 70% α-helix. The<br />

remainder of the molecule is either predicted to be random or a combination of β sheet and turn. Since it<br />

is known that algorithms that predict secondary structures from CD spectra are most accurate at<br />

identifying α helices, we are confident that IFN-τ is mainly α helical. The CD spectra for the synthetic<br />

peptides of IFN-τ were also obtained. The peptides IFN-τ(1–37), IFN-τ(62–92), IFN-τ(119–150), and<br />

IFN-τ(139–172) all show the presence of α helix, while IFN-τ(34–64) and IFN-τ(90–122) are mainly<br />

random. The presence of an α helix in the peptides supports the CD analysis of the intact protein and<br />

also roughly indicates the location of helical segments.<br />

The secondary structure of IFN-τ, including the location of the α helices and loop region, was then<br />

predicted using a neural network-<strong>based</strong> computer program called PHD that relies on sequence<br />

alignments of all proteins related to the target sequence [25,26]. When this prediction is correlated with<br />

the CD data, peptides that possess considerable α helicity are predicted to contain entire helical<br />

segments, and conversely, peptides with little helicity are predicted to be within loop regions.<br />

A model of the 3-D structure of IFN-τ was constructed using a distance geometry-<strong>based</strong> homology<br />

modeling method with mouse IFN-β acting as a template. The distance constraints were generated<br />

between residues within IFN-τ that are homologous to residues of IFN-β. Dihedral-angle restraints of α<br />

helices were generated from the secondary-structure prediction of IFN-τ. No constraints were applied to<br />

the 13-residue carboxy tail of IFN-τ, which is absent in IFN-β, since it is likely to be flexible in a<br />

manner similar to other proteins such as IFN-γ. Additional distance constraints were added from putative<br />

disul-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_444.html (1 of 2) [4/9/2004 12:11:24 AM]


Document<br />

Figure 4<br />

Stereo view of IFN-τ model. Highlighted sequences are from 1–37, 62–92,<br />

and 139–172.<br />

Page 445<br />

fide bridges between residues 1 and 99 and residues 29 and 139. Several structures were generated using<br />

distance geometry routines, and the energy was minimized and averaged to yield a final model [16]. A<br />

similar model was built <strong>by</strong> Senda et al. (unpublished results) using a homology modeling method. This<br />

model was also built using the x-ray coordinates of IFN-β and shows a similar topology to the IFN-β<br />

three-dimensional structure (Figure 4). The most striking feature of both models is that those<br />

discontinuous regions, previously determined to be functionally important, are localized to one side of<br />

the molecule and found to be spatially contiguous (Figure 4). This observation is consistent with<br />

multiple binding sites on IFN-τ interacting simultaneously with the Type I IFN receptor and emphasizes<br />

the importance of structural modeling in the understanding and interpretation of functional data.<br />

III. Type II IFN<br />

A. Functional Sites on the IFN-γ Molecule<br />

The production of IFN-γ-neutralizing antibodies specific for an N-terminal peptide of human IFN-γ<br />

provided the first evidence that the N-terminus of IFN-γ contained an important functional site [27]. A<br />

similar approach was used to produce N-terminus-specific neutralizing antisera against murine IFN-γ<br />

[28]. Subsequent studies using IFN-γ synthetic peptides to map the epitope specificity of monoclonal<br />

antibodies to murine IFN-γ showed that N-terminal specific monoclonal antibodies neutralize IFN-γ<br />

antiviral activity [29]. In receptor-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_445.html [4/9/2004 12:11:30 AM]


Document<br />

Page 446<br />

competition studies, murine IFN-γ N-terminal peptide consisting of residues 1–39 [IFN-γ (1–39)]<br />

blocked both binding to receptor and antiviral activity of IFN-γ [30]. Overlapping peptides of other<br />

regions of the IFN-γ molecule failed to block binding and function of IFN-γ [31]. Thus the combination<br />

of peptide mapping of epitope specificities and receptor competition using peptides has identified the Nterminus<br />

as a structurally and functionally important region of the IFN-γ molecule. This region is<br />

highlighted in the sequence of human IFN-γ found in Table 3.<br />

Interestingly, site-specific antibodies to the C-terminus of murine IFN-γ, which were induced using the<br />

peptide consisting of residues 95–133 [IFN-γ (95–133)], also neutralized IFN-γ activity, however IFN-γ<br />

(95–133) failed to block binding of IFN-γ to receptor and IFN-γ activity simultaneously. Antibodies to<br />

internal peptides failed to block both antiviral activity and binding of IFN-γ to receptor. In studies with<br />

recombinant murine IFN-γ receptor, which consisted of the entire α chain except for the transmembrane<br />

domain, the C-terminal peptide did block binding of IFN-γ to receptor [32]. Thus we have the interesting<br />

paradox wherein the IFN-γ C-terminal peptide blocked binding of IFN-γ to the recombinant, soluble<br />

receptor and yet did not block binding to the cell-surface receptor. One interpretation of these findings<br />

has allowed us to formulate the “velcro-key” model of binding to receptor that involves both N- and Cterminal<br />

domains of IFN-γ (Figure 5). The N-terminus binds in the “lock and key” manner characterized<br />

<strong>by</strong> specific ligand-receptor binding. The hydrophilic C-terminus binds to a region of the receptor distinct<br />

from that for the N-terminus, most likely through its polycationic region, which is conserved across<br />

species barriers. Binding of this type would exhibit high affinity and low specificity, similar to a piece of<br />

velcro. The C-terminal peptide of IFN-γ would therefore act as a poor competitor for cell-surface<br />

binding due to its low specificity alone. This interaction becomes specific in the context of the whole<br />

IFN-γ molecule and may increase the affinity of receptor binding.<br />

An alternative explanation that may also account for the inability of the C-terminal peptide to compete<br />

for cell-surface interactions is that its binding site is located not on the extracellular domain of the<br />

receptor, but rather on the intracellular domain. The primary differences between the cell-surface form<br />

of the IFN-γ receptor and (2) the accessibility of the recombinant receptor's cytoplasmic domain. A<br />

synthetic peptide corresponding to the membrane proximal region of the cytoplasmic domain of the<br />

murine IFN-γ receptor was able to bind IFN-γ and specifically compete with the binding of IFN-γ<br />

(95–133) to fixed/permeabilized cells [33].<br />

Studies <strong>by</strong> others have reaffirmed the importance of both the N- and C-terminal regions of IFN-γ in<br />

function. Using recombinant DNA techniques, it has been shown that deletion of residues from the Nterminus<br />

of the molecule<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_446.html [4/9/2004 12:11:32 AM]


Document<br />

Figure 5<br />

Velcro-key model of IFN-γ binding to its<br />

receptor. (From Reference 25. Copyright<br />

1992. The American Association of<br />

Immunologists.)<br />

Page 447<br />

results in decreased receptor binding [34]. Deletions or substitutions at the C- terminus have a direct<br />

effect on function of the molecule [35–38]. Epitope mapping of neutralizing monoclonal antibodies has<br />

also revealed an internal region of the molecule (from residues 84–94) as being functionally important<br />

[39]. This sequence bears strong homology to the nuclear localization sequence (NLS) of the SV40 large<br />

T antigen and has recently been demonstrated to be fully functional as an NLS for IFN-γ [40]. Thus,<br />

internal regions of the IFN-γ molecule are also likely to play an important functional role.<br />

B. IFN-γ Receptor α Chain Sites of Interaction with IFN-γ<br />

Both the human and the murine IFN-γ receptors consist of a ligand-binding subunit and a speciesspecific<br />

cofactor molecule. It is through interaction with this cell-surface receptor complex that IFN-γ<br />

exerts its biological effects. The IFN-γ molecule and its N-terminal peptide IFN-γ (1–39) bind<br />

specifically to the<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_447.html (1 of 2) [4/9/2004 12:11:36 AM]


Document<br />

Page 448<br />

cell-surface receptor and to a recombinant, soluble form of the ligand-binding chain of the receptor [25].<br />

Synthetic peptides corresponding to the sequence of the extracellular domain of the ligand-binding<br />

subunit were used to define the region of the receptor to which the N-terminus of IFN-γ binds. Receptor<br />

peptide MIR (95–120) competed most strongly with IFN-γ binding to both cell-surface and recombinant,<br />

soluble receptor [41]. Additionally, antisera to this peptide and the adjacent overlapping peptide, MIR<br />

(118–143), inhibited the binding of IFN-γ to the recombinant, soluble receptor. Therefore, the receptor<br />

domain responsible for binding the N-terminus of IFN-γ is defined <strong>by</strong> the region encompassing residues<br />

95–120 of the ligand-binding subunit of the IFN-γ receptor and may extend further into the neighboring<br />

sequence.<br />

Antibodies to the C-terminal region of IFN-γ have been shown to be potent neutralizers of IFN-γ activity<br />

[29]. However, no cell-surface binding site for the C-terminus of IFN-γ could be localized using either<br />

antisera or synthetic peptides. Furthermore, as indicated above, the C-terminal IFN-γ peptide, IFN-γ<br />

(95–133), competed specifically with the intact IFN-γ molecule for binding to a recombinant, soluble<br />

form of the receptor, which consists of both the extracellular and the intracellular domains [42]. It was<br />

hypothesized that since the intracellular portion of the soluble receptor was accessible, in contrast to that<br />

of the cell-surface receptor, the C-terminus of IFN-γ might indeed be binding to this region. In studies<br />

using synthetic peptides corresponding to the cytoplasmic domain of the murine IFN-γ receptor, only<br />

peptide MIR (253–287) specifically bound both murine IFN-γ and its C-terminal peptide, MuIFN-γ<br />

(95–133) [33]. This peptide corresponds to the membrane proximal region of the receptor's cytoplasmic<br />

region. Antibodies to this receptor peptide inhibited the binding of the C-terminus of murine IFN-γ to<br />

the receptor in cells which had been fixed and permeabilized. Analogous binding studies with human<br />

IFN-γ and its C- terminal peptide, HuIFN-γ (95–134), yielded a similar result [43]. Surprisingly, the<br />

binding of the IFN-γ C-terminal peptides to their cytoplasmic binding sites is not species restricted,<br />

which is in contrast to the binding of the whole molecule at the cell surface. Both human and murine<br />

IFN-γ and their C-terminal peptides bound equally well to receptor peptides of either human or murine<br />

origin [43]. Thus, a receptor binding site for the C-terminus of the IFN-γ molecule has been localized to<br />

the membrane proximal region of the ligand-binding subunit's cytoplasmic domain (Figure 6).<br />

Previously, there have been several reports of human IFN-γ having activity on murine cells when<br />

administered cytoplasmically [44–46]. With the identification of a cytoplasmic binding site for IFN-γ,<br />

which is not species restricted, the question arose as to whether this might be the basis for these earlier<br />

observations. Thus, C-terminal IFN-γ peptides of both human and murine origin were used to stimulate<br />

murine macrophage lines P388D 1 WEHI-3. Macrophages were chosen particularly for their capacity to<br />

nonspecifically endocytose material,<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_448.html [4/9/2004 12:11:38 AM]


Document<br />

Figure 6<br />

Proposed receptor activation pathway for IFN-γ. (From Reference 53. Copyright 1995.<br />

The American Association of Immunologists.)<br />

Page 449<br />

and we took advantage of this as a means of introducing the IFN-γ peptides into these cells. The IFN-γ Cterminal<br />

peptides induced a potent antiviral state in the murine macrophages and upregulated expression<br />

of MHC class II molecules, both in a dose-dependent fashion [43]. These effects were demonstrated to<br />

be sequence specific, as a scrambled version of the murine C-terminal peptide lacked activity.<br />

Furthermore, a truncated form of the murine C-terminal peptide, lacking the sequence of basic amino<br />

acids (RKRKR), was also without activity. The absence of activity of this truncated peptide was linked<br />

directly to a loss of its ability to bind to the receptor [43]. Therefore, interaction of IFN-γ, via its Cterminus,<br />

with its cytoplasmic binding site is important for function and requires the presence of a<br />

region of basic amino acids near the C-terminus of the molecule.<br />

C. Structural Biology of IFN-γ and the IFN-γ Receptor<br />

<strong>Structure</strong>-function studies of IFN-γ carried out using the synthetic peptide approach and site-specific<br />

antibodies indicated that both the N- and C-terminal regions of the protein were not only functionally<br />

important, but also accessible at the surface of the molecule. The x-ray crystal structure of human IFN-γ<br />

has been determined and reveals that in the IFN-γ homodimer both the N- and the<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_449.html [4/9/2004 12:11:51 AM]


Document<br />

Figure 7<br />

Stereo view of IFN-γ [40]. Regions 1–39 of chain A and 95–119 of chain B of the dimer<br />

are highlighted.<br />

Page 450<br />

C-terminus were indeed accessible [3]. Figure 7 illustrates the close proximity of the N- and C-termini<br />

of IFN-γ. The subunits of the homodimer are oriented head to tail, such that the N-terminal helix-loophelix<br />

(corresponding to residues 1–39) of one IFN-γ molecule interacts with the C-terminus of the<br />

second IFN-γ molecule.<br />

As mentioned above, synthetic peptides were also instrumental in identifying the region of the receptor<br />

to which the N-terminus of IFN-γ binds. Recently, the crystal structure of a complex between murine<br />

IFN-γ and the murine IFN-γ Rα subunit has been determined [5]. The synthetic peptides and<br />

corresponding antisera had predicted an interaction of murine IFN-γ residues (1–39) with receptor<br />

region (95–143). The crystal structure confirmed these observations, indicating an interaction of IFN-γ<br />

residues (1–42) with receptor residues (108–132). However, the crystal structure did not define an<br />

extracellular binding site for the C-terminus of IFN-γ. There has been some speculation that the basic<br />

amino acid residues of the C-terminus may interact with an acidic patch on the receptor's extracellular<br />

domain, which would support the previously mentioned “velcro-key” model, but that the crystallization<br />

conditions precluded this interaction [5]. It is quite possible that such an interaction may occur as a<br />

transitional state prior to the internalization of the C-terminal portion of IFN-γ and interaction with the<br />

cytoplasmic region of the receptor. An alternative explanation for the apparent lack of a binding site for<br />

the IFN-γ C-terminus on the receptor's extracellular face is that its primary site of interaction is<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_450.html (1 of 2) [4/9/2004 12:11:58 AM]


Document<br />

Page 451<br />

with the cytoplasmic portion of the receptor as described above. Thus, the orientation of the C-terminal<br />

portion of the IFN-γ molecules in the receptor complex should be such that they are situated near to the<br />

cell membrane. When one examines the structure of the receptor-ligand complex, it is easy to see that<br />

this is indeed the case. Studies are currently under way to determine the crystal structure of the complex<br />

between the cytoplasmic domain of the IFN-γ receptor and the IFN-γ molecule, in particular the Cterminus.<br />

With regards to the definition of sites of interaction between receptors and ligands, the<br />

synthetic peptide approach has repeatedly proven to be an accurate indicator of structurally important<br />

regions of the IFN-γ/IFN-γ R system.<br />

D. Signal Transduction <strong>by</strong> IFN-γ<br />

Within the past several years some of the immediate-early signal transduction events initiated in<br />

response to IFN-γ stimulation have been elucidated. Treatment of cells with IFN-γ leads to the rapid<br />

activation of two protein tyrosine kinases, JAK1 and JAK2 [47]. The JAK kinases are a newly emerging<br />

family of protein kinases important in signaling via cytokines and growth factors. These proteins are<br />

unrelated to the src family of tyrosine kinases and are characterized as being larger, having two putative<br />

phosphotransferase domains and containing no characteristic SH2 or SH3 domains [48–51]. Members of<br />

the Janus kinase family are found associated with the cytoplasmic domains of cytokine and growth<br />

factor receptors at or near to the membrane proximal region [51].<br />

In the resting cell, JAK1 and JAK2 are found associated with the α and β/AF-1 chains of the IFN-γ<br />

receptor, respectively [52]. These kinases as well as the ligand-binding chain of the IFN-γ receptor are<br />

tyrosine phosphorylated in response to IFN-γ treatment [47,53,54]. This leads in turn to the tyrosine<br />

phosphorylation of a latent cytoplasmic transcription factor, known variously as p91, Stat 91, or Stat 1 α<br />

on tyrosine residue 701 [55]. It is interesting to note that the IFN-α signal-transduction pathway partially<br />

overlaps with that of IFN-γ. Stimulation of cells <strong>by</strong> IFN-α leads to the activation of JAK1 and another<br />

Janus family kinase, Tyk2 [56,57]. In turn, this cascade leads to phosphorylation of two latent<br />

cytoplasmic transcription factors, p84 (Stat 1β) and p113 (Stat 2β) in addition to the p91 (Stat 1α)<br />

activated <strong>by</strong> IFN-γ.<br />

The identification of tyrosine kinases that directly associate with the subunits of the IFN-γ receptor lead<br />

to the question of how the binding of IFN-γ might affect these proteins. Recently, the synthetic peptide<br />

method was used to identify two regions of the murine IFN-γ receptor's α chain as being important for<br />

interaction with the kinase JAK2 [58]. One of these regions lies in the distal portion of the cytoplasmic<br />

tail (residues 404–432), while the other (residues 283–309) is nearer to the membrane proximal region to<br />

which the C-terminal part of IFN-γ binds (residues 253–287). The fact that there are adjacent binding<br />

sites for JAK2 and IFN-γ implied a potential for interaction between the IFN-γ<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_451.html [4/9/2004 12:12:00 AM]


Document<br />

Page 452<br />

ligand and the machinery of signal transduction, namely JAK2. It was found that both intact murine IFNγ<br />

and its C-terminal peptide (95–133) are capable of specifically mediating an increase in the degree of<br />

association between the recombinant, soluble IFN-γ receptor and JAK2. These findings were further<br />

supported as IFN-γ and IFN-γ (95–133) caused an increase in the amount of JAK2 coprecipitating with<br />

the receptor from intact murine macrophages [58]. This has been the first such demonstration of an<br />

extracellular cytokine ligand participating directly in interaction with cytoplasmic signaling elements.<br />

E. IFN-γ as a Candidate for Rational <strong>Drug</strong> <strong>Design</strong><br />

The IFN-γ molecule is a potentially attractive model for the rational application of drug-design<br />

strategies. Reagents exist that are capable of either positively or negatively modulating the in vivo<br />

effects of IFN-γ. An initial target is quite simply at the level of receptor-ligand interaction. Synthetic<br />

peptide analogs of the N-terminal region have been successfully applied in vitro to inhibit interaction of<br />

intact IFN-γ with cell-surface receptors [30]. Interestingly, it has been observed that the N-terminal<br />

region of mouse IFN-γ has the ability to interact with the human receptor [59]. It was shown that the<br />

mouse peptide IFN-γ (1–39) had a 10-fold greater ability to block the binding of human IFN-γ to cellsurface<br />

receptors. This was shown to be correlated with a more stable structure in solution for the<br />

murine peptide and illustrates the importance of stable structure to receptor binding, which may be<br />

exploited when designing peptide mimetics. The solution structure of this peptide has also been<br />

determined and could provide the beginning steps for determining the structural requirements of an<br />

antagonist [7]. Future studies could focus on cocrystallization of the peptides with receptor or NMR<br />

studies of peptide domains of the receptor and IFN-γ. Recombinant, soluble forms of the extracellular<br />

domain of the ligand-binding subunit of the receptor have also been used in analogous fashion both in<br />

vitro to inhibit cell-surface binding and in vivo to interfere with disease progression [60]. Therefore<br />

prevention of potentially deleterious effects of IFN-γ may be achieved <strong>by</strong> preventing initial interactions<br />

with receptor molecules at the cell surface.<br />

A second candidate region lies within amino acid residues 84–94 of human IFN-γ and the corresponding<br />

region of its murine homologue. This portion of the molecule functions as a nuclear localization signal<br />

and, therefore, is also an attractive target for drug design. In cells treated with IFN-γ, the IFN-γ molecule<br />

traffics rapidly to the nucleus of the cell, usually within one to two minutes. When the IFN-γ molecule is<br />

crosslinked to its receptor, the resultant receptor-ligand complex migrates to the nucleus [40]. The<br />

implication is that this sequence may therefore be of use in artificially targeting proteins from the<br />

cytoplasm directly to the nucleus. This is potentially a very attractive method<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_452.html [4/9/2004 12:12:03 AM]


Document<br />

for directed subcellular targeting of synthetic transcriptional activators or repressors that might not<br />

otherwise be directed to the nucleus.<br />

Page 453<br />

Finally, the C-terminal region of IFN-γ offers another possibility. With respect to the induction of an<br />

antiviral state, or the upregulation of MHC class II molecule expression, synthetic peptides<br />

corresponding to the C-terminal 39 amino acids of either human or murine IFN-γ function as potent<br />

agonists. This activity is due to the interaction of these peptides with the cytoplasmic domain of the IFNγ<br />

receptor and the associated protein tyrosine kinases. Furthermore, in contrast to the stringent species<br />

specificity of intact IFN-γ, the action of the C-terminal peptides agonists is not limited <strong>by</strong> species<br />

constraints. This property therefore renders the C-terminal portion of the IFN-γ molecule an attractive<br />

model for the development of IFN-γ agonists and antagonists. The identification of the C-terminus of<br />

IFN-γ as that part of the molecule that contacts the cytoplasmic portion of the receptor implies that in<br />

developing IFN-γ agonists, the primary focus should be on this region of the protein. Corresponding<br />

antagonists may be developed <strong>based</strong> upon the portion of the receptor to which the IFN-γ C-terminus<br />

binds.<br />

IV. Conclusion<br />

The interest in IFNs as therapeutics has existed from their initial discovery in 1957. Since then scientists<br />

have been trying to understand the mechanism of their action and apply that knowledge to the treatment<br />

of many different diseases, meeting with some success. The effort now is to understand how IFNs work<br />

at the molecular level, with the goal being to design better, more specific therapeutics. Through<br />

structure/function studies, we now know where the functional sites lie on many of the IFNs. We also<br />

know the sites of interaction with their receptors and second messenger systems. From these studies,<br />

initial candidates for structure-<strong>based</strong> drug design have been identified. Although more work is needed to<br />

further characterize the IFNs and their receptor systems, the challenge now is to apply our existing<br />

knowledge and create second generation molecules that can modulate the many activities of these<br />

fascinating proteins.<br />

References<br />

1a. Isaacs A and Lindenmann J. Virus Interference. I. The interferon. Proc R Soc London Ser B 1957;<br />

147:258.<br />

1b. Johnson HM, Bazer FW, Szente BE, Jarpe MA. How interferons fight disease. Scientific American<br />

1994; 270:40–47.<br />

2. Senda T, Shimazu T, Matsuda S, Kawano G, Shimizu H, Nakamura KT, Mitsui Y. Three-dimensional<br />

crystal structure of recombinant murine interferon-β. EMBO J 1992; 11:3193–3201.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_453.html (1 of 2) [4/9/2004 12:12:07 AM]


Document<br />

3. Ealick SE, Cook WJ, Vijay-Kumar S, Carson M, Nagabhushan TL, Trotta PP, Bugg CE. Threedimensional<br />

structure of recombinant human interferon-g. Science 1991; 252:698–702.<br />

Page 454<br />

4. Samudzi CT, Burton LE, Rubin JR. Crystal structure of recombinant rabbit interferon-gamma at 2.7 A<br />

resolution. J Biol Chem 1991; 266:21791–21797.<br />

5. Walter MR, Windsor WT, Nagabhushan TL, Lundell DJ, Lunn CA, Zavodny PJ, Narula SK. Crystal<br />

structure of a complex between interferon-g and its soluble high-affinity receptor. Nature 1995;<br />

376:230–235.<br />

6. Grzesiek S, Dobeli H, Gentz R, Garotta G, Labhardt AM, Bax A. 1H, 13C, and 15N NMR backbone<br />

assignments and secondary structure of human interferon-gamma. Biochemistry 1992; 31 (35):8180–90.<br />

7. Sakai TT, Jablonski MJ, DeMuth PA, Krishna NR, Jarpe MA, Johnson HM. Proton NMR sequence<br />

specific assignments and secondary structure of a receptor binding domain of mouse γ-interferon.<br />

Biochemistry 1993; 32:5650.<br />

8. Murgolo NJ, Windsor WT, Hruza A, Reichert P, Tsarbopoulos A, Baldwin S, Huang E, Pramanik B,<br />

Ealick S, Trotta PP. A homology model of human interferon alpha-2. Proteins 1993; 17(1):62–74.<br />

9. Seto MH, Harkins RN, Adler M, Whitlow M, Church WB, Croze E. Homology model of human<br />

interferon-alpha-8 and its receptor complex. Protein Science 1995; 4:655–70.<br />

10. Jarpe MA, Johnson HM, Bazer FW, Ott TL, Curto EV, Rama Krishna N, Pontzer CH. Predicted<br />

structural motif of IFN-τ. Protein Engineering 1994; 7:863–867.<br />

11. Bazer<br />

FW.<br />

Mediators of<br />

maternal<br />

recognition<br />

of pregnancy<br />

in mammals.<br />

Proc Soc Exp<br />

Biol Med<br />

1992;<br />

199:373–384.<br />

12. Imakawa K, Anthony RV, Kazemi M, Marotti KR, Polites HG, Roberts RM. Interferon-like<br />

sequence of ovine trophoblast protein secreted <strong>by</strong> embryonic trophectoderm. Nature 1987; 330:377–379.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_454.html (1 of 2) [4/9/2004 12:12:10 AM]


Document<br />

13. Stewart HJ, McCann SHE, Barker PJ, Lee KE, Lamming GE, Flint APF. Interferon sequence<br />

homology and receptor binding activity of ovine trophoblast antileuteolytic protein. J Endocrinol 1987;<br />

115:R13–R15.<br />

14. Pontzer CH, Ott TL, Bazer FW, Johnson HM. <strong>Structure</strong>/function studies with interferon tau:<br />

Evidence for multiple active sites. J Interferon Res 1994; 14:133–141.<br />

15. Fish EN, Banerjee K, Stebbing N. The role of three domains in the biological activity of human<br />

interferon-α. J Interferon Res 1989; 9:97–114.<br />

16. Novick D, Cohen B, Rubinstein M. The human interferon α/β receptor: characterization and<br />

molecular cloning. Cell 1994; 77:391–400.<br />

17. Pontzer CH, Yamamoto JK, Bazer FW, Ott TL, Johnson HM. Potent anti-feline immunodificiency<br />

virus and anti-human immunodeficiency virus effect of interferon tau. J Immunol 1995; (in press).<br />

18. Pontzer CP, Bazer FW, Johnson HM. Antiproliferative activity of a pregnancy recognition hormone,<br />

ovine trophoblast protein-1. Cancer Res 1991; 51:5304–5307.<br />

19. Pestka S, Langer JA, Zoon KC, Samuel CE. Interferons and their actions. Annu Rev Biochem 1987;<br />

56:727–777.<br />

20. Barasoain I, Portolès A, Aramburu JF, Rojo JM. Antibodies against a peptide representative of a<br />

conserved region of human IFN-α. Differential effects on the antiviral and antiproliferative effects of<br />

IFN. J Immunol 1989; 143:507–512.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_454.html (2 of 2) [4/9/2004 12:12:10 AM]


Document<br />

Page 455<br />

21. McInnes B, Chambers PJ, Cheetham BF, Beilharz MW, Tymms MJ. <strong>Structure</strong>-function studies of<br />

interferons-α: Amino acid substitutions at the conserved residue tyrosine 123 in human interferon-α 1. J<br />

Interferon Res 1989; 9:305–314.<br />

22. Waine GJ, Tymms MJ, Brandt ER, Cheetham BF, Linnane AW. <strong>Structure</strong>-function study of the<br />

region encompassing residues 26–40 of human interferon-α 4: Identification of residues important for<br />

antiviral and antiproliferative activities. J Interferon Res 1992; 13:42–48.<br />

23. Ott TL, Van Heeke G, Johnson HM, Bazer FW. Cloning and expression in Saccharomyces<br />

cerevisiae of a synthetic gene for the type-1 trophoblast interferon ovine trophoblast protein-1:<br />

Purification and antiviral activity. J Interferon Res 1990; 11:357–364.<br />

24. Subramaniam PS, Khan SA, Pontzer CH, Johnson HM. Differential recognition of the type In IFN<br />

receptor <strong>by</strong> IFN-τ and IFN-α is responsible for their differential cytotoxicities. 1995; (submitted).<br />

25. Sander C, Schneider R. Database of homology-derived structure and the structural meaning of<br />

sequence alignment. Proteins 1991; 9:56–68.<br />

26. Rost B, Sander J. Prediction of protein structure at better than 70% accuracy. J Mol Biol 1993;<br />

232:544–599.<br />

27. Johnson HM, Langford MP, Lakchaura B, Chan TS, Stanton GJ. Neutralization of native human<br />

gamma interferon <strong>by</strong> antibodies to a synthetic peptide encodded <strong>by</strong> the 5' end of human gamma<br />

interferon cDNA. J Immunol 1982; 129:2357–2359.<br />

28. Langford MP, Gray PW, Stanton GJ, Lakchaura B, Chan T-S, Johnson HM. Antibodies to a<br />

synthetic peptide corresponding to the N-terminal end of mouse gamma interferon (IFN-α). Biochem<br />

Biophys Res Comm 1983; 117:866–871.<br />

29. Russell JK, Hayes MP, Carter MJ, Torres BA, Dunn BM, Russell SW, Johnson HM. Epitope and<br />

functional specificity of monoclonal antibodies to mouse interferon gamma: the synthetic peptide<br />

approach. J Immunol 1986; 136:3324–3328.<br />

30. Magazine HI, Carter JM, Russell JK, Torres BA, Dunn BM, Johnson HM. Use of synthetic peptides<br />

to identify an N-terminal epitope on mouse gamma interferon that may be involved in function. Proc<br />

Natl Acad Sci USA 1988; 185:1237–1241.<br />

31. Jarpe MA, Johnson HM. Topology of receptor binding domains of mouse IFN-α. J Immunol 1990;<br />

145:3304–3309.<br />

32. Griggs ND, Jarpe MA, Pace JL, Russell SW, Johnson HM. The N-terminus and C- terminus of<br />

interferon gamma are binding domains for cloned soluble interferon gamma receptor. J Immunol 1992;<br />

149:517–520.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_455.html (1 of 2) [4/9/2004 12:12:22 AM]


Document<br />

33. Szente BE, Johnson HM. Binding of IFN-α and its C-terminal peptide to a cyto-plasmic domain of<br />

its receptor that is essential for function. Biochem Biophys Res Comm 1994; 201:215–221.<br />

34. Zavodny PJ, Petro ME, Chiang TR, Narula SK, Leibowitz PJ. Alterations of the amino terminus of<br />

murine interferon gamma: expression and biological activity. J Interferon Res 1988; 8:483–494.<br />

35. Arakawa T, Hsu YR, Parker CG, Lai PH. Role of polycationic C-terminal portion in the structure<br />

and activity of recombinant human interferon gamma. J Biol Chem 1986; 261:8534–8539.<br />

36. Leinikki PO, Calderon J, Luquette MH, Schreiber RD. Reduced receptor binding <strong>by</strong> a human<br />

interferon gamma fragment lacking 11 carboxyl-terminal amino acids. J Immunol 1987;<br />

139:3360–3366.<br />

37. Wetzel R, Perry LJ, Veilleux C, Chang G. Mutational analysis of the C-terminus of human<br />

interferon gamma. Prot Eng 1990; 3:611–623.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_455.html (2 of 2) [4/9/2004 12:12:22 AM]


Document<br />

Page 456<br />

38. Lundell D, Lunn C, Dalgarno D, Fossetta J, Greenberg R, Reim R, Grace M, Narula S. The carboxylterminal<br />

region of human interferon-gamma is important for biological activity: mutagenic and NMR<br />

analysis. Prot Eng 1991; 4(3):335–341.<br />

39. Zu X, Jay FT, The E 1 functional epitope of the human interferon-α is a nuclear targeting signal-like<br />

element. J Biol Chem 1991; 266:6023–6026.<br />

40. Bader T, Wietzerbin J. Nuclear accumulation of interferon-gamma. Proc Natl Acad Sci USA<br />

1994;91:11831–11835.<br />

41. Van Volkenburg MA, Griggs ND, Jarpe MA, Pace JL, Russel SW, Johnson HM. Binding site on the<br />

murine interferon-gamma receptor for interferon-gamma has been identified using the synthetic peptide<br />

approach. J Immunol 1993; 151:6206– 6213.<br />

42. Fernando LP, LeClaire RD, Obici S, Zavodny PJ, Russell SW, Pace JL. Stable expression of a<br />

secreted form of the mouse IFN-α receptor <strong>by</strong> rate cells. J Immunol 1991; 147:541–547.<br />

43. Szente BE,<br />

Soos JM,<br />

Johnson HM.<br />

The C-terminus<br />

of IFN-α is<br />

sufficient for<br />

intracellular<br />

function.<br />

Biochem<br />

Biophys Res<br />

Comm 1994;<br />

203:1645–1654.<br />

44. Fidler IJ, Fogler WE, Kleinerman ES, Saiki I. Abrogation of species specificity for activation of<br />

tumoricidal properties in macrophages <strong>by</strong> a recombinant mouse or human interferon gamma<br />

encapsulated in liposomes. J Immunol 1985; 135:4289–4296.<br />

45. Sancéau J, Sondermeyers P, Béranger F, Falcoff R, Vaquero C. Intracellular human interferon<br />

triggers an antiviral state in transformed murine L cells. Proc Natl Acad Sci USA 1987;84:2906–2910.<br />

46. Smith MR, Muegge K, Keller JR, Kung HF, Young HA, Durum SK. Direct evidence for an<br />

intracellular role for interferon-gamma. J Immunol 1990; 144:1777–1782.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_456.html (1 of 2) [4/9/2004 12:12:24 AM]


Document<br />

47. Igarashi K, Garotta G, Ozmen L, Ziemiecki A, Wilks AF, Harpur AG, Larner AC, Finbloom DS.<br />

Interferon gamma induces tyrosine phosphorylation of interferon gamma receptor and regulated<br />

association of protein tyrosine kinases, Jak1 and Jak2, with its receptor. J Biol Chem 1994;<br />

269:14333–14336.<br />

48. Firmbach-Kraft I, Byers M, Shows T, Dalla-Favera R, Krolewski JJ. Tyk2, prototype of a novel<br />

class of non-receptor tyrosine kinase genes. Oncogene 1990; 5:1329–1336.<br />

49. Bernards A. Predicted tyk2 protein contains two tandem protein kinase domains. Oncogene 1991;<br />

6:1185–1187.<br />

50. Wilks AF, Harpur AG, Kurban RR, Ralph SJ, Zurcher G, Ziemiecki A. Two novel protein tyrosine<br />

kinases, each with a second phosphotransferase-related catalytic domain, define a new class of protein<br />

kinase. Mol Cell Biol 1991; 11:2057–2065.<br />

51. Ihle JN, Witthuhn BA, Quelle FW, Yamamoto K, Silvennoinen O. Signaling through the<br />

hematopoietic cytokine receptors. Annu Rev Immunol 1995; 13:369–398.<br />

52. Sakatsume M, Igarashi K-I, Winestock KD, Garotta G, Larner AC, Finbloom DS. The Jak Kinases<br />

differentially associate with the a and b (accessory factor) chains of the interferon-g receptor to form a<br />

functional receptor unit capable of activating STAT transcription factors. J Biol Chem 1995;<br />

270:17528–17534.<br />

53. Khurana Hershey GK, McCourt DW, Schreiber RD. Ligand-induced phosphorylation of the human<br />

interferong receptor. J Biol Chem 1990; 265:17868–17875.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_456.html (2 of 2) [4/9/2004 12:12:24 AM]


Document<br />

Page 457<br />

54. Greenlund AC, Farrar MA, Viviano BL, Schreiber RD. Ligand induced interferon-gamma receptor<br />

tyrosine phosphorylation couples the receptor to its signal transduction system ( p912). EMBO J 1994;<br />

13:1591–1600.<br />

55. Shuai K, Start GR, Kerr IM, Darnell JE. A single phosphotyrosine residue of Stat91 required for<br />

gene activation <strong>by</strong> interferon gamma. Science 1993; 261:1744–1746.<br />

56. Muller M, Briscoe J, Laxton C, Guschin D, Ziemiecki A, Silvennoinen O, Harpur AG, Barbieri G,<br />

Witthuhn BA, Schindler C, Pellegrini S, Wilks AF, Ihle JN, Stark GR, Kerr IM. The protein tyrosine<br />

kinase JAK1 complements defects in interferon alpha/beta and gamma signal transduction. Nature 1993;<br />

366:129–135.<br />

57. Barbieri G, Velazquez L, Scrobogna M, Fellous M, Pellegrini S. Activation of the protein kinase<br />

tyk2 <strong>by</strong> interferon α/β. Eur J Biochem 1994;223:427–435.<br />

58. Szente BE, Subramaniam PS, Johnson HM. Identification of IFN-γ receptor binding sites for JAK2<br />

and enhancement of binding <strong>by</strong> IFN-γ and its' C-terminal peptide IFN-γ(95–133). J Immunol 1995;<br />

155:95–133.<br />

59. Jarpe MA, Johnson HM. Stable conformation of IFN-γ receptor binding peptide in aqueous solution<br />

is required for IFN-γ antagonist activity. J Interferon Res 1993; 13:99.<br />

60. Ozmen L, Roman D, Fountoualakis M, Schmid G, Ryffel B, Garotta G. Soluble interferon-gamma<br />

receptor: a therapeutically useful drug for systemic lupus erythematosus. J Interferon Res 1994;<br />

14(5):283–284.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_457.html [4/9/2004 12:12:28 AM]


http://legacy.netlibrary.com/reader/message.asp?message=811&BookID=12640&FileName=Page_458.html<br />

The requested page could not be found.<br />

Return to previous page<br />

http://legacy.netlibrary.com/reader/message.asp?message=811&BookID=12640&FileName=Page_458.html [4/9/2004 12:12:31 AM]


Document<br />

18<br />

The <strong>Design</strong> of Anti-Influenza Virus <strong>Drug</strong>s from the X-ray Molecular<br />

<strong>Structure</strong> of Influenza Virus Neuraminidase<br />

Joseph N. Varghese<br />

Biomolecular Research Institute, Melbourne, Victoria, Australia<br />

I. Introduction<br />

Page 459<br />

Influenza has plagued humankind since the dawn of history and continues to affect a significant<br />

proportion of the population irrespective of age or previous infection history. These periodic epidemics<br />

that reinfect otherwise healthy people have devastated communities world wide. Some pandemics like<br />

the 1917–1919 “Spanish flu” were responsible for the death of tens of millions of people throughout the<br />

world. The origins, spread, and severity of influenza epidemics have been a puzzle that has only in the<br />

last two decades been adequately addressed. In early times it was thought that the disease was the evil<br />

influence (sic) of the stars, and other extraterrestial objects. At present it is generally accepted that the<br />

disease is of viral origin, spread <strong>by</strong> aerosols produced <strong>by</strong> infected animals, and the continual production<br />

of new strains of the virus results in reinfection of the disease (reviewed in Reference 1).<br />

There are three types of influenza virus classified on their serological cross-reactivity with viral matrix<br />

proteins and soluble nucleoprotein (A, B, and C). Only type A and B are known to cause severe human<br />

disease. Type B is only found in humans, while type A has a natural reservoir in birds and some<br />

mammals like pigs and horses [2]. Influenza, an orthomyxovirus, is a 100 nm lipid-enveloped virus<br />

(Figure 1) containing an eight-segment negative single-stranded genome [3]. Two of the segments code<br />

for the surfaces glycoproteins, hemagglutinin (which binds to terminal sialic acid), and neuraminidase<br />

(which<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_459.html [4/9/2004 12:12:33 AM]


Document<br />

Figure 1<br />

A schematic diagram of an influenza virus particle that illustrates<br />

its constituent components and morphology. The surface antigens hemagglutinin<br />

and neuraminidase are attached to the lipid and matrix protein shell that<br />

encapsulates the eight negative-stranded RNA genes of the virus and<br />

associated nucleoprotein and polymerase.<br />

Page 460<br />

cleaves terminal sialic acid) and which appear as spikes protruding out of the viral envelope. The viral<br />

target in humans is the upper respiratory tract epithelial cells. Replication (see Figure 2) begins with<br />

penetration of the virion through the mucin layer covering the epithelial surface, followed <strong>by</strong> attachment<br />

to the viral receptor <strong>by</strong> the hemagglutinin. Penetration of the cell is achieved <strong>by</strong> endocytosis and the<br />

virion core is released after the fusion of the virion and vesicle membrane mediated <strong>by</strong> the<br />

hemagglutinin. Fusion is enabled <strong>by</strong> a conformational change in the hemagglutinin made possible <strong>by</strong><br />

lowering the pH of the endosome <strong>by</strong> the M2 ion channel protein. Following replication, the progeny<br />

virions are released <strong>by</strong> budding off the cell membrane [4,5].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_460.html (1 of 2) [4/9/2004 12:12:42 AM]


Document<br />

Figure 2<br />

A simplified schematic of the replication cycle of an influenza virion in the host<br />

respiratory epithelia. Details of viral transport through mucin and pathways of<br />

viral spread on budding from the epithelial cell membrane during replication is<br />

not understood. Neuraminidase activity is important for release of budding<br />

progeny virions, desialyation of viral glycoproteins, and probably facilitates<br />

transport through sialic-acidrich mucin.<br />

Page 461<br />

Release of virions occur 8 hours post infection and the onset of infection is sudden, resulting in pyroxia,<br />

muscular and joint pain, and a dry cough [6]. Virus shedding continues for up to a week, when a rise in<br />

virus-specific antibody clears the virus from the host. The vulnerability of the host succumbing to<br />

viremea during this week of rising viral titer is mediated <strong>by</strong> interferon induction [7] 48 hours post<br />

infection, which attenuates viral replication until the cell-mediated immune response begins to clear the<br />

virus. The severity of the illness is thought to depend on the level of cross protection arising from<br />

antibodies raised from previous influenza infections [19]. The course of the illness can be debilitating,<br />

and no effective treatment is available at present to halt the progression of the disease. Death can result<br />

for susceptible populations (neonate and elderly) primarily as a result of secondary infections [8]. This<br />

chapter shall examine a structural basis for the continual emergence of new influenza strains, and the<br />

reasons current vaccines against influenza fail to protect against all<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_461.html (1 of 2) [4/9/2004 12:13:07 AM]


Document<br />

Page 462<br />

strains of influenza. The discovery of the active site of influenza neuraminidase and the exploitation of<br />

its structural conservation shall be discussed in terms of the design of potent neuraminidase inhibitors.<br />

The potential therapeutic use of these inhibitors as antiviral drugs against influenza virus infections shall<br />

be examined.<br />

A. Antigenic Variation<br />

The plethora of different strains of virus that are responsible for the continued reinfection of virus in<br />

humans is primarily related to mutations in the viral genes of two surface glycoproteins, hemagglutinin<br />

and neuraminidase [9]. The current paradigm for this genetic variation [10,11] is that these mutations<br />

arise primarily from incremental changes in the amino acid sequences of these glycoproteins <strong>by</strong><br />

selection pressure of the immune system of the infected host. This mechanism termed “Antigenic Drift”<br />

accounts for most of the strain variation within a particular subtype of influenza.<br />

However, infrequently a mutation arises <strong>by</strong> genetic reassortment of viruses from different animal hosts<br />

(“Antigenic Shift”) where<strong>by</strong> an entirely new gene for one of the surface glycoproteins is generated that<br />

is significantly different (~50%) in amino acid sequence from the parent virus. This is the mechanism <strong>by</strong><br />

which new subtypes of influenza arise and are primarily responsible for the major pandemics that occur.<br />

Strains of influenza virus are classified <strong>by</strong> type (A, B, or C), geographic location, date of original<br />

isolation, and the subtype of the hemagglutinin and neuraminidase antigens. There exist 9 known<br />

subtypes (N1 to N9) of neuraminidase and 13 known subtypes (H1 to H13) of hemagglutinin for<br />

influenza A in all animal populations. Two neuraminidase (N1 and N2) and three hemagglutinin (H1,<br />

H2, and H3) subtypes of influenza A have occurred in strains that have infected humans since 1933<br />

when isolates were first characterized [12]. Prior to 1933 there is indirect evidence of antigenic shift<br />

occurring in human populations [13]. The N1 subtype was associated with virus isolated between 1933<br />

and 1957, after which time the N2 subtype appeared in the Asian influenza. No major change in the<br />

structure of neuraminidase has occurred since, although the hemagglutinin subtype has changed from H2<br />

to H3 in 1968 in the Hong Kong pandemic, and H1N1 reappeared in 1978 as the Russian pandemic.<br />

Influenza B, which infects only human hosts, has only one subtype, but like type A undergoes continual<br />

antigenic drift.<br />

B. Current Therapeutics<br />

Amantidine and Rimantidine are the only class of drugs that have been approved for therapy. At high<br />

concentration (>50 mg/mL) Amantidine is thought to buffer the pH of the endosome and prevent the<br />

conformational<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_462.html [4/9/2004 12:13:12 AM]


Document<br />

Page 463<br />

change of the hemagglutinin necessary for fusion. <strong>Drug</strong>-resistant mutants arise where the hemagglutinin<br />

trimers are thought to be less stable than the wild type [14]. At low concentrations (


Document<br />

Page 464<br />

could not again be agglutinated <strong>by</strong> either the eluted virus or fresh virus preparations. This activity is now<br />

attributed to neuraminidase, which is one of the two integral membrane glycoproteins of influenza virus<br />

(for reviews see References 24 and 25).<br />

A. Function<br />

Neuraminidase is an exoglycosidase that destroys the hemagglutinin receptor <strong>by</strong> cleaving the αketosidic<br />

linkage of terminal sialic acid [(N-actylneuraminic acid (Neu5Ac))] to an adjacent sugar<br />

[26,27]. Viral hemagglutinin binds specifically to Neu5Ac-containing receptors on the surface of<br />

susceptible cells [28]. Neuraminidase, which also removes terminal sialic acid from a range of<br />

glycoconjugates, plays an important, but not completely understood, role in the viral replication cycle.<br />

Without neuraminidase activity viruses [29] were thought to be immobilized <strong>by</strong> mucosal secretions in<br />

the upper respiratory tract. By removing terminal sialic acid from the sialic-acid-rich mucous layer<br />

[27,30] protecting target cells, neuraminidase could facilitate penetration of the virus to the cell surface.<br />

It has been shown that neuraminidase-deficient virus [31] can still replicate in vivo, albiet at a much<br />

reduced rate [32]. This shows that neuraminidase does not play an essential role in viral entry,<br />

replication, assembly or budding in mice, but has an important role in the spread of the infection <strong>by</strong><br />

preventing aggregation at the cell surface and possible immobilization in the mucin <strong>by</strong> hemagglutinin.<br />

Once replication is initiated in the infected cell, the freshly synthesised viral glycoproteins have to be<br />

desialylated to prevent self-aggregation at the infected host cell surface <strong>by</strong> hemagglutinin binding to<br />

terminal sialic acid on these glycoproteins. Finally on elution of progeny virions from infected cells,<br />

neuraminidase activity is required to facilitate viral escape from the cell surface.<br />

Inactivation or inhibition of neuraminidase during budding has been observed to result in aggregation of<br />

virions on the cell surface [33–35]. Inhibition of this glycohydrolase could provide a means of<br />

controlling this disease <strong>by</strong> slowing the rate of viral attachment and subsequent release of progeny virions<br />

allowing the host immune system to eliminate the virus while the number of infected cells is low.<br />

B. Morphology<br />

There are between 50 to 100 neuraminidase spikes per virion [36] which is approximately 10% of the<br />

visible spikes projecting out of the surface of the virion [37]. These spikes can be removed from the<br />

virus <strong>by</strong> treatment with detergent [38]. Electron microscopic images of the neuraminidase spikes [39]<br />

reveal a mushroom-shaped molecule made up of a boxlike head of about 80 ×<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_464.html [4/9/2004 12:13:17 AM]


Document<br />

Page 465<br />

80 × 40 Å with a narrow centrally attached stalk (15 Å wide and 100 Å long), which terminates into a<br />

hydrophobic knob anchored in the viral envelope. The detergent-released spikes can be digested <strong>by</strong><br />

pronase to release the neuraminidase “heads,” which retain full antigenic and enzyme activity [40,41].<br />

The molecule was found to be a tetramer of molecular weight 240,000, reducing to 200,000 when<br />

treated with pronase [42]. A low resolution x-ray image of crystallized neuraminidase heads [43]<br />

established that the enzyme had circular 4-fold symmetry.<br />

III. Molecular <strong>Structure</strong> of Neuraminidase<br />

Crystals of pronase-released heads of the N2 human strains of A/Tokyo/3/67 [44] and A/RI/5+/57 were<br />

used for an x-ray structure determination. The x-ray 3-dimensional molecular structure of neuraminidase<br />

heads was determined [45] for these two N2 subtypes <strong>by</strong> a novel technique of molecular electron density<br />

averaging from two different crystal systems, using a combination of multiple isomorphous replacement<br />

and noncrystallographic symmetry averaging. The structure of A/Tokyo/3/67 N2 has been refined [46]<br />

to 2.2 Å as has the structures of two avian N9 subtypes [47–49]. Three influenza type B structures [50]<br />

have also been determined and found to have an identical fold with 60 residues (including 16 conserved<br />

cysteine residues) being invariant. Bacterial sialidases from salmonella [51] and cholera [52] have<br />

homologous structures to influenza neuraminidase, but few of the residues are structurally invariant.<br />

A. Structural Topology<br />

The protein fold consists of a symmetric arrangement of six four-stranded antiparallel β sheets arrange<br />

as blades of a propeller (Figure 3), the propeller axis being approximately parallel to but titled away<br />

from the circular 4-fold axis of the tetramer. This tilt angle varies between the known subtypes. This<br />

topology has now also been found in the seven β sheet propeller structure of bacterial methylamine<br />

dehydrogenase [53] and galactose oxidase [54], and the eight β sheet propeller structure of methanol<br />

dehydrogenase [55].<br />

Each sheet of neuraminidase has a “W” topology (+1,+1,+1) [56] with four strands connected <strong>by</strong> reverse<br />

turns (Figure 3). The first strand of each sheet enters from the top, approximately parallel to and near the<br />

propeller axis; the fourth strand exits from the bottom, approximately perpendicular to the propeller axis.<br />

Top and bottom surfaces of the head refer to the faces of the tetramer away and towards the viral<br />

membrane respectively. Each sheet thus has a characteristic 90° right-hand twist between the inner- and<br />

outermost strand. The six sheets and their connections to each other are topologically identical.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_465.html [4/9/2004 12:13:21 AM]


Document<br />

Figure 3<br />

(a) A MOLSCRIPT [107] stereo diagram of the neuraminidase tetramer viewed<br />

from above down the symmetry axis. The different shaded arrows represent β<br />

strands comprising the six β sheets that form the “propeller” framework of<br />

each subunit. (b) A stereo diagram of a neuraminidase monomer viewed down<br />

the 4-fold axis, and α-sialic acid is shown bound in the active site<br />

of the enzyme, which lies in a pocket formed <strong>by</strong> the six β<br />

sheets near the pseudo six-fold axis of each subunit.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_466.html (1 of 2) [4/9/2004 12:15:39 AM]<br />

Page 466


Document<br />

The first strand of each sheet is connected across the top of the submit to the fourth strand of the<br />

preceding sheet. The N-terminus lies across the bottom of the subunit, and builds the fourth strand of the<br />

sixth sheet before entering the first sheet. The C-terminus strand builds the third strand of the sixth<br />

sheet, enters the subunit interface from the bottom, and runs parallel with the outermost (fourth) strand<br />

of the sixth β sheet.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_466.html (2 of 2) [4/9/2004 12:15:39 AM]


Document<br />

B. Protein <strong>Structure</strong><br />

Page 467<br />

The 6 β sheets of the subunit, although topologically identical, vary in size and detailed structure.<br />

However, the six sheets can be considered as maintaining an approximate 6-fold symmetry relationship<br />

to each other about an axis parallel to the mean direction of the first strand of each sheet and through the<br />

centroid of these directions. Four disulfide bridges are formed between adjacent sheets as well as two<br />

between the sheets, which distort the sheet structure. Similar pairing has now been found in one of the<br />

domains of CD4 [57]. Also short β bulges occur on the inner and outer strands of most of the sheets. The<br />

packing of the β sheets is stabilized <strong>by</strong> hydrophobic interactions at the sheet interfaces and hydrogen<br />

bonding at the periphery of the sheets. In addition the intersheet disulfide bonds and ion pairs stabilize<br />

the structure.<br />

In general the loops connectioning the β sheets are shorter and tighter on the bottom of the subunit<br />

compared to the outer loops, which tends to be more elaborate. The intersheet loop connecting the fourth<br />

and fifth loop is the most extensive in the structure and is stabilized primarily <strong>by</strong> a disulfide bridge<br />

between Cys318 (sequence numbering of A/Tokyo/3/67 will be used throughout) and Cys337, a<br />

conserved ion pair Asp330—Arg364, and a putative Ca 2+ ion binding site. This calcium binding site has<br />

approximate octahedral coordination with the main-chain oxygens of residues 293, 297, 345, and 348, a<br />

carboxylate oxygen of Asp324, and a water molecule in N2 and type B neuraminidases. In N9, the mainchain<br />

oxygen at 343 is replaced <strong>by</strong> a water molecule. Calcium has been shown to be necessary for<br />

neuraminidase activity [58] and this site is connected to the active site via conserved residues; however,<br />

the functional role of calcium in the structure is unknown, although it has been shown that calcium is<br />

essential for the thermostability of the molecule [59].<br />

C. Carbohydrate <strong>Structure</strong><br />

Carbohydrate at four N-linked glycosylation sites were observed in N2 neuraminidase at residues 86,<br />

146, 200, and 234 in the x-ray structure. Two N-acetylglucosamines were resolved at Asp86 and<br />

Asp234, both at the bottom surface of the monomer. The carbohydrate at Asp200 consists of eight sugar<br />

residues with linkages consistent with known mannose-rich simple N-linked carbohydrates [60]. This<br />

oligosaccharide emerges from the side of the monomer and covers a neighboring subunit (see Figure 4).<br />

The oligosaccharide site at Asn146 is the most conserved of all neuraminidase glucosylation sites,<br />

except that of the neurovirulent virus A/WSN/33 [61]. The absence of this glycosylation site in<br />

A/WSN/33 has been shown to confer neurovirulence in mice [62]. It is a complex sugar containing Nacetylgalactosamine<br />

[63] that is not found in any other of the known oligosaccharides of influenza virus<br />

glycoproteins and is the only glycopeptide antigenically related to chick embryo “host antigen”<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_467.html [4/9/2004 12:16:34 AM]


Document<br />

Figure 4<br />

A CPK model of two tetramers of A/Tokyo/3/67 neuraminidase heads as seen<br />

in the crystal structure. The four carbohydrate spikes emanating from the top<br />

of the molecule (at Asn146) interdigitate and form an open “barrel” structure.<br />

They form an unusual crystal contact. The dark spheres represent carbohydrate<br />

and the light spheres represent protein. The carbohydrate at Asn200 starts<br />

from one subunit and covers a neighboring subunit on the same tetramer. The<br />

other carbohydrates lie on the<br />

underside of the tetramer.<br />

Page 468<br />

[63,64]. The oligosaccharide appears as a spike emanating from the top of the monomer, forming a<br />

crystal contact with a neighboring tetramer in crystals of A/Tokyo/3/67 neuraminidase. The four<br />

carbohydrate spikes of a tetramer form an open “barrel” structure of eight carbohydrate chains with the<br />

neighboring tetramer, with no apparent intercarbohydrate contacts. This oligosaccharide may play an<br />

important but as yet unidentified role in neuraminidase structure or activity.<br />

D. Antigenic Variation in Neuraminidase <strong>Structure</strong>s<br />

Comparison of all known sequences of approximately 390 residues of the neuraminidase globular head<br />

[24], indicates that only 54 (excluding 16 conserved cysteine residues) are invariant (Figure 5a). Apart<br />

from 21 residues involved<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_468.html (1 of 2) [4/9/2004 12:17:07 AM]


Document<br />

Figure 5<br />

(a) Stereo image of a CPK atomic model of an influenza virus neuraminidase<br />

tetrameric head viewed distal to the viral membrane. The darker-shaded atoms<br />

represent totally conserved residues that for the most part form the enzyme<br />

active-site pocket. The lighter shaded atoms represent strain-variable<br />

residues and carbohydrate. (b) A stereo image of the enzyme active-site<br />

pocket of a subunit of neuraminidase with the same shading scheme.<br />

Page 469<br />

in preserving the structural integrity of the molecule [46], the main clustering of these invariant residues<br />

is within the enzyme active site (Figure 5b), where 17 are in the active site and 16 are neighboring the<br />

active site. This is a cavity on the upper surface of the molecule into which sialic acid has been observed<br />

to bind [65,66,50]. Excluding the active-site pocket, strain variation occurs over the entire surface of the<br />

neuraminidase heads. The active site was found to be in<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_469.html (1 of 2) [4/9/2004 12:21:29 AM]


Document<br />

Figure 6<br />

Stereo image of a ball-and-stick model of sialic acid bound to the active site of<br />

Tern/N9 neuraminidase. The hydrogen-bond interactions with conserved residues<br />

are shown as dotted lines. Nitrogen atoms are shaded black, oxygen atoms<br />

are shaded dark gray, and carbon atom are shaded light gray.<br />

Page 470<br />

a pocket of totally conserved (over all animal subtypes) residues [65]. In this way the enzyme active-site<br />

pocket is surrounded <strong>by</strong> highly variable surface residues that prevent immune recognition of the active<br />

site <strong>by</strong> antibody molecules [25]. The x-ray diffraction studies of neuraminidase—antibody complexes<br />

have shown that the footprint of an antibody in the complex is larger than the exposed surface of the<br />

conserved region of the active site [67–69]. These structural results indicate that antibodies are unable to<br />

exert mutational pressure on the conserved active site because they cannot bind there without engaging<br />

strain-variable residues as part of the binding surface. However, as antibodies bind to strain-variable<br />

elements of the structure, the virus can overcome host immune pressure <strong>by</strong> point mutations of the<br />

residues that do not have a catalytic or structural role [70,48] but are able to disrupt the<br />

antigen—antibody binding interface. The rapid emergence of these escape mutants explains the failure<br />

in producing a universal vaccine for influenza.<br />

E. The Enzyme Active Site<br />

The structures of N-acetyl neuraminic acid (sialic acid Neu5Ac) and the 2-deoxy-2,3-dehydro-N-acetyl<br />

neuraminic acid (Neu5Ac2en) inhibitor complexed with N2 neuraminidase [66] revealed the nature of<br />

the interactions of the<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_470.html [4/9/2004 12:22:22 AM]


Document<br />

Page 471<br />

molecules in the active-site pocket (Figure 6). Sialic acid binds in the active site in the α-anomer and, in<br />

a distorted half-chair conformation, through the same face as used in its interaction with hemagglutinin<br />

[71]. The carboxylate group of the sugar interacts with three guanidinium groups of argine residues 118,<br />

292, and 371 and has an equatorial conformation with respect to the sugar ring (this group, axial toward<br />

the floor in the undistorted structure, is probably held equatorial <strong>by</strong> interactions with these arginine<br />

residues). The NH group of the 5-N-acetyl side chain interacts with the floor of the active-site cavity via<br />

a bound water molecule. The oxygen of the 5-N-acetyl side chain is hydrogen bonded to the N ε of Arg<br />

152, while the methyl group lies in a hydrophobic pocket near Ile222 and Trp178. The last two hydroxyl<br />

groups of the 6-glycerol side chain are hydrogen bonded to carboxylate oxygens of Glu276 and the 4hydroxyl<br />

is directed to a carboxylate oxygen of Glu119. The glycosidic oxygen O2 interacts with a<br />

carboxylate oxygen of Asp151. Similar binding of sialic acid in the active site was observed in type B<br />

virus [50].<br />

Comparison of active sites of N2, N9, and type B neuraminidase [72] show there are no significant<br />

differences between active-site orientations, except for some minor displacements of Arg224 and<br />

Glu276, where the major interactions with the 6-glycerol group of sialic acid occur. However there are<br />

differences in the water structure in the active sites of the different subtypes. The Gly405 residue (in N2<br />

and N9) is replaced <strong>by</strong> a tryptophan in type B, which displaces four water molecules that lie in a solvent<br />

pocket bounded <strong>by</strong> arginines at residues 371 and 118. The Val240 residue (in N2 and N9) is replaced <strong>by</strong><br />

a methionine in type B, which displaces two water molecules that form a channel under Arg 224,<br />

decreasing the flexibility of the active site of type B in this region. These waters are not displaced in the<br />

sialic acid/neuraminidase complex in N9 and N2 and would alter the hydrogen-bonding pattern of the<br />

complexes when compared to type B.<br />

Comparison of the active sites of influenza neuraminidases and bacterial sialidases [51,52] indicates that<br />

there is considerable conservation of the catalytic site at the carboxylate-binding end. The residues<br />

Asp151, Arg118, Glu277, Arg292, Val or Ile349, Arg371, Tyr406, and Glu425 are conserved over all<br />

known viral and bacterial strains. The arginyl residues 118, 292, and 371 position the 2-carboxylate<br />

group and the Val(or Ile)349, Glu425, and Glu277 are important in positioning the triarginyl cluster. The<br />

residues Asp151 and Tyr406 are presumably important in bond cleavage, but the precise mechanism is<br />

still unclear. These eight residues (Figure 7) are thus most likely to be conserved in all neuraminidases.<br />

Differences between viral, bacterial, and mammalian neuraminidase structures may correspond with the<br />

different role these enzymes have in vivo. These differences are likely to be in the interactions of the 6glycerol,<br />

5-N-acetyl, and 4-OH groups of silaic acid. In the influenza virus, the turnover rate must be<br />

balanced against the requirement to maintain<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_471.html [4/9/2004 12:22:26 AM]


Document<br />

Figure 7<br />

Stereo image of the active site of Tern/N9 showing the active-site residues<br />

surrounding 4-guanidino-Neu5Ac2en (black). Those residues that are conserved<br />

in both influenza and bacterial neuraminidase are shaded gray, and those that are<br />

conserved only in influenza virus neuraminidase are not shaded.<br />

sufficient sialic acid at the cell surface to enable attachment via the hemagglutinin. This balance may<br />

require some configuration of residues in the active site not directly responsible for catalysis but only<br />

involved in the binding and release of sialic acid.<br />

IV. Neuraminidase Inhibitor <strong>Design</strong><br />

Page 472<br />

Earlier screening programs [73] failed to identify potent inhibitors of viral neuraminidase. The first<br />

inhibitor synthesized [74] with a K i value in the micromolar range was Neu5Ac2en. This was <strong>based</strong> on<br />

the proposed transition state of the reaction, where the anomeric carbon (C2) bound to the ketosidic<br />

oxygen has a trigonal state. Several analogues of Neu5Ac2en were synthesized soon after, and the most<br />

potent of these, a trifluoracetyl derivative, had a K i of only 0.8 μM [75]. While this compound showed<br />

that in cell culture it retarded virus shedding [34,76], it failed as an effective antiviral agent in animals<br />

[77].<br />

The x-ray structure of Neu5Ac2en/neuraminidase complexes have been determined for N2 [66], type B<br />

[78], and N9 [79]. The Neu5Ac2en molecule binds in the active site of neuraminidase with the<br />

carboxylate oxygen atoms placed in the same location as the carboxylate of sialic acid. The 5-N-acetyl, 4-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_472.html [4/9/2004 12:22:48 AM]


Document<br />

hydroxy, and 6 Neu5Ac2en -glycerol are positioned isosterically in the two molecules.<br />

Page 473<br />

An alternative approach to develop sialidase inhibitors has been made using synthetic thioglycoside<br />

analogs of gangliosides such as Neu5Aca(2-S-6)Glcb(1-1)Ceramide [80] were shown to inhibit different<br />

subtypes of human and animal influenza virus with K i values of up to 2.8 μM. These metabolically<br />

stable ganglioside analogs contain a thioglycosidic linkage to the terminal neuraminic acid that resists<br />

cleavage <strong>by</strong> the enzyme. Several flavonoid neuraminidase inhibitors have been isolated from plant<br />

extracts [81], one of which 5,7,4'-trihydroxy-8-methoxyflavone, was a more potent inhibitor than<br />

Neu5Ac2en. Recently, in vivo anti-influenza virus activity of a Kampo (Japanese herbal medicine)<br />

preparation has shown promising results in inhibiting influenza virus replication in mice [82], but the<br />

mode of action of these compounds is unclear.<br />

A. Enzyme Mechanism<br />

The similar positioning of the carboxylate oxygens and ring of Neu5Ac2en and sialic acid suggests that<br />

Neu5Ac2en is probably a transition state analog. As Neu5Ac2en has a higher affinity for neuraminidase<br />

than sialic acid, a mechanism of catalysis was proposed [83] that involves the distortion of the substrate<br />

<strong>by</strong> the formation of a oxycarbonium ion intermediate, which has a similar structure to Neu5Ac2en.<br />

However the structural results from sialic acid/neuraminidase complexes suggest that the tighter binding<br />

of Neu5Ac2en more likely comes from the relaxation of the conformational strain arising from the<br />

transition from chair to boat of the pyranose ring of sialic acid in the active site [66].<br />

Evidence for a sialyl cation transition state <strong>by</strong> isotopic effects [84] support the existence of a<br />

oxycarbonium ion intermediate. However the structural basis for neuraminidase activity is still unclear.<br />

It has been suggested [66] that the tyrosyl oxygen of Tyr406, assisted <strong>by</strong> the sialic acid carboxylate<br />

itself, could stabilize the developing charge on the oxycarbonium ion intermediate. The reaction would<br />

be completed <strong>by</strong> the activation of a water molecule <strong>by</strong> a deprotonated Asp151 and its attack on the<br />

carbonium, resulting in the formation of the α-anomer of sialic acid [85]. However the pH-activity<br />

profile of neuraminidase [86,87,78] suggests a bell-shaped profile, which indicates normal activity from<br />

a pH range of about 4.5 to 9. This would indicate that the role of Asp151 as the acid group in the<br />

catalysis is unclear. A nonspecific proton donor has been proposed [78], probably a water molecule as<br />

the acid group, with a deprotonated Tyr406 stabilizing the oxycarbonium ion, and a proton transferred<br />

from water to the departing aglycon group. It has also been postulated that a proton is eliminated at C3<br />

leading to the transformation<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_473.html [4/9/2004 12:23:02 AM]


Document<br />

Page 474<br />

of the oxycarbonium ion into Neu5Ac2en, which is produced irreversibly at low levels from sialic acid<br />

<strong>by</strong> the enzyme [78]. An S N1-type mechanism has been suggested [88] that is facilitated <strong>by</strong> an activated<br />

water molecule, which can be expelled upon inhibitor binding. The catalytic mechanism could possibly<br />

proceed without an acid group: the electrostatic potential of the enzyme could lower the barrier<br />

preventing the breaking of the ketosidic bond and the solvent could protonate the aglycon after release<br />

[89]. Clearly details of the enzyme mechanism have yet to be elucidated definitively. Structural<br />

considerations indicate that only Tyr406 (and possibly Glu277) and the triarginyl cluster are essential in<br />

the enzyme mechanism and that Asp151 is implicated.<br />

B. Inhibitor <strong>Design</strong> Principles<br />

All the nearest-neighbor interactions between sialic acid or Neu5Ac2en and the protein are with totally<br />

conserved amino acids. Thus an inhibitor designed to bind only to the conserved active-site residues of<br />

neuraminidase would inhibit neuraminidase activity across all strains of influenza. This would enable<br />

the development of an antiviral drug that would affect the spread of viral replication potentially in three<br />

ways, i.e., transport through the protective mucosal layer, desialyation of freshly synthesized viral<br />

glycoproteins, and elution of progeny virions from infected cells.<br />

The development of potent inhibitors was <strong>based</strong> on the structural information of the N2 neuraminidase<br />

conserved active site and its complex with sialic acid and Neu5Ac2en [66]. There are no reports of de<br />

novo molecules designed to fit into the cavity, and the most useful approach was to consider molecules<br />

that were structurally related to Neu5Ac2en. This involved the design of molecules that would bind<br />

isosterically to Neu5Ac2en but which were modified to increase the number of favorable interaction<br />

with the protein. The method of Goodford [90] enabled the calculation of favorable binding sites for a<br />

variety of chemical probes. The validity of this method was indicated <strong>by</strong> its ability to identify the<br />

positions of the carboxylate binding site of sialic acid as an energy minima for a carboxylate probe<br />

(Figure 8a), and the successful prediction of known bound-water sites in the active sites <strong>by</strong> a water<br />

probe. Utilizing this methodology, predictions of energetically favorable substitutions to Neu5Ac2en<br />

were examined [91]. A replacement of the hydroxyl at the 4-position of the pyranose ring of Neu5Ac2en<br />

<strong>by</strong> an amino group was identified <strong>by</strong> this procedure as an energetically favorable substitution. A<br />

protonated primary amine probe identified a favorable binding site of -16 kcal mol -1 at this location and<br />

in a pocket in the active site near two conserved glutamate residues Glu119 and Glu227 (Figure 8b).<br />

This suggested that the substitution of the 4-hydroxyl group <strong>by</strong> an amino group would increase the<br />

overall binding interactions <strong>by</strong> forming a salt link with Glu119. Furthermore the substitution of the 4hydroxyl<br />

group with a much bulkier guanidinyl group would lead to even tighter binding as a result<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_474.html [4/9/2004 12:23:05 AM]


Document<br />

Figure 8<br />

Stereo image of the residues in the active site of Tern/N9 complexed with<br />

4-guanidino—Neu5Ac2en overlayed with GRID maps [90] (caged mesh contours)<br />

for (a) a carboxylate oxygen probe, contoured at -12 kcal/mol; (b) an amino<br />

nitrogen probe, contoured at -12 kcal/mol.<br />

of lateral interactions of the terminal nitrogens of the guanidinyl group and the carboxylate groups of<br />

Glu119 and Glu227.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_475.html (1 of 2) [4/9/2004 12:25:17 AM]<br />

Page 475


Document<br />

This led to the design and syntheses of 4-amino-Neu5Ac2en and 4-guanidino-Neu4Ac2en [91] which<br />

bound to A/Tokyo/3/67 with a K i of 50 nM and 0.2<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_475.html (2 of 2) [4/9/2004 12:25:17 AM]


Document<br />

Page 476<br />

nM, respectively. These compounds were later shown <strong>by</strong> x-ray studies of 4-guanidino and 4-amino-<br />

Neu4Ac2en complexed with A/Tokyo/3/67 neuraminidase to bind close to that predicted <strong>by</strong> the design<br />

studies. However details of the interactions of the guanidinyl group of the 4-guanidino-Neu4Ac2en with<br />

the glutamic acid groups (Glu119 and Glu277) in the floor of the active site were slightly different. This<br />

was confirmed on a higher resolution x-ray study (Figure 9) of a 4-guanidino-Neu4Ac2en complexed<br />

with Tern N9 neuraminidase [72]. One of the primary guanidinyl nitrogens of 4-guanidino-Neu4Ac2en<br />

is hydrogen bonded to the main-chain oxygen at residue 178, a carboxylate oxygen of Glu227, and a<br />

water molecule. The other primary guanidinyl nitrogen interacts with the main-chain oxygen of residues<br />

178 and 151. The secondary guanidinyl nitrogen interacts with the carboxylate of Glu119 and Asp151.<br />

The interactions with Glu119 are electrostatic and van der Waals in character and lack hydrogenbonding<br />

geometry (postulated in the design study) as the carboxylate group of Glu119 stacks parallel to<br />

the guanidinyl group. Furthermore, theoretical energy-minimized structures of the complex using<br />

AMBER [92] converged to the x-ray structure only if the protein nonhydrogen atoms were kept rigid in<br />

the x-ray structure [72]. Otherwise this resulted in active site residues showing large distortions in their<br />

conformation. This is an example of the difficulty in correctly modeling even modest changes in the<br />

interactions of an inhibitor/active site complex.<br />

The 4-guanidino analog shows potent inhibition of neuraminidase activity in all known wild strains of<br />

influenza. Furthermore it is very specific to in-<br />

Figure 9<br />

Stereo image of the Tern/N9 4-guanidino-Neu5Ac2en complex showing the<br />

hydrogen-bond interactions (dotted lines) of the inhibitor with conserved residues<br />

in the active site of the enzyme. Nitrogen, oxygen, and carbon atoms are<br />

shaded black, dark gray, and light gray, respectively.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_476.html [4/9/2004 12:26:21 AM]


Document<br />

Page 477<br />

fluenza, as it shows weak inhibition to bacterial, para influenza, and mammalian neuraminidases<br />

[91,93]. This is possibly due to the specific interactions of the 4-guanidino group within the subpocket<br />

of the active site of influenza neuraminidase that is not conserved in other neuraminidases. In bacterial<br />

neuraminidases [51] this pocket is much smaller, and would prevent the binding of the 4-guanidino<br />

group in this region of the active site. This is consistent with the proposition that the interactions of<br />

sialic acid with the active site of neuraminidases are function specific at the C4, C5, and C6 position of<br />

sialic acid and that modification at these positions confer specificity to the target enzyme [72]. Other<br />

approaches [94] to structure-<strong>based</strong> design of inhibitors have to date produced only millimolar inhibition<br />

of neuraminidase activity.<br />

V. Antiviral Activity<br />

In vitro inhibition of viral replication in tissue culture was demonstrated earlier [34] for the trifluro<br />

derivative of Neu5Ac2en, but its antiviral activity in vivo was not demonstrated [77]. As a consequence<br />

of this, efforts were directed towards hemagglutinin, which was then considered a better target for<br />

antiinfluenza drugs. The interest in neuraminidase inhibitors as anti-influenza drugs has only been<br />

revived with the success of 4-guanidino-Neu5Ac2en and its analogs in attenuating viral titer in mice<br />

when administered directly into the lungs [91,95].<br />

A. Inhibition In Vitro<br />

Von Itzstein and co-workers [91] have shown that the 4-amino-and 4-guanidino-Neu5Ac2en inhibit<br />

influenza strains A/Singapore/1/57 and B/Victoria/102/95 in MDCK cells with IC 50 values (the<br />

concentration required to inhibit plaque formation in MDCK cells <strong>by</strong> 50%) of 1.5 mM and 0.065 mM (4amino)<br />

and 0.014 mM and 0.005 mM (4-guanidino) respectively. These IC 50 values, in particular for the<br />

4-guanidino compound, are well below those found for amantadine, ribovarin, and Neu5Ac2en.<br />

Furthermore in comparison to Neu5Ac2en, the 4-guanidino-Neu5Ac2en inhibitor was 100-fold less<br />

active against human lysosomal sialidase and over 1000-fold more active against a wide range of<br />

clinical isolates of influenza A and B, including amantidine and rimantadine resistant variants [93].<br />

The inhibition of virus replication in MDCK cells has been confirmed [96] and this knowledge<br />

prompted the extension of the inhibition studies to human respiratory epithelium cells in vitro [97],<br />

which indicated high antiviral activity for strains of A(HINI) and A(N3N2) isolates. They also found<br />

that delayed administration of the drug—after viral replication was well estab-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_477.html [4/9/2004 12:26:55 AM]


Document<br />

Page 478<br />

lished—was associated with inhibition of virus replication. However viral titer was higher (1.3 log 10<br />

compared to 4.0 log 10 at 10 mg/mL concentration of the drug) for the delayed administration compared<br />

with viral titer when the drug was present throughout the period of viral exposure. The clinical<br />

significance of this in the treatment of established infections has yet to be explored in detail, although<br />

preliminary clinical trails [98] have indicated some positive results.<br />

B. Administration and Inhibition In Vivo<br />

The earlier work of Palase and Schulman [77] indicated that the failure in inhibiting viral replication in<br />

mice after intranasal and subcutaneous treatment with Neu5aC2en would also occur for Neu5Ac2en<br />

analogs. This failure of Neu5Ac2en as an antiviral treatment in animal models can now be ascribed to<br />

the rapid excretion of the compound [99] there<strong>by</strong> not delivering sufficient concentration of the inhibitor<br />

to the infected tissue. It had been shown [91] that the antiviral activity in mice is considerably more,<br />

when the 4-guanidino compound is administered intranasally than when the drug is injected<br />

intraperitoneally. This can be attributed to the localization of sufficiently high concentration of inhibitor<br />

in the lining of the nasal and respiratory epithelia where influenza virus replication is believed to occur.<br />

Animal trials with ferrets challenged with influenza virus have shown the 4-guanidino Neu5Ac2en<br />

compound is effective in studies [91] involving prophylactic administration of the drug. When the drug<br />

is administrated intranasally, 50 μg/kg, twice daily, one day before infection with the virus and the<br />

succeeding six days, it substantially reduces virus titer in nasal washing and abolishes fever that usually<br />

appears 3 days after infection.<br />

The drug is currently undergoing clinical trials, and the initial results from a double blind, randomized,<br />

placebo-controlled trial using this compound have been positive both for early treatment and<br />

prophylaxis of experimental inoculation of human volunteers [98] with influenza A/Texas/91.<br />

C. <strong>Drug</strong> Resistance<br />

Although the active site of influenza virus has been conserved in all known field strains of the virus, the<br />

possibility of drug resistance needs to be addressed. Experience with influenza and other viruses, in<br />

particular HIV, have shown [100] that drug-resistant mutants arise very rapidly, resulting in the<br />

effectiveness of antiviral drugs being short lived. One attempted solution to the problem is the use of<br />

several drugs during therapy [101], making it more difficult for the virus the develop resistance.<br />

In the case of influenza virus, there have to date been no reports of drug resistance from field strains.<br />

However it has recently been reported [102,103]<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_478.html [4/9/2004 12:27:14 AM]


Document<br />

Page 479<br />

that 4-guanidino-Neu5Ac2en-resistant mutants can arise from multiple serial passages of virus in<br />

MDCK cells in the presence of the inhibitor. It has been demonstrated [104] that almost all of the<br />

mutations arise in the hemagglutinin receptor binding site and not on the neuraminidase. These altered<br />

HA variants, which have weaker binding to HA receptors, appear to arise as a result of increased<br />

inhibition of the neuraminidase <strong>by</strong> the drug. This is consistent with the earlier proposition that the rate of<br />

desialylation of receptor is critically related to the rate of attachment to receptor, for the virus infection<br />

and elution. The decreased activity of the neuraminidase <strong>by</strong> the drug selects HA mutants with decreased<br />

binding to receptors.<br />

However a drug-resistant neuraminidase mutation has been isolated [103,105] that results in a single<br />

active-site residue mutation—with apparently unaltered activity—of glutamic acid 119 to glycine. The<br />

crystal structure of this mutant and its complex (Figure 10) with 4-guanidino-Neu5Ac2en has been<br />

determined [105] and the structure suggests that the decrease in inhibitor binding arises from the loss of<br />

stabilizing interaction with the 4-guanidino group of the drug [72] and alterations in the solvent structure<br />

of the active site. This alteration arises from a water molecule that binds near the location of one of the<br />

carboxylate oxygens of the glutamic acid in the wild type molecule. The location of the 4-guanidino-<br />

Neu5Ac2en drug in the complex with the mutant enzyme is isosteric compared to the drug/wildtype<br />

complex [72]. The only differences are the interactions with residue 119.<br />

Figure 10<br />

Stereo image of the Glu119Gly Tern/N9 mutant complexed with<br />

4-guanidino-Neu5Ac2en. The inhibitor binds in a similar conformation as with<br />

the wild type enzyme. The atom labeled “Wat” represents the water<br />

molecule found in the mutant enzyme that assumes the role of the glutamic acid<br />

in the wild type enzyme. Nitrogen, oxygen, and carbon atoms are shaded<br />

black, dark gray and light gray, respectively.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_479.html (1 of 2) [4/9/2004 12:27:37 AM]


Document<br />

VI. Conclusion<br />

Page 480<br />

It has now been over a decade since the structure of influenza virus neuraminidase was determined<br />

[45,65]. The development of the 4-substituted Neu5ac2en analog inhibitors dates from 1987 when the<br />

structure of sialic acid and Neu5Ac2en complexed with neuraminidase were determined to sufficient<br />

accuracy to permit modeling of potential inhibitors [66]. The development was a multidisciplinary<br />

collaboration of biochemists, crystallographers, molecular modelers, and synthetic chemists and<br />

culminated in the synthesis [106] and biological testing of the compounds [91]. Preliminary data on the<br />

efficacy of these drugs on humans indicate its effectiveness in both prophylaxis and treatment of the<br />

disease [98]. The 4-guanidino compound is in Phase 2 trials (October, 1995). While drug resistance of<br />

the virus has been observed in vitro, it will be interesting to see if variants arise in animal studies as they<br />

do for Amantidine. This is one of the first rationally designed antiviral drugs to be synthesized and<br />

portends well for this methodology to be used as an additional weapon in controling the many pathogens<br />

that have plagued humanity for so long.<br />

Acknowledgments<br />

The author would like to acknowledge Peter Colman, my collaborator in neuraminidase crystallography,<br />

Mike Lawrence for the GRID maps shown here and reading this manuscript, Brian Smith for discussions<br />

on enzyme mechanisms, Jenny McKimm-Breschkin for discussions on drug resistance, Bert van<br />

Donkelaar for technical support, and Paul Davis for computing support.<br />

References<br />

1. Kilbourne ED. Influenza. New York: Plenum, 1987.<br />

2. Webster RG, Schafer JR, Suss J, Bean Jr WJ, Kawaoko Y. Evolution and ecology of influenza<br />

viruses. In: Hannoun C, et al. eds. Options for the Control of Influenza Virus II. Amsterdam: Excerpta<br />

Medica, 1993:177–185.<br />

3. Lamb RA. Genes and proteins of the influenza virus. In: Krug RM, ed. The Influenza Viruses, ed.<br />

New York: Plenum, 1989: 1–87.<br />

4. Murphy JS, Bang FB. Observations with the electron microscope on cells of chick chorio-allantoic<br />

membrane infected with influenza virus. J Exp Med 1952; 95:259.<br />

5. Compans RW, Dimmock NJ. An electron microscopic study of single-cycle infection of chick<br />

embryo fibroblasts <strong>by</strong> influenza virus. Virology 1969; 39:499.<br />

6. Murphy BR, Webster RG. Orthomyxovirus. In: Fields BN, Knipe DM, eds. Virology. New York:<br />

Raven Press, 1990:1091–1152.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_480.html (1 of 2) [4/9/2004 12:27:51 AM]


Document<br />

7. Ennis FA, Meager A. Immune interferon produced to high levels antigenic stimulation of human<br />

lymphocytes with influenza virus. J Exp Med 1981; 154:1279–1289.<br />

Page 481<br />

8. Sprenger MJW, Beyer WEP, Kempen BM, Mulder PGH, Masurel N. Risk factors for influenza<br />

mortality. In: Hannoun C, et al. eds. Options for the Control of Influenza Virus II. Amsterdam: Excerpta<br />

Medica, 1993:15–23.<br />

9. Smith FI, Palase P. Variation in influenza virus genes. Epidemiological, pathogenic, and evolutionary<br />

consequences. In: Krug RM, ed. The Influenza Viruses. New York: Plenum, 1989:319–359.<br />

10. Skehel JJ. The origin of pandemic influenza virus. Symp Soc Gen Microbiol 1974; 24:321–342.<br />

11. Webster RG, Laver WG. Antigenic variation in influenza viruses. In: Kilbourne ED, ed. Influenza<br />

Virus and Influenza. New York: Academic, 1975:269–314.<br />

12. Smith W, Andrewes CH, Laidlaw PP. A virus obtained form influenza patients. Lancet 1933;<br />

2:66–68.<br />

13. Beveridge WI. Influenza: the last great plague. London: Heinemann, 1977.<br />

14. Daniels RS, Downie JC, Hay AJ, Knossow M, Skehel JJ, Wang ML, Wiley DC. Fusion mutants of<br />

the influenza virus haemagglutinin glycoprotein. Cell 1985; 40:431–439.<br />

15. Hay AJ, Thompson CA, Geraghty A, Hayhurst S, Grambas S, Bennett MS. The role of the M2<br />

protein in influenza virus infection. In: Hannoun C, et al. eds. Options for the Control of Influenza Virus<br />

II. Amsterdam: Excerpta Medica, 1993:281–288.<br />

16. Pinto LH, Holsinger LJ, Lamb RA, Influenza virus M2 protein has ion channel activity. Cell 1992;<br />

69:517–528.<br />

17. Hayden FG. Update on antiviral agents and viral drig resistance. In: Mandell GL, Douglas RG,<br />

Bennett JE, eds. Principles and Practice of Infectious Disease. Vol. 2. 2d ed. New York: Churchill<br />

Livingston. 1993:3–15.<br />

18. Tyrrell DAJ. Influenza vaccines. Phil Trans R Soc Lond 1980; B288:449–460.<br />

19. Fox JP, Cooney MK, Hall CE, Foy HM. Influenza virus infections in Seattle families, 1975–1979. II.<br />

Pattern of infection of invaded households and relation of age and prior antibody to occurrence of<br />

infection <strong>by</strong> time and age. Am J Epidemiol 1982; 116:228–242.<br />

20. Chakraverty et al. Influenza Activity—United States and worldwide, and composition of the<br />

1993–94 influenza vaccine. JAMA 1993; 269:1778–1779.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_481.html (1 of 2) [4/9/2004 12:28:23 AM]


Document<br />

21. Both GW, Sleigh MJ, Cox NJ, Kendal AP. Antigenic drift in influenza virus H3 haemagglutinin<br />

from 1968 to 1980: multiple evolutionary pathways and sequential amino acid changes at key antigenic<br />

sites. J Virol 1983; 48:52–60.<br />

22. Donnelly JJ, Friedman A, Martinez D, Montgomery DL, Shiver JW, Motzel SL, Ulmer JB, Liu MA.<br />

Preclinical efficacy of a prototype DNA vaccine: enhanced protection against antigenic drift in influenza<br />

virus. Nature Med 1995; 1:583–586.<br />

23. Hirst GK. Adsorption of influenza hemagglutinins and virus <strong>by</strong> red blood cells. J Exp Med 1942;<br />

76:1740–1743.<br />

24. Colman PM, Influenza virus neuraminidase: Krug RM, ed. Enzyme and antigen. In: The Influenza<br />

Viruses. New York: Plenum, 1989:175–218.<br />

25. Colman PM. Influenza virus neuraminidase: structure, antibodies, and inhibitors. Protein Science<br />

1994; 3:1687–1696.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_481.html (2 of 2) [4/9/2004 12:28:23 AM]


Document<br />

26. Klenk E, Faillard H, Lempfrid H. Uber die enzymatishe Wirkung von Ifluenza. Z Physiol Chem<br />

1955; 301:235–246.<br />

Page 482<br />

27. Gottschalk A. Neuraminidase: the specific enzyme of influenza virus and vibrio cholerae. Biochem<br />

Biophys Acta 1957; 23:645–646.<br />

28. Rogers GN, Paulson JC. Receptor determinants of human and animal influenza virus isolates:<br />

difference in receptor specificity of the H3 haemagglutinin <strong>based</strong> on species of origin. Viroloy 1983;<br />

127:361–373.<br />

29. Burnett FM. Mucins and mucoids in relation to influenza virus action. IV. Inhibition <strong>by</strong> purified<br />

mucoid of infection and haemagglutinin with the virus strain WSE. Aust J Exp Biol Med Sci 1947;<br />

26:381–387.<br />

30. Gottschalk A. Neuraminidase: its substrate and mode of action. Adv Enzymol 1958; 20:135–145.<br />

31. Liu C, Air GM. Selection and characterization of a neuraminidase-minus mutant of influenza virus<br />

and its rescue <strong>by</strong> cloned neuraminidase genes. Virol 1993; 194:403–407.<br />

32. Liu C, Eichelberger MC, Compans RW, Air GM. Influenza type A virus neuraminidase does not<br />

play a role in viral entry, replication, assembly, or budding. J Virol 1995; 69:1099–1106.<br />

33. Palase P, Tobita K, Ueda M, Compans RW. Characterization of temperature sensitive influenza<br />

virus mutants defective in neuraminidase. Virology 1974; 61:397–410.<br />

34. Palase P, Compans RW. Inhibition of influenza virus replication in tissue culture 2-deoxy-2,3dehydro-N-trifluroacetylneuraminic<br />

acid (FANA): mechanism of action. J Gen Virol 1976; 33:159–163.<br />

35. Griffin JA, Compans RW. Effect of cytochalsin B on the maturation of enveloped viruses. J Exp<br />

Med 1979; 150:379–391.<br />

36. Bucher DJ, Palase P. The biologically active proteins of influenza virus neuraminidase. In:<br />

Kilbourne ED, ed. Influenza virus and influenza. New York: Academic, 1975:83–123.<br />

37. White DO. Influenza viral proteins: identification and synthesis. Curr Top Microbiol Immunol 1974;<br />

63:1–48.<br />

38. Laver WG. The structure of influenza virus 3. Disruption of the virus particle and separation of<br />

neuraminidase activity. Virology 1963; 20:251–262.<br />

39. Laver WG, Valentine RC. Morphology of the isolated hemagglutinin and neuraminidase subunits of<br />

influenza virus. Virology 1969; 38:105–119.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_482.html (1 of 3) [4/9/2004 12:28:41 AM]


Document<br />

40. Drzenick R, Frank H, Rott R. Electron microscopy of purified influenza virus neuraminidase.<br />

Virology 1968; 36:703–707.<br />

41. Laver WG. Crystallization and peptide maps of neuraminidase “heads” from H2N2 and H3N2<br />

influenza virus strains. Virology 1978; 86:78–87.<br />

43. Colman PM, Laver WG. The structure of influenza virus neuraminidase at 5A resolution. In:<br />

Structural aspects of recognition and assembly in biological macromolecules. I.S.S. Rehovot,<br />

1981:869–872.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_482.html (2 of 3) [4/9/2004 12:28:41 AM]<br />

42. Blok J,<br />

Air GM,<br />

Laver WG,<br />

Ward CW,<br />

Lilley GG,<br />

Woods EF,<br />

Roxburgh<br />

CM, Inglis<br />

AS. Studies<br />

on the size,<br />

chemical<br />

composition,<br />

and partial<br />

sequence of<br />

the<br />

neuraminidase<br />

(NA) from<br />

type A<br />

influenza<br />

virus show<br />

that the Nterminal<br />

region of the<br />

NA is not<br />

processed and<br />

serves to<br />

anchor the<br />

NA in the<br />

viral<br />

membrane.<br />

Virology<br />

1982;<br />

119:109–121.


Document<br />

44. Wright CE, Laver WG. Preliminary crystallographic data for influenza virus neuraminidase “heads”.<br />

J Mol Biol 1978; 120:133–136.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_482.html (3 of 3) [4/9/2004 12:28:41 AM]


Document<br />

45. Varghese JN, Laver WG, Colman PM. <strong>Structure</strong> of the influenza virus glycoprotein antigen<br />

neuraminidase at 2.9Å resolution. Nature 1983; 303:35–40.<br />

46. Varghese JN, Colman PM. Three-dimensional structure of the neuraminidase of influenza virus<br />

A/Tokyo/3/67 at 2.2Å resolution. J Mol Biol 1991; 221:473–486.<br />

47. Baker AT, Varghese JN, Laver WG, Air GM, Colman PM. The three-dimensional structure of<br />

neuraminidase of subtype N9 from an avian influenza virus. Proteins 1987; 1:111–117.<br />

Page 483<br />

48. Tulip WG, Varghese JN, Baker AT, van Donkelaar A, Laver WG, Webster RG, Colman PM.<br />

Refined atomic structures of N9 subtype influenza virus neuraminidase and escape mutants. J Mol Biol<br />

1992; 221:487–497.<br />

49. Tulip WG, Varghese JN, Laver WG, Webster RG, Colman PM. Refined crystal structure of the<br />

influenza virus N9 neuraminidase-NC41 Fab complex. J Mol Biol 1992; 227:122–148.<br />

50. Burmeister WP, Ruigrok RWH, Cusack S. The 2.2Å resolution crystal structure of influenza B<br />

neuraminidase and its complex to sialic acid. EMBO J 1991; 11:49–56.<br />

51. Crennell SJ, Garman EF, Laver WG, Vimr ER, Taylor GL. Crystal structure of a bacterial sialidase<br />

(from Salmonella typhimurium LT2) shows the same fold as an influenza virus neuraminidase. Proc<br />

Natl Acad Sci USA 1993; 90:9852–9856.<br />

52. Crennell SJ, Garman E, Laver G, Vimr E, Taylor GL. Crystal structure of Vibro Cholerae<br />

neuraminidase reveals dual lectin-like domains in addition to the catalytic domain. <strong>Structure</strong> 1994;<br />

2:535–544.<br />

53. Vellieux FMD, Huitema F, Groendijk HN, Kalk KH, Frank J, Jonjejan JA, Duine JA, Petratos K,<br />

Drenth J, Hol WGJ. <strong>Structure</strong> of quinoprotein methylamine dehydrogenase at 2.25A resolution. EMBO<br />

J 1989; 8:2171–2178.<br />

54. Ito N, Phillips SEV, Stevens C, Ogel ZB, McPherson MJ, Keen JN, Yadav KDS, Knowles PF.<br />

Novel thioether bond revealed <strong>by</strong> a 1.7A crystal structure of galactose oxidase. Nature 1991; 350:87–90.<br />

55. Xia Z-X, Dia W-W, Xion J-P, Hao Z-P, Davidson VL, White S, Mathews FS. The three dimensional<br />

structure of methonol dehydrogenase from two methylotrophic bacteria at 2.6A resolution. J Biol Chem<br />

1992; 267:22289.<br />

56. Richardson JS. The anatomy and taxonomy of protein structure. Advan Protein Chem 1981;<br />

34:167–339.<br />

57. Wang J, Yan Y, Garrett TPJ, Liu J, Rogers DW, Garlick, Tarr GE, Husain Y, Reinherz EL, Harrison<br />

SC. Atomic structure of a fragment of human CD4 containing two immunogloblin-like domains. Nature<br />

1990; 348:411–418.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_483.html (1 of 2) [4/9/2004 12:28:58 AM]


Document<br />

58. Baker NJ, Gandhi SS. Effect of Ca++ on the stability of influenza virus neuraminidase. Arch Virol<br />

1976; 52:7–18.<br />

59. Burmeister WP, Cusack S, Ruigrok RWH. Calcium is needed for the thermostability of influenza B<br />

virus neuraminidase. J Gen Virol 1994; 75:381–388.87.<br />

60. Wagh PV, Bahl OP. Sugar residues on proteins. Crit Rev Biochem 1981; 307–377.<br />

61. Hitte AL, Nayak DP. Complete nucleotide sequences of the neuraminidase gene of human influenza<br />

virus A/WSN/33. J Virol 1982; 41:730–734.<br />

62. Li S, Schulman J, Itamura S, Palase P. Glycosylation of neuraminidase determines the<br />

neurovirulence of influenza A/WSN/33 virus. J Virol 1993; 67:6667–6673.<br />

63. Ward CW, Murry JM, Roxburgh CM, Jackson DC. Chemical and antigenic characterisation of the<br />

carbohydrate side chains of an Asian (N2) influenza virus neuraminidase. Virology 1983; 126:370–375.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_483.html (2 of 2) [4/9/2004 12:28:58 AM]


Document<br />

64. Ward CW, Elleman TC, Azad AA. Amino acid sequence of the pronase-released heads of<br />

neuraminidase subtype N2 from the Asian strain A/Tokyo/3/67 of influenza virus. Biochem J 1982;<br />

207:91–95.<br />

Page 484<br />

65. Colman PM, Varghese JN, Laver WG. <strong>Structure</strong> of the catalytic and antigenic sites in influenza virus<br />

neuraminidase. Nature 1983; 303:41–44.<br />

66. Varghese JN, McKimm-Breschkin JL, Caldwell JB, Kortt AA, Colman PM. The structure of the<br />

complex between influenza virus neuraminidase and sialic acid, the viral receptor. Proteins 1992;<br />

14:327–332.<br />

67. Colman PM, Laver WG, Varghese JN, Baker AT, Tullock PA, Air GM, Webster RG. Threedimensional<br />

structure of a complex of antibody with influenza virus neuraminidase. Nature 1987;<br />

326:358–363.<br />

68. Colman PM, Tulip WR, Varghese JN, Tulloch PA, Baker AT, Laver WG, Air GM. 3-D structures of<br />

influenza virus neuraminidase-antibody complexes. Proc Roy Soc Lond 1989; B,323:511–518.<br />

69. Mal<strong>by</strong> RL, Tulip WR, Harley VR, McKimm-Breschkin JL, Laver WG, Webster RG, Colman PM.<br />

The structure of a complex between the NC10 antibody and influenza virus neuraminidase and<br />

comparison with the overlapping binding site of the NC41 antibody. <strong>Structure</strong> 1994; 2:733–746.<br />

70. Varghese JN, Webster RG, Laver WG, Colman PM. The three-dimentional structure of an escape<br />

mutant of the neuraminidase of influenza virus A/Tokyo/3/67. J Mol Biol 1988; 200:201–203.<br />

71. Weis W, Brown JH, Cusack S, Paulson JC, Skehel JJ, Wiley DC. <strong>Structure</strong> of the influenza virus<br />

haemagglutinin complexed with its receptor, sialic acid. Nature 1988; 333:426–431.<br />

72. Varghese JN, Epa VC, Colman PM. Three-dimensional structure of 4-guanidino-Neu5Ac2en and<br />

influenza virus neuraminidase. Protein Science 1995; 4:1081–1087.<br />

73. Edmond JD, Johnston RG, Kidd D, Rylance HJ, Sommerville RG. The inhibition of neuraminidase<br />

and antiviral action. Br J Pharmacol Chemother 1966; 27:415–426.<br />

74. Meindl P, Tuppy H. 2-Deoxy-2,3-dehydrosialic acids. I. Synthesis and properties of 2-deoxy-2,3dehydro-N-acetylneuraminic<br />

acids and their methy esters. Monatsh Chem 1969; 100:1295–1306.<br />

75. Meindl P, Bodo G, Palase P, Schulman J, Tuppy H. Inhibition of neuraminidase activity <strong>by</strong><br />

derivatives of 2-deoxy-2,3-dehydro-N-acetylneuraminic. Virology 1974; 58:457–463.<br />

76. Palase P, Schulman JL, Bodo G, Meidl P. Inhibition of influenza and parainfluenza virus replication<br />

in tissue culture <strong>by</strong> 2-deoxy-2,3-dehydro-N-trifluroacetylneuraminic acid (FANA). Virology 1974;<br />

59:490–498.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_484.html (1 of 2) [4/9/2004 12:29:15 AM]


Document<br />

77. Palase P, Schulman JL. Inhibitors of viral neuraminidase as potential antiviral drugs. In:<br />

Chemoprophylaxis and virus infections of the upper respiratory tract. vol. 1. Boca Raton, Florida: CRC<br />

Press, 1977:189–205.<br />

78. Burmeister WP, Henrissat B, Bosso C, Cusack S, Ruigrok RWH. Influenza B virus neuraminidase<br />

can synthesize its own inhibitor. <strong>Structure</strong> 1993; 1:19–26.<br />

79. Bossart-Whitaker P, Carson M, Babu YS, Smith CD, Laver WG, Air GM. Three dimensional<br />

structure of influenza A N9 neuraminidase and its complex with the inhibitor 2-deoxy 2,3-dehydro-Nacetyl<br />

neuraminic acid. J Mol Biol 1993; 232:1069–1083.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_484.html (2 of 2) [4/9/2004 12:29:15 AM]


Document<br />

80. Suzuki Y, Sato K, Kiso M, Hasegawa A. New ganglioside analogs that inhibit influenza virus<br />

sialidase. Glycoconjugate J 1990; 7:349–356.<br />

Page 485<br />

81. Nagai T, Miyaichi Y, Tomimori T, Suzuki Y, Yamada H. Inhibition of influenza virus sialidase and<br />

anti-influenza virus activity <strong>by</strong> plant flavonoids. Chem Pharm Bull 1990; 38(5):1329–1332.<br />

82. Nagai T, Yamada H. In vivo anti-influenza virus activity of Kampo (Japanese herbal) medicine “Shoseiryu-to”<br />

and its mode of action. Int J Immunopharmacol 1994; 16(8):605–613.<br />

83. Miller CA, Wang P, Flashner M. Mechanism of Arthro-bacter sialophilus neuraminidase: the<br />

binding of substrates and transition state anologs. Biochem Biophys Res Commun 1978; 83:1479–1487.<br />

84. Chong AKJ, Pegg MS, Taylor NR, von Itzstein M. Evidence for a sialyl cation transition state<br />

complex in the reaction of sialidase from influenza. Eur J Biochem 1992; 207:335–343.<br />

85. Friebolin H, Brossmer R, Keilich G, Ziegler D, Supp M. 1H-NMR spektroskopischer nachweis der<br />

N-acetyl-a-D-neuraminsaure als primares spaltprodukt der neuraminidasen. Hoppe Seyler's Z Physiol<br />

Chem 1980; 361:697–702.<br />

86. Lentz MR, Webster RG, Air GM. Site-directed mutation of the active site of influenza<br />

neuraminidase and implications for the catalytic mechanism. Biochemistry 1987; 26:5351–5358.<br />

87. Chong AKJ, Pegg MS, von Itzstein M. Characterization of an ionisable group involved in the<br />

binding and catalysis <strong>by</strong> sialidase from influenza virus. Biochem Int 1991; 24:165–171.<br />

88. Taylor NR, von Itzstein M. Molecular modeling studies on ligand binding to sialidase from influenza<br />

virus and the mechanism of catalysis. J Med Chem 1994; 37:616–624.<br />

89. Janakiraman MN, White CL, Laver WG, Air MA, Luo M. <strong>Structure</strong> of influenza virus<br />

neuraminidase B/Lee/40 complexed with sialic acid and a dehydro anolog at 1.8A-resolution:<br />

implications for the catalytic mechanism. Biochem 1994; 33:8172–8179.<br />

90. Goodford PJ. A computational procedure for determining energetically favorable binding sites on<br />

biologically important macromolecules. J Med Chem 1985; 28:849–857.<br />

91. Von Itzstein M, Wu W-Y, Kok GB, Pegg MS, Dyson JC, Jin B, VanPhan T, Symthe ML, White HF,<br />

Oliver SW, Colman PM, Varghese JN, Ryan DM, Woods JM, Bethell R C, Hotham VJ, Cameron JM,<br />

Penn CR. Rational design of potent sialidase-<strong>based</strong> inhibitors of influenza virus replication. Nature<br />

1993; 363:418–423.<br />

92. Pearlman DA, Case DA, Caldwell J, Seibel G, Singh UC, Weiner PK, Kollman PA. AMBER 4.0.<br />

San Francisco, California: University of California, 1990.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_485.html (1 of 2) [4/9/2004 12:29:59 AM]


Document<br />

93. Woods JM, Bethell RC, Coates JAV, Healy N, Hiscox SA, Pearson BA, Ryan DM, Ticehurst J,<br />

Tilling J, Walcott SM, Penn CR. 4-Guanidino-2-4-Dideoxy-2,3-Dehydro-N-Acetylneuraminic Acid is a<br />

highly effective inhibitor both of the sialidase (Neuraminidase) and growth of a wide range of influenza<br />

A and B viruses In Vitro. Antimicro Agents and Chemotherapy 1993; 37:1473–1479.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_485.html (2 of 2) [4/9/2004 12:29:59 AM]<br />

94. Luo M,<br />

Jedrzejas MJ,<br />

Singh S, White<br />

CL, Brouillette<br />

WJ, Air GM,<br />

Laver WG.<br />

Benzoic acid<br />

inhibitors of<br />

influenza virus<br />

neuraminidase.<br />

Acta Cryst<br />

1995;<br />

D51:504–510.


Document<br />

Page 486<br />

95. Ryan DM, Ticehurst J, Dempsey MH, Penn CR. Inhibition of influenza virus replication in mice <strong>by</strong><br />

GG167 (40guanidino-2,4-dideoxy-2,3-dehydro-N-acetylneuraminic acid) is consistent with extracellular<br />

activity of viral neuraminidase. Antimicrob Chem Chemother 1994; 38(10),2270–2275.<br />

96. Thomas GP, Forsyth M, Penn CR, McCauley JW. Inhibition of growth of influenza virus in vitro <strong>by</strong><br />

4-guanidino-2,4-dideoxy-2,3-dehydro-N-acetylneuraminic acid. Antiviral Res 1994; 24:351–356.<br />

97. Hayden FG, Rollins BS, Madren LK. Anti-influenza activity of the neuraminidase inhibitor 4guanidino-Neu5Ac2en<br />

in cell culture and in human respiratory epithelium. Antiviral Res 1994;<br />

25:123–131.<br />

98. Hayden FG, Treanor JJ, Betts RF, Lobo M, Esinhart JD, Hussey EK. Safety and efficacy of the<br />

neuraminidase inhibitor GG167 in experimental human influenza, JAMA 1996; 275:295–299.<br />

99. Nohle U, Beau J-M, Schauer R. Eur J Biochem 1982; 126:543–548.<br />

100. Kimberlin DW, Crumpacker CS, Straus SE, Biron KK, Drew WL, Hayden FG, McKinlay M,<br />

Richman DD, Whitley RJ. Antiviral resistance in clinical practice. Antiviral Res 1995; 26:423–438.<br />

101. Madren LK, Shipman C, Hayden FG. In vitro inhibitory effects of combinations of anti-influenza<br />

agents. Antiviral Chem Chemotherapy 1995; 6(2):109–113.<br />

102. McKimm-Breschkin JL, Marshall D, Penn CR. Phenotypic changes observed in influenza viruses<br />

passaged in 4-amino or 4-guanidino-Neu5Ac2en in vitro. Abstracts, 9th Internat. Conf. on Negative<br />

Strand Viruses. Estoril, Portugal, 1994:260.<br />

103. Gubareva LV, Bethell R, Hart GJ, Penn CR, Webster RG. Characterization of mutants of influenza<br />

virus selected with 4-guanidino-Neu5Ac2en. Abstracts, 14th Annual Meeting, Amer. Soc. for Virol.<br />

Austin, Texas, 1995:W44-1.<br />

104. McKimm-Breschkin JL, Blick TJ, Sarasrabudhe AV, Tiong T, Marshall D, Hart GJ, Bethell RC,<br />

Penn CR. Generation and characterisation of variants of the NWS/G70C influenza virus after in vitro<br />

passage in 4-amino-Neu5Ac2en and 4-guanidino-Neu5Ac2en. Antimicrob Agents Chemotherapy 1995;<br />

in press.<br />

105. Blick TJ, Tiong T, Sahasrabudhe A, Varghese JN, Colman PM, Hart GJ, Bethell RC, McKimm-<br />

Breschkin JL. Generation and characterization of an influenza virus variant with decreased sensitivity to<br />

the neuraminidase specific inhibitor 4-guanidino-Neu5Ac2en. Virology 1995; 214:475–484.<br />

106. von Itzstein M, Wu W-Y, Jin B. The synthesis 2,3-didehydro-2,4-dideoxy-4-guanidiny-Nacetylneuraminic<br />

acid: a potent influenza virus inhibitor. Carbohydr Res 1994; 259:301–305.<br />

107. Kraulis PJ. MOLSCRIPT: a program to produce both detailed and schematic plots of protein<br />

structures. J Appl Cryst 1991, D50:869–873.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_486.html (1 of 2) [4/9/2004 12:30:20 AM]


Document<br />

Page 487<br />

19<br />

Rhinoviral Capsid-Binding Inhibitors: Structural Basis for Understanding<br />

Rhinoviral Biology and for <strong>Drug</strong> <strong>Design</strong><br />

Vincent L. Giranda<br />

Abbott Laboratories, Abbott Park, Illinois<br />

Guy D. Diana<br />

ViroPharma, Inc., Malvern, Pennsylvania<br />

I. Introduction<br />

A cure for the common cold has been sought after so long that it has become a clicheé: “We can<br />

_____(do something difficult) but we can't cure the common cold”. There are, unfortunately, a number<br />

of factors that conspire to make the cure for the common cold ephemeral. In spite of these difficulties,<br />

dramatic progress has been made in producing chemotherapies for human rhinoviruses (HRVs), which<br />

are the major cause of the common cold in humans [1]. Although most colds are generally both mild and<br />

self-limiting, they are responsible for both a large proportion of visits to physicians and lost work time<br />

[2]. Billions of dollars are spent in the United States alone on symptomatic relief from this disease. The<br />

ubiquitousness of this disease has led many people to seek a cure for many years (the discovery of the<br />

class of HRV inhibitors described here is over 20 years old). This chapter will describe the influence of<br />

structure-<strong>based</strong> approaches in the design of a class of antipicornaviral agents called capsid-binding<br />

inhibitors.<br />

Any effort to inhibit HRV replication, and thus cure many common colds, is made more difficult <strong>by</strong><br />

three factors. First, there are at least 102 described serotypes of HRVs, and it seems likely that there are<br />

many more serotypes not yet described [3]. Rhinoviruses are responsible for about 40–60% of the colds<br />

in humans [1,4]. Therefore, a chemotherapeutic agent would need to be effective<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_487.html [4/9/2004 12:30:25 AM]


Document<br />

Page 488<br />

against most of the HRV serotypes to be useful in approximately 50% of common colds. This<br />

percentage may be somewhat higher because some of the other causes of cold-like illnesses, particularly<br />

the enteroviruses (e.g., Coxsackie and echoviruses), are also inhibited <strong>by</strong> capsid-binding compounds<br />

[5–7]. In order to be efficacious, drugs must necessarily have a broad spectrum of activity.<br />

The second difficulty in producing HRV chemotherapy stems from the relatively innocuous nature of<br />

HRV infection. Compounds must be able to be very safely administered, with a minimum of drug-drug<br />

interactions, if therapy is to be acceptable. An analogy may be drawn between common headache<br />

remedies and common cold chemotherapy. Common headache cures such as nonsteroidal<br />

antiinflammatory agents and acetaminophen clearly can cause serious side effects (gastrointestinal<br />

bleeding or catastrophic liver failure) particularly if misused [8]. In spite of this possibility, serious<br />

complications from these agents are quite rare. One would suspect that an antirhinoviral agent would<br />

need to be at least as safe as headache remedies.<br />

The third major difficulty in developing cold cures arises from the fact that the HRVs are RNA viruses.<br />

When presented with any selective pressure, including chemotherapeutic or antibody challenge, RNA<br />

viruses mutate rapidly [9]. This ability to mutate is most clearly illustrated in influenza viruses (RNA<br />

viruses), where new strains continuously arise to circumvent immunity in a population. Influenza A<br />

viruses have been shown to mutate around the anti-influenza drug Amantadine, after a single passage<br />

through a susceptible human host. The mutated viruses shed from a host treated with Amantadine are<br />

now resistant to Amantadine. These mutated viruses appear to be as virulent as the parent strain of virus<br />

[10].<br />

Any effort at antirhinoviral therapy must attend to these three issues: (1) serotypic diversity; (2)<br />

exceptional safety; and (3) viral resistance. Therefore, any structure-<strong>based</strong> approach cannot concentrate<br />

on potency alone, but must also attend to these three issues as well. The requirement for inhibiting<br />

multiple targets has been addressed in tangible ways using structure-<strong>based</strong> design and will be discussed<br />

here. Safety issues have been addressed in limited published data from clinical and preclinical studies,<br />

but structure-<strong>based</strong> design has not played any significant role in addressing these problems. One might<br />

argue that structure-<strong>based</strong> approaches have aided in the design of clinical backups. These backups are<br />

then brought forward after an initial drug fails for safety reasons. <strong>Drug</strong>-resistant mutations created in the<br />

laboratory have also been examined structurally and will be discussed. The importance of resistance<br />

developing in the clinical setting has not yet been answered.<br />

This chapter will first introduce the target, the HRVs, and describe their anatomy and life cycle.<br />

Emphasis will be placed on the viral capsid and its disassembly or uncoating. How structural<br />

information has helped in our under-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_488.html [4/9/2004 12:30:28 AM]


Document<br />

Page 489<br />

standing viral physiology will be highlighted. This will be followed <strong>by</strong> a description of capsid-binding<br />

antirhinoviral compounds. We will leave discussions of other HRV targets, e.g., proteases, to other<br />

authors more knowledgeable in these subjects. <strong>Drug</strong> structure—activity relationships will be discussed,<br />

followed <strong>by</strong> a discussion of drug resistance. Finally clinical trials and future prospects for capsidbinding<br />

inhibitors in any antirhinoviral armamentarium will be discussed.<br />

II. The Human Rhinovirus Anatomy<br />

The human rhinoviruses are picornaviruses (pico = small; rna = RNA), a family of small (300 Å in<br />

diameter), positive sense, single-stranded RNA viruses. Other generas in this family include the<br />

enteroviruses (e.g., polioviruses); the aphthoviruses (e.g., foot-and-mouth disease virus); the<br />

cardioviruses (e.g., mengovirus, EMC virus); and the heparnaviruses (e.g., hepatitis A virus). With the<br />

exception of the heparnaviruses, a crystallographic structure is known for at least one of the viruses in<br />

each genera [5,11–17]. These structures show remarkable similarities, which will be described. In spite<br />

of these similarities, caution should be exercised when trying to generalize data gathered in one genus to<br />

other members of the picornavirus family.<br />

The picornaviruses share an icosahedral structure (Figure 1). The icosahedral protein coat that<br />

encapsidates the viral RNA is made up of 60 symmetrically arranged protomers. Each of these<br />

protomers is comprised of four viral polypeptides, termed VP1 through VP4. This was the extent of our<br />

structural knowledge of the picornaviruses until the 1985 structure of HRV14 was published <strong>by</strong> Michael<br />

Rossmann and coworkers [16]. This was followed rapidly <strong>by</strong> structural determination of a variety of<br />

other picornaviruses. These structures provided the framework on which to base current paradigms for<br />

the picornaviral life cycles, particularly with regard to their assembly, attachment, and uncoating.<br />

The VP1, VP2, and VP3 all contain a core eight-stranded antiparallel β barrel (Figure 2). These<br />

polypeptides have surfaces on both the exterior and interior of the virion particle, facing both solvent<br />

and viral RNA. The fourth polypeptide, VP4, is considerably smaller than the others and does not<br />

contain the eight-stranded barrel motif. The VP4 resides entirely on the interior surface of the virion, in<br />

close association with the viral RNA. The N-terminus of VP4 is known to be myristoylated in both rhino-<br />

and polioviruses [18,19].<br />

Three regions of the picornavirus structure deserve special attention because they appear to play crucial<br />

roles in the viral life cycle as well as bear on the function of the capsid-binding compounds. These<br />

regions are the canyon, the VP1 hydrophobic pocket, and the β cylinder.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_489.html [4/9/2004 12:32:00 AM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_490.html (1 of 2) [4/9/2004 12:32:47 AM]<br />

Page 490


Document<br />

Figure 1<br />

Schematic illustration of the icosahedral rhinovirus 14.<br />

(a) Shown is the icosahedron comprised of 60 copies<br />

each of VP1 (light gray), VP2 (black), and VP3 (gray).<br />

The shaded circles around each five-fold axis indicate<br />

the canyon positions. Also indicated is the approximate<br />

position of the VP1 hydrophobic pocket that lies<br />

underneath the surface of the virion. (b) An<br />

icosahedral pentamer is expanded with one viral protomer<br />

shown as a protein ribbon diagram. (c) This pentamer is<br />

seen in a cutaway view. Here VP1 is white, VP2 and VP4<br />

black, and VP3 gray. A capsid-binding compound is<br />

depicted as black spheres inside the VP1 ribbon diagram.<br />

The cross hatched regions on the (c) schematic (right)<br />

indicate areas that disorder when HRV14 crystals are<br />

exposed to acid.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_490.html (2 of 2) [4/9/2004 12:32:48 AM]


Document<br />

A. The Canyon<br />

Page 491<br />

The HRVs have been divided into the major and minor receptor groups <strong>based</strong> on two identified cellular<br />

receptors [3]. The major group, which is comprised of approximately 90 serotypes, binds to the<br />

intercellular adhesion molecule 1 (ICAM-1) [20]. The minor group, about 10 serotypes, binds to the low<br />

density lipoprotein receptor family [21].<br />

The canyons are depressions approximately 15 to 20 Å deep that encircle each icosahedral five-fold axis<br />

(Figure 1). When first seen in HRV 14, these canyons were postulated to be the site at which a cellular<br />

receptor would bind. Subsequent electron-microscopic data revealed that ICAM-1 does indeed bind in<br />

the canyon as predicted, although in a somewhat different orientation than early models [22,23]. These<br />

canyons allow the receptor binding sites to escape immunological surveillance because the canyons are<br />

too narrow to allow an immunoglobulin to contact the canyon floor. Directly underneath the floor of the<br />

canyon lies a second important structure, the VP1 hydrophobic pocket.<br />

B. The VP1 Hydrophobic Pocket<br />

The VP1 hydrophobic pocket is the site where the capsid-binding compounds reside. This hydrophobic<br />

pocket in VP1 was not initially apparent in the HRV14 structure because it exists in a closed<br />

conformation in the native HRV14. The addition of a capsid-binding antiviral agent induces the pocket<br />

to open. This was first seen <strong>by</strong> Smith and coworkers when the first crystal structure of a capsid-binding<br />

drug, WIN 51711, was solved bound in HRV14 [24]. (WIN is the designation for a Sterling Winthrop<br />

compound.) This was a seminal event in the structure-<strong>based</strong> design of these capsid-binding compounds;<br />

before this structure was solved the exact site at which the compounds bound was unknown.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_491.html [4/9/2004 12:33:24 AM]


Document<br />

Figure 2<br />

Panels (a), (b), and (c) depict VP1, VP2, and VP3 β barrels, respectively. In all<br />

cases the view is such that the virion exterior would be on the right side of<br />

the page. A capsid-binding compound is depicted as gray spheres as it is<br />

found inside VP1.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_492.html (1 of 2) [4/9/2004 12:34:08 AM]<br />

Page 492


Document<br />

Subsequent studies have shown this pocket is present in every known HRV and enterovirus structure.<br />

Page 493<br />

This pocket is inside the β barrel of VP1, directly underneath the canyon floor where ICAM-1 binds<br />

(Figure 1c). The proximity of the hydrophobic drug-binding site to the receptor-binding site explains the<br />

effects these compounds have on viral attachment in specific HRV serotypes (see below).<br />

C. The β Cylinder<br />

The β cylinder is a complicated structure that lies on the inside of the viral coat (Figure 1c). There is one<br />

β cylinder at each icosahedral five-fold axis. The β cylinder is formed <strong>by</strong> winding the five-fold related<br />

VP3 N-termini around the symmetry axis. In close association with this cylinder is the N-terminal<br />

regions of VP1 and VP4.<br />

The myristoyl moiety attached to VP4 is observed in the poliovirus structures. This moiety is in close<br />

association with the cylinder formed <strong>by</strong> VP3 [19]. In HRVs, density consistent with the myristoyl<br />

moiety is also seen near the β cylinder [25]. In the HRVs, structural disorder of the first 25–28 Nterminal<br />

residues of VP4 obscures the connection between VP4 and the myristoyl group. The addition of<br />

myristic acid to protein is seen in many viral and cellular<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_493.html [4/9/2004 12:34:17 AM]


Document<br />

proteins. It is typically a signal moiety that directs the protein to which it is attached towards cellular<br />

membranes [26].<br />

Page 494<br />

Above the β cylinder is an ion, thought to be Ca 2+, which lies right on the five-fold axis and coordinates<br />

to the five adjacent VP1 polypeptides [25,27]. The proximity of the β cylinder to structures shown later<br />

to become external during uncoating (portions of VP1, VP4) suggests that it may be important in<br />

uncoating.<br />

III. The Human Rhinovirus Life Cycle<br />

The HRVs must undergo a number of transitions to replicate (Figure 3). First, they must attach to the<br />

cell surface at a cellular receptor. They are then internalized into the cell via the endosomal<br />

compartment. Following internalization, they must eject their RNA out of the viral capsid, through a<br />

lipid bilayer, and into the cytosol in a manner that preserves the integrity of the RNA. Replication<br />

Figure 3<br />

A schematic diagram depicting some of the<br />

required steps in viral replication.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_494.html (1 of 2) [4/9/2004 12:34:39 AM]


Document<br />

Page 495<br />

of the RNA and production of a polyprotein ensues. The polyprotein is then processed <strong>by</strong> viral proteases<br />

to form the viral polypeptides. The coat must then assemble, package RNA, and leave the cell. The<br />

capsid-binding compounds' effect on propagation have been shown to occur at the attachment,<br />

uncoating, and assembly steps (25,27–30). The relative effect on each of these steps appears to be<br />

variable and has not yet been extensively characterized. The most universal effect appears to be on the<br />

uncoating step.<br />

During uncoating several different rhinovirus subparticles are observed. These are thought to be<br />

intermediates in the uncoating process [32–33]. The fully infectious 149S particle becomes an 125S Aparticle,<br />

which has lost VP4 but still maintains viral RNA. This A-particle is no longer infectious. The Aparticle<br />

then releases RNA to become an 80S empty shell.<br />

The formation of these types of particles can be induced <strong>by</strong> association of the virion with the cell<br />

receptor or acidification [32,34]. In poliovirus, attachment to cells has been shown to lead to a particle<br />

that has lost VP4 and has externalized the N-terminal region of VP1, which normally resides on the<br />

virion interior.<br />

There has been considerable debate about how the HRVs accomplish the transfer of RNA from inside<br />

the virion into the cell cytosol. This step is crucial for productive uncoating. An important question<br />

concerns the requirement for acidification of the endosome for HRVs to release their RNA. Evidence<br />

that appeared to conflict was found in a number of studies using either entero- or rhinoviruses (35–39).<br />

This question was later addressed <strong>by</strong> experiments that specifically separated entero- and rhinovirus<br />

behavior [40]. These experiments showed that HRVs, unlike poliovirus, require a pH-lowering step for<br />

productive infection. This pH lowering is likely to occur in the endosomal compartment. It should be<br />

noted that HRV and enteroviruses have been classified historically <strong>based</strong> on their resistance to acid:<br />

HRVs are acid-labile, while enteroviruses are stable in acid. Consequently, differences in behavior<br />

between the rhino- and enteroviruses in an acidic environment within the cell are not surprising.<br />

To replicate the changes in HRV that might be induced <strong>by</strong> acidification in the endosome, HRV14 was<br />

acidified in the crystalline state and examined via x-ray diffraction. When compared to the native<br />

HRV14, the acidified HRV14 capsid becomes disordered in three regions: the Ca 2+ ion on the five-fold<br />

axis; a region of the β cylinder and the adjacent portion of VP4; and the GH loop [41]. The GH loop is<br />

the region of structure that connects β-strands G and H and lies directly between the hydrophobic pocket<br />

and the receptor binding site (Figure 4). It forms the roof of the VP1 hydrophobic pocket and the floor of<br />

the canyon at the receptor binding site.<br />

Mutants that are resistant to acid in vitro were isolated [41,42]. These mutants cluster about the GH loop<br />

(Table 1, Figure 4), and typically would be thought of as mutants that stabilize protein structure (e.g.,<br />

larger or more<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_495.html [4/9/2004 12:34:53 AM]


Document<br />

Table 1 VP1 Acid- and <strong>Drug</strong>-Resistant Mutations<br />

Phenotype Mutations in VP1<br />

Acid resistant H1078L, H1078Y, N1100K, N1100T, D1101E, N1145S,<br />

W1163R, V1188A, V1191A, T1216I, M1221L, M1224L,<br />

A1225V<br />

<strong>Drug</strong> resistant (compensation) N1100S, N1105S, V11531, N1219S, S1223G<br />

<strong>Drug</strong> Resistant (exclusion) V1188L, V1188M, C1199F, C1199Y, C1199W, C1199R<br />

branched amino acids) [43–46]. The proximity of these mutations to the GH loop coupled with the<br />

loop's movement on acidification suggest that the GH-loop movement plays an important role in<br />

uncoating.<br />

It has also been suggested that binding of ICAM-1 to the virions may also play a role in uncoating. In<br />

support of this hypothesis, it has also been observed<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_496.html (1 of 2) [4/9/2004 12:35:37 AM]<br />

Page 496


Document<br />

Figure 4<br />

A ribbon diagram of HRV14 VP1 bound to WIN 61605 (small gray spheres<br />

with black bonds). The VP1 region that disorders under acid conditions is<br />

depicted <strong>by</strong> black color on the ribbon diagram. Residues that can be<br />

mutated to be acid stable, drug resistant (compensation type) or to either<br />

phenotype are shown as black, white, and gray ball-and-stick models,<br />

respectively. The majority of mutations are near the site of drug binding<br />

as well as the site of acid-induced disorder in VP1.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_496.html (2 of 2) [4/9/2004 12:35:37 AM]


Document<br />

Page 497<br />

that in some HRVs (particularly HRV14 and to a lesser extent in HRV3, but not HRV16) the addition of<br />

ICAM-1 itself, in the absence of acid, can induce structural changes that mimic uncoating in the capsid<br />

[34]. However, these changes are unlikely to lead to productive uncoating because they occur in the<br />

extracellular space. Therefore, the binding of ICAM alone, at the cell surface, is unlikely to be sufficient<br />

to cause productive uncoating of HRVs.<br />

It seems that the hydrophobic pocket in HRVs is maintained to allow an induced transition around the<br />

GH loop that is required for productive uncoating. This transition could be induced <strong>by</strong> acidification in<br />

the endosome, receptor binding, or a combination of the two. Filling this pocket in VP1 with a drug or<br />

naturally occurring factor would inhibit this transition, thus inhibit uncoating. As expected, binding of<br />

compounds in the VP1 pocket has also been shown to inhibit intracellular uncoating as well as either<br />

acid-or heat-induced uncoating of the virus [28,29,47].<br />

An attractive hypothesis suggests that the GH loop transition precedes or allows the externalization of<br />

VP4, which would be required for the formation of the uncoating intermediate particles. Remember,<br />

VP4 contains the myristoyl moiety, which can signal VP4 to associate with a membrane. The VP4<br />

would drag the N-terminal region of VP1 (to which it is closely associated) to the exterior of the virion.<br />

This would result in a particle with the N-terminus of VP1 exposed, as has been observed in poliovirus<br />

[48–50]. The sequence of this exposed region of VP1 suggests that it can form an amphipathic helix.<br />

The VP1 helices could then insert into the membrane and form a pore, which could allow the passage of<br />

RNA through the lipid bilayer into the cytosol. This is reminiscent of the pore formation <strong>by</strong> colicin [51].<br />

The observation that both the Ca 2+ ion plus the β cylinder and VP4 become disordered under acidic<br />

conditions are consistent with this hypothesis. These are regions that would need to disorder to allow the<br />

externalization of VP4 and the N-terminus of VP1.<br />

IV. Capsid-Binding Compounds<br />

Capsid-binding compounds were discovered long before the emergence of the HRV crystal structure<br />

[52]. They were initially discovered on screening of compounds that had been produced <strong>by</strong> an insect<br />

pheromone project at Sterling Winthrop. Examples of compounds known or presumed to bind in this<br />

pocket are shown in Figure 5 [52–69]. The prototypical WIN drug contains an oxazoline ring attached to<br />

a phenoxy group, which is in turn linked <strong>by</strong> an aliphatic chain to an isoxazole ring. The three rings will<br />

be termed A, B, and C here (Figure 6). Compounds of this type were the first to be shown to inhibit the<br />

viral uncoating and also to stabilize the virion to heat-induced denaturation [29].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_497.html [4/9/2004 12:35:57 AM]


Document<br />

Figure 5<br />

Some compounds which have been shown to inhibit picornavirus<br />

replication. These are thought to bind in the VP1 hydrophobic<br />

pocket. References are indicated on the figure.<br />

Page 498<br />

The structure of the first compound to be solved in complex with HRV 14 was found to be bound in an<br />

extended conformation within the VP1 hydrophobic pocket [24]. The compound is almost entirely<br />

buried within the capsid of the virion. Since in the native HRV14 the pocket exists in a closed<br />

configuration, binding required large motions in VP1 to accommodate the drug. These<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_498.html (1 of 2) [4/9/2004 12:36:34 AM]


Document<br />

Figure 6<br />

Some WIN compounds depicting different rings<br />

and linkers. Note WIN 52452 has no C ring.<br />

motions, of up to approximately 4.5 Å, are most pronounced in the region of the GH loop.<br />

In contrast, HRV1A and HRV16, a minor and major receptor group virus respectively, have their<br />

pockets in open conformations even in the absence of drug [13,15]. For these serotypes, drug binding<br />

typically shifts positions of capsid atoms a maximum of 1 to 2 Å, smaller than those shifts seen in<br />

HRV14. <strong>Drug</strong>s bind in these pockets in a fashion similar to that seen in HRV14, extended and almost<br />

entirely buried <strong>by</strong> VP1.<br />

Page 499<br />

The structures of five HRVs have been solved to date: HRVs 1A, 3, 14, 16, and 50. In all of these<br />

HRVs, as well as the polio- and coxsackie viruses, VP1 hydrophobic pockets have been observed<br />

[5,12,13,15,17,24,70,71] (HRV3, Zhao, R. et al., personal correspondence; HRV50, Giranda V. L. et al.,<br />

unpub-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_499.html (1 of 2) [4/9/2004 12:36:54 AM]


Document<br />

Figure 7<br />

HRV14 VP1 hydrophobic pocket schematic with WIN 61605<br />

illustrating some of the terminology commonly used to describe<br />

this pocket. Notice the GH loop separates the pocket from the<br />

canyon floor.<br />

Page 500<br />

lished data). These pockets all share similar features and have been described as foot shaped, with a<br />

hydrophobic toe region, a heel region capable of hydrogen bonding, and a pore region near the ankle of<br />

the foot (Figure 7).<br />

Many of the picornavirus structures have been shown to have electron density in their VP1 pockets even<br />

in the absence of any added drug. These densities have been modeled as fatty acids or similar<br />

compounds [12,15,56,70–72]. The occurrence of these pocket factors have led some to hypothesize that<br />

these factors perform a similar function as do capsid-binding inhibitors, that is, to stabilize the virions<br />

[15,24,41,73,74].<br />

Teleologically one could argue that the virion would pick up a fatty acid in its VP1 pocket before its<br />

egress from the cell, which would then stabilize the virion in transit to new hosts. The HRVs are known<br />

to be stable for long periods of time on surfaces and the dominant mode of transmission is thought to be<br />

hand-to-hand contact [75,76]. When a new host cell is reached, the stabilization factor might exit the<br />

pocket. This would allow the necessary conformational transition (probably at the GH loop) to occur and<br />

allow productive uncoating.<br />

A. Multiple Targets<br />

The use of these structures in a traditional structure-<strong>based</strong> drug design approach has been limited <strong>by</strong> the<br />

large number of unknown target serotype structures. There have been useful studies that have grouped<br />

picornaviruses <strong>based</strong> on their susceptibility to various antiviral agents (see below) [7,77]. However, in<br />

order to design a truly broad-spectrum single drug, the structural elements common among many<br />

serotypes must be considered.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_500.html (1 of 2) [4/9/2004 12:37:20 AM]


Document<br />

Page 501<br />

Because of the large number of serotypes, a successful inhibition strategy needs to consider whether it is<br />

better to have a drug that is exceedingly potent against a small subset of HRV serotypes (e.g. 30%) or a<br />

drug that is somewhat less potent but effective against most serotypes (e.g. 90%). Because of these<br />

considerations, different parameters are required to describe viral inhibition.<br />

Two important values that have been used extensively to describe potency for antipicornaviral<br />

compounds include the mean inhibitory concentration (MIC) and the MIC 80. The MIC is the<br />

concentration that inhibits the viral progeny production <strong>by</strong> 50% in a cell-<strong>based</strong> plaque assay. The mean<br />

MIC is the average MIC over the number of serotypes against which the compounds have been tested.<br />

The MIC 80 is the MIC at which at least 80% of the serotypes tested will be inhibited <strong>by</strong> at least 50%.<br />

Another way to think about the MIC 80 is that it is the MIC value for the serotype that is at the 80th<br />

percentile rank for viruses inhibited. For example, if ten viruses were tested, the MIC 80 would be the<br />

MIC concentration of the drug that inhibits the 8th most sensitive virus [15].<br />

B. Potency and Binding Energetics<br />

It would be reasonable to assume that the activity of a drug (MIC) against a specific serotype would be<br />

related to its binding energy. This is an important consideration because algorithms that are used to<br />

predict potency rely on estimations of binding energy. The only experiment completed to directly study<br />

this correlation suggests a rough correlation in a small number of samples. This study however was<br />

limited to a small number of compounds in a single chemically similar series [78]. It is unclear whether<br />

this relationship will hold over diverse chemical entities.<br />

One could imagine a series of compounds that binds, but is ineffective at stabilizing the virion to any<br />

extent, thus ineffective in inhibiting viral replication. This appears to be what was observed when<br />

fragments of WIN compounds that only contained the A and B rings were examined. These fragments<br />

were less able to stabilize the virus to heat-induced denaturation than intact compounds. Although the<br />

binding constant for these compounds has not been determined, the diminished thermostabilization<br />

occurred at concentrations of compound sufficient to allow the compounds to be seen bound in the VP1<br />

pocket via x-ray crystallography [79]. This observation suggests that the drug binding and inhibition<br />

have been decoupled to some degree.<br />

Another question concerning correlating binding affinity to viral inhibition is the number of sites per<br />

virion required to be occupied before the virion can no longer uncoat: all 60, or a few? Further, is there<br />

any positive or negative cooperativity in the interaction? These questions have not yet been answered.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_501.html [4/9/2004 12:37:57 AM]


Document<br />

C. <strong>Structure</strong>-Activity Relationships<br />

Page 502<br />

Structural features, as stated above, have been found that are common to the known HRV structures.<br />

These features place constraints on compounds that have been demonstrated in a large number of<br />

structure—activity relationships. Shape considerations, hydrophobicity requirements, hydrogen-bonding<br />

requirements and compound flexibility will be discussed.<br />

Pocket-Shape Considerations<br />

Compound Length. The results of x-ray studies on several HRV serotypes have shown that the<br />

hydrophobic pockets vary in size, but all pockets impose structural constraints on compounds. The<br />

pockets, which almost completely enclose the compounds, require that both drug length and width must<br />

be limited. The results of structure—activity relationships for one series of WIN compounds have shown<br />

that the optimum length of the aliphatic linking region between the phenoxy (B ring) and isoxazole (C<br />

ring) ring is between 3 and 6 carbons (Table 2). This optimum length will vary in other series <strong>based</strong> on<br />

the extent of substitution on the A and C rings (see below) [54,80–82]. This effect on linker length is<br />

obvious both for individual serotypes like HRV14 and for the MIC 80, which measures many serotypes.<br />

This effect of pocket length can also be seen when substituents are added to either the A or C ring which<br />

would lengthen the compounds. Additions of alkyl chains of four or more carbons tend to decrease<br />

potency, whereas shorter chains may increase potency (Table 3) [54,83].<br />

In studies on compound length, HRV14 appears to be more sensitive to longer compounds than are<br />

many other rhinoviruses. This is consistent with the observation that the hydrophobic pocket of HRV14<br />

is longer than that of HRVs<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_502.html (1 of 2) [4/9/2004 12:38:14 AM]


Document<br />

Page 503<br />

1A, 16, and 50. Cluster analysis examining the sensitivity of 100 different picornaviruses against 15<br />

different inhibitors showed that the sensitivities of different viruses depends on the shape of the pocket,<br />

with shorter pockets preferring shorter compounds [7]. This analysis is consistent with the structural<br />

observations, i.e., the viruses with long pockets (e.g., HRV14 and polioviruses) cluster together. A<br />

cluster distinct from that which contains HRV14, and prefers shorter compounds, is the group that<br />

contains the shorter pocket viruses: HRVs 1A, 16, and 50 (Figure 8).<br />

It is interesting that these drug sensitivity groups do not correspond to the receptor binding groups of the<br />

viruses. For example, both HRV14 and 50 are major receptor group viruses, but the pocket of HRV50 is<br />

clearly shaped more like the pocket of the minor receptor group virus 1A.<br />

The argument that the two different drug sensitivity groups may constitute distinct classes of HRVs has<br />

been bolstered <strong>by</strong> genetic examination, which has shown that clusters with similar amino acid sequences<br />

correlate well with the drug sensitivity [84]. As with the pocket shapes, these genetic clusters do not<br />

correlate with the two receptor groups into which the HRVs are placed.<br />

Attempts have been made to break through the length barrier imposed <strong>by</strong> the pocket structure. One<br />

could conceive of a drug that could pass through the pore at the heel of the pocket and connect the VP1<br />

hydrophobic pocket to the canyon floor. Such a drug may have activity that could inhibit uncoating via<br />

stabilization of VP1 as well as inhibit attachment <strong>by</strong> directly blocking the receptor site. Attempts have<br />

been made to synthesize just such compounds, with extended “tails” attached to the C-ring. While these<br />

compounds bind, their tail does not extend through the pocket pore, but rather coils up within the pocket<br />

[85].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_503.html [4/9/2004 12:39:29 AM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_504.html (1 of 2) [4/9/2004 12:40:23 AM]<br />

Page 504


Document<br />

Figure 8<br />

Solvent accessible surfaces (dot surfaces) of (a) HRV1A, (b) 14,<br />

(c) 16, (d) 50 filled with WINs 56291, 61605, 56291, 61209<br />

respectively (ball and stick). In all cases the cutaway view of the<br />

pocket has the viewer looking from inside the virion. The pocket<br />

of HRV14 is narrower and longer than the others and this<br />

is reflected in the drug structure—activity relationships (Tables 1–3).<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_504.html (2 of 2) [4/9/2004 12:40:23 AM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_505.html (1 of 2) [4/9/2004 12:41:31 AM]<br />

Page 505


Document<br />

Page 506<br />

Compound Width. Bulk considerations have also been examined for the phenoxy or B ring [80]. In<br />

HRV14, which has a long narrow pocket, compounds with no substituents on the phenoxy ring were the<br />

most potent. However most other serotypes were more sensitive to compounds with substituents on the<br />

phenoxy ring, either a dimethyl or a dichloro. The serotypes that preferred disubstituted compounds<br />

included HRV1A and 50, both known to have wider, shorter pockets (Table 4). This is also in agreement<br />

with the drug-clustering analysis discussed above [7].<br />

Induced VP1 Pocket Changes. The shape constrains discussed above are <strong>by</strong> nature somewhat<br />

qualitative. The precision with which one could predict the binding energy of a compound <strong>based</strong> on its<br />

shape is limited <strong>by</strong> the ability of the pocket to conform to its occupant. Conformational changes within<br />

the pocket are most pronounced between native and drug-bound HRV14, but present to a lesser extent in<br />

both HRV1A and 16. The extent of conformational changes produced <strong>by</strong> different drugs on the same<br />

virus may differ. This observation is illustrated in a study of a compound SCH 38057 in HRV14. This<br />

compound has a substantially different structure than the WIN compounds, and when bound to HRV14,<br />

induces changes in HRV14 that are also quite different from changes resulting from the binding of a<br />

WIN compound [62].<br />

The ability of compounds to effect conformational changes in the capsid when bound and the variability<br />

of these changes both between drug classes and viral serotypes have made a priori predictions of<br />

potency <strong>based</strong> on shape considerations very difficult. Comparative Molecular Field Analysis (CoMFA)<br />

studies has been useful, when examining similar compounds in a single serotype (HRV14), in<br />

demonstrating spatial constraints [85], but generalizing these constraints over many serotypes and drug<br />

classes is much more difficult.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_506.html [4/9/2004 12:41:41 AM]


Document<br />

Hydrophobicity Requirements<br />

<strong>Drug</strong> binding is enhanced <strong>by</strong> hydrophobicity in that portion of the drug that binds to the pocket toe.<br />

Quantitative structure-activity relationship (QSAR) analysis of these compounds have consistently<br />

shown that the most predictive parameter of antiviral activity is a measure of hydrophobicity, the<br />

octanol:water partition coefficient (logP) [80,82,85]. These studies have also consistently shown that<br />

there is no apparent correlation between electrostatic potential or dipole moment and potency.<br />

Page 507<br />

This evidence suggests that the drug—pocket interactions in the toe of the pocket are low-intensity<br />

hydrophobic interactions. Supporting this hypothesis is the finding that many structurally diverse<br />

molecules can be accommodated in this region, and even structurally similar molecules can bind in<br />

distinctly different conformations. Closely related structures have been shown to shift along their long<br />

axis <strong>by</strong> as much as 1.6 Å with respect to their phenoxy substituents [55]. There is even more variability<br />

with respect to the isoxazole placement (Table 5, Figure 9). No single hydrogen-bonding or electrostatic<br />

interaction appears to predominate within the pocket toe.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_507.html [4/9/2004 12:42:05 AM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_508.html (1 of 2) [4/9/2004 12:42:46 AM]<br />

Page 508


Document<br />

Figure 9<br />

A solvent-accessible surface of HRV14 (from the 61605 bound<br />

structure, dot surface) overlaid with WIN 61605 (ball and stick),<br />

51711, 56826, and 56291 (a). The atoms closest to the toe pocket<br />

wall (top) for each compound are in approximately the same<br />

position, while the isoxazole ring (C ring) positions vary significantly.<br />

The atom closest to the toe pocket wall is quite different in (b),<br />

which compares WIN 61605 (ball and stick) to R61837 (tubes) and<br />

SCH 38057 (lines).<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_508.html (2 of 2) [4/9/2004 12:42:46 AM]


Document<br />

Page 509<br />

The correlation of potency with logP might suggest that A rings with the lowest hydrogen-bonding<br />

potential would be the most potent. This has not been shown to be true. In fact, rings with heterocyclic<br />

nitrogen atoms are preferred to furan and thiophene rings, which would be expected to have less<br />

hydrogen-bonding potential [86,87]. This observation even extends to tetrazole rings, which have often<br />

been used in pharmaceutical design to replace hydrophilic groups such as carboxylates or esters.<br />

Explanations for the variability in potency due to A-ring changes have not been satisfactory. These<br />

heterocycles lack any consistent pattern of hydrogen bonds with the pocket residues (Table 6) [88]. Like<br />

QSAR analysis, structural analysis has not been able to show any relationship between hydrogenbonding<br />

groups in the A ring, or dipole moment, and potency [80,85].<br />

Extensive structural characterization of many different A-ring heterocycles has not yet been done.<br />

Difficulties predicting relative potency of these compounds a priori stem from the lack of understanding<br />

of solvation/desolvation effects as well as difficulties in characterizing the low-intensity hydrophobic<br />

interactions. Consequently, it seems likely that new structure—activity relationships about the A-ring<br />

heterocycle will continue to be determined <strong>based</strong> on empirical findings.<br />

Hydrogen-Bonding Requirements<br />

Capsid-binding compounds with a hydrogen-bond accepting atom in the region of the drug that binds to<br />

the VP1 pocket heel appear to demonstrate greater potency than do their non-hydrogen-bonding<br />

counterparts (Table 7) [55,89]. Studies using HRV14 where Asn1219, a hydrogen-bond donor in the<br />

pore region, has been mutated to Ala have shown that the compounds bind as well in the mutated virus<br />

as in the native virus [27]. This might suggest that the hydrogen bonding to Asn1219 at the pocket pore<br />

is unimportant. Recent structures of compounds in HRV14 have shown that other hydrogen-bond donors<br />

can function in place of Asn1219, even if Asn1219 is still present. These groups in HRV14 are the<br />

hydroxyl of Ser1107 and the backbone nitrogen of Leu1106. Other hydrogen-bonding groups are present<br />

in other rhinoviruses that coordinate to tightly bound waters, which can also act as hydrogen-bond<br />

donors [55,56]. This provides a flexible hydrogen-bonding network that can then<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_509.html [4/9/2004 12:42:48 AM]


Document<br />

accommodate a variety of different hydrogen-bonding groups in the C ring of the drug (Figure 10).<br />

Page 510<br />

This hydrogen-bonding network has also been observed in all classes of capsid-binding compounds that<br />

have had their structures determined in HRVs [58,62]. Reports of compounds that bind to the capsid and<br />

inhibit picornavirus uncoating, but would seem to have little hydrogen-bonding potential (e.g.,<br />

dichloroflavan) have not been structurally examined [60]. The effect of adding a hydrogen-bonding<br />

group to these compounds near the pocket pore is unknown, but one might expect that the potency of the<br />

compound would improve. Remember that WIN compounds without a hydrogen-bonding group at the<br />

pore region still have antiviral activity, albeit this activity is much weaker than that of an analogous drug<br />

with the hydrogen-bonding group present.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_510.html (1 of 2) [4/9/2004 12:43:03 AM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_511.html (1 of 2) [4/9/2004 12:44:11 AM]<br />

Page 511


Document<br />

Figure 10<br />

Possible hydrogen bonds in the pore region of HRV14 bound to WIN<br />

61605 (ball and stick). The waters (spheres W1 and W2) could<br />

potentially form hydrogen bonds to WIN 61605 as well as the side<br />

chains of Asn1219 and Ser1107 as well as the backbone of Leu1106<br />

(residues highlighted as tubes). The viewer is looking<br />

from the virion exterior.<br />

Page 512<br />

The fluidity of the hydrogen-bond donation network at the pore could allow a large number of<br />

structurally distinct molecules to bind in this location. If the function of the pocket is to bind fatty acids<br />

or other structurally diverse pocket factors and thus stabilize the virion. The flexibility of the hydrogenbonding<br />

network at the pore seems ideally suited for such a purpose.<br />

<strong>Drug</strong> Flexibility<br />

The WIN compounds all contain a flexible linking region that allows them to conform to differently<br />

shaped interiors of VP1 hydrophobic pockets. This flexibility is seen in many compounds that share the<br />

WIN-drug mechanism of action. However, some compounds do not contain a region as obviously<br />

flexible as an aliphatic linker (e.g., those with unsaturated rings as linkers, Figure 5).<br />

Flexibility in the aliphatic linking region of the WIN compounds has been explicitly studied and the<br />

results suggest that such a property is important for<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_512.html (1 of 2) [4/9/2004 12:44:42 AM]


Document<br />

Page 513<br />

broad spectrum activity (Table 8) [56,90]. While certain less-flexible compounds may have increased<br />

potency for a particular serotype of HRV, these compounds have lower potency versus other serotypes.<br />

It is not known if the effect of flexibility is an equilibrium or kinetic effect. The flexibility might allow<br />

the compounds to expand or contract to fill available space in the VP1 hydrophobic pocket.<br />

Alternatively, the flexibility may allow the compounds to achieve a conformation required to enter or<br />

leave the pocket, but this conformation would not be seen in the crystallographic experiment. If this is<br />

true, modeling of the equilibrium structure of compounds in the pocket will not be accurate predictors of<br />

compound potency.<br />

<strong>Structure</strong>—Activity Relationship <strong>Summary</strong><br />

The combination of classical structure—activity relationships combine with the structures of several<br />

compounds in a number of HRV serotypes have led to a paradigm for designing potent compounds. The<br />

effects of pocket shape, requirements for hydrophobicity in the toe of the pocket, as well as the potential<br />

for hydrogen bonding in the heel region appear to be strong determinants of antiviral potency. The<br />

requirement for a broad spectrum has so far limited a more detailed paradigm, such as one that might<br />

exist for designing inhibitors to a specific serine protease. In spite of this vagueness of antirhinoviral<br />

structure—activity relationships, they have been useful in guiding new compound synthe-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_513.html [4/9/2004 12:44:57 AM]


Document<br />

Page 514<br />

sis. Large numbers of compounds, which might have been synthesized in the absence of any structural<br />

knowledge, have been eliminated from the list of inhibitors to be created, because these compounds<br />

would not have fit well into the VP1 pocket due to its finite size and hydrophobic nature.<br />

The goal of being able to predict with greater accuracy the potency and spectrum of compounds before<br />

they are synthesized awaits three developments: the structures of more HRVs and compounds, the<br />

ability to more accurately model hydrophobic interactions, and probably the most difficult, the ability to<br />

predict changes in the HRVs that occur due to drug binding.<br />

V. Viral Resistance<br />

As would be expected for any virus, particularly an RNA virus, resistance to capsid-binding compounds<br />

has been observed in the laboratory. The effect such mutations have on the life cycle of the virus is<br />

important. Mutations that occur in regions where changes are poorly tolerated, because that region<br />

serves a vital function in the viral life cycle, are likely to lead to less virulent viruses. The almost<br />

universal inhibition of rhino- and enteroviruses <strong>by</strong> capsid-binding compounds, coupled with the<br />

observation that all of the viruses with determined structures have a VP1 hydrophobic pocket, suggests<br />

that these pockets serve an important and similar function. It seems unlikely that so many viruses would<br />

maintain such a pocket if it were not a selective advantage. <strong>Drug</strong>-resistance mutants have been classified<br />

into two groups, exclusion mutants and compensation mutants.<br />

A. Exclusion Mutants<br />

The exclusion mutants' behavior has been readily and adequately determined <strong>by</strong> biochemical and<br />

crystallographic means [91,92]. The mutations occur within the hydrophobic pocket of VP1 and<br />

thermostabilization studies have shown that these mutations preclude the binding of drug in the VP1<br />

pocket. One mutation site in HRV14 has been located at position 1188, which is on the side of the<br />

pocket closest to the viral interior, away from the canyon. Mutations at this site that convey resistance<br />

are Val rarrow.gif Leu or Val rarrow.gif Met. In both cases the mutation is to a larger side chain,<br />

which would be expected to fill the pocket. The crystal structure of the Val rarrow.gif Leu mutation<br />

confirms this hypothesis and demonstrates that the Leu side chain occupies space that would normally<br />

be occupied <strong>by</strong> an antiviral drug.<br />

The second site found in HRV14 is at Cys1199. Mutations to Phe, Tyr, Trp, and Arg all confer<br />

resistance at this site. Again, all of these mutations are to larger side chains. The hypothesis that these<br />

mutations function <strong>by</strong> excluding<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_514.html [4/9/2004 12:45:35 AM]


Document<br />

drug from the pocket is confirmed <strong>by</strong> crystallographic analysis showing that the Phe side chain in the<br />

Cys rarrow.gif Phe mutation occupies the site which would be occupied <strong>by</strong> a drug if it were bound.<br />

B. Compensation Mutants<br />

Page 515<br />

More intriguing and more difficult to understand are the compensation mutants. These viruses bind<br />

compounds within their hydrophobic pockets, but are still able to replicate and are therefore able to<br />

compensate for the bound drug (Table 1, Figure 4). These mutations, like the mutations that convey acidresistance<br />

phenotypes, cluster about the hydrophobic pocket of VP1 [27,28,92].<br />

Two hypothesis have been presented in order to explain the behavior of the compensation mutants. The<br />

first suggests that there is a link between the binding of receptor <strong>by</strong> virion and its ability to bind<br />

compounds in the VP1 pocket. The second hypothesis suggests that the effect of a mutation on protein<br />

stability leads to the drug-resistant phenotype. These two hypotheses are not mutually exclusive.<br />

Linked-Binding Hypothesis<br />

The compensation mutants can be subdivided into those that occur on the canyon floor and those that are<br />

inside the VP1 hydrophobic pocket. The linked-binding hypothesis suggests that the subset of<br />

compensation mutations that occur on the canyon floor increase the binding affinity of receptor. An<br />

increase in binding for one of these mutants, Val1153 rarrow.gif Ile, has been observed [28]. This<br />

increased binding would then in turn cause a decrease in the affinity of the capsid for the drug. This<br />

would result in the drug being less effective in inhibiting uncoating.<br />

The second subset of compensation mutants are those that occur inside the hydrophobic pocket but do<br />

not interact with the receptor binding site. These mutations might directly decrease drug-binding affinity<br />

in the VP1 pocket due to decreased hydrophobic interactions caused <strong>by</strong> the smaller amino acid side<br />

chains [27]. In both subsets of mutations, the binding affinity of a drug is reduced either directly or<br />

through the effects of receptor binding. This decreased affinity for drug would be displayed as a drugresistant<br />

phenotype.<br />

Stabilization—Destabilization Hypothesis<br />

An alternative explanation for the behavior of the drug compensation mutants suggests that these<br />

mutants allow for increased conformational flexibility in the capsid. This increased conformational<br />

flexibility would compensate for the decrease in flexibility, which is manifested <strong>by</strong> decreased acid or<br />

thermal lability, induced <strong>by</strong> the binding of drug. The increase in flexibility would allow a<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_515.html [4/9/2004 12:45:42 AM]


Document<br />

conformational transition to occur, presumably near the GH loop, that would be required for viral<br />

uncoating. The conformational transition would lead to productive uncoating, even in the presence of<br />

bound drug.<br />

Page 516<br />

This hypothesis suggests that drug-resistant mutations would be of the type that destabilize protein<br />

structures. Such mutations are typically from larger to smaller side chains or from highly branched to<br />

less branched side chains [43–45]. Four of five observed compensation mutants are of this type (Table<br />

1). The remaining mutant, Val1153 rarrow.gif Ile, while a priori might not be predicted to destabilize<br />

the capsid, has been shown experimentally to decrease viral thermostability when compared with native<br />

HRV14 [28].<br />

The advantage of the stabilization-destabilization hypothesis is that it allows for a single mechanism of<br />

resistance for both compensation mutants occurring in the hydrophobic pocket as well as those that are<br />

outside the pocket near the receptor binding site. Both of these sites are close to the GH loop of VP1 that<br />

becomes disordered under acidic conditions. This stabilization—destabilization hypothesis also explains<br />

the behavior of acid-resistant mutants, which are predominately of the type which should stabilize a<br />

protein to conformational changes. Wild-type viruses have a preferred stability profile, which allows<br />

uncoating under the proper conditions. Mutations or addition of drug can perturb this profile, resulting in<br />

decreased replication. A second perturbation might then compensate for the first, restoring virulence<br />

(Figure 11). This behavior has been observed in HRV1A, in which mutants have been isolated that<br />

require presence of drug to replicate. These mutations are more acid-labile<br />

Figure 11<br />

The effects of various rhinovirus manipulations. The solid line depicts<br />

the native rhinovirus uncoating profile. Mutations and drugs can<br />

effect this profile to make the virus more or less stable to<br />

pH- or temperature-induced changes.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_516.html (1 of 2) [4/9/2004 12:45:46 AM]


Document<br />

Page 517<br />

than the wild-type virus. At the optimal drug concentration for growth, the pH-stability of these<br />

mutations is equivalent to that of native HRV1A in the absence of drug. Increasing drug concentration<br />

beyond that which is optimal for growth further increases the acid-stability of these mutant viruses so<br />

that they are more acid-stable than native virus without drug (Daniel Pevear, personal communication).<br />

C. Clinical Resistance and Virulence<br />

The clinical importance of these mutations has not been clearly demonstrated. In at least one case a drugresistant<br />

mutant was able to grow in vitro in single-cycle growth experiments as well as wild-type<br />

HRV14 [28]. Frequently, however, drug-resistant or acid-stable mutants will not grow as well in cell<br />

culture as native virus [93].<br />

The final analysis of the clinical importance of resistant mutants awaits the results of clinical trails. In<br />

the one study published to date, HRV2 mutants of the compensation type (resistant to the capsid-binding<br />

compound chalcone) were tested in healthy human volunteers [65]. The drug-resistant mutants caused<br />

significantly fewer colds than the normal virus. Mutants that require the presence of drug to grow did<br />

not cause any apparent disease. In this single case it appears that the mutant viruses were not as virulent<br />

as the parent strain.<br />

VI. Animal and Clinical Studies<br />

The idea that an inhibitor that binds to a nonenzymatic, nonreceptor site of a virion could inhibit viral<br />

replication in vivo would a priori be considered to be an unlikely scenario <strong>by</strong> most in the field of<br />

structure-<strong>based</strong> design. If this idea were proposed knowing only the structure of the native virus<br />

(especially HRV14, which has a closed-pocket conformation) skepticism would abound. This type of<br />

project arose not from a structure-<strong>based</strong> approach, but from the tried-and-true screening approach. The<br />

compounds were first shown to be effective in inhibiting viral replication in vitro [52]. This was<br />

followed <strong>by</strong> an experiment showing that these compounds could inhibit at least one enterovirus in a<br />

mouse model [94].<br />

In this study, suckling mice were inoculated with an echovirus and if untreated they developed a fatal<br />

paralytic illness. When treated either prophylactically or within a few days of the viral inoculation,<br />

virtually all of the mice were protected. This is dramatic proof of the concept that these compounds can<br />

inhibit picornaviral infection in vivo.<br />

There are important differences between mice and men and between enteroviral paralytic illness and<br />

rhinoviral upper respiratory tract infections.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_517.html [4/9/2004 12:45:48 AM]


Document<br />

Page 518<br />

The differences between mice and men should be obvious to the casual reader. Enterovirus-induced<br />

paralytic illnesses are systemic infections, and the virus must at some point move through the body and<br />

be subject to circulating drug. Upper respiratory infections resulting from rhino-or enteroviruses tend to<br />

be localized to the pharynx, which would require that drug titers remain high in this region of the body.<br />

In spite of these difficulties, there have been some dramatic successes in clinical trials of antirhinovirals,<br />

although clearly there are still obstacles to overcome. Two different classes of antipicornavirus<br />

compounds have been shown to be successful at inhibiting at least some types of infection when<br />

administered prophylactically. The WIN compound 54954, given orally, has been shown to inhibit<br />

Coxsackie virus A21 in humans [6]. The Jaansen compound R77975 has been shown to inhibit HRV9<br />

when administered intranasally 6 times daily [95,96], but R77975 has not been shown to inhibit<br />

infection when used therapeutically (after symptoms occur). Further clinical trials with WIN 54954 were<br />

suspended due to the developments of side effects. More recently VP 63843 has been shown to have<br />

dramatically improved effect when compared with WIN 54954 in the prophylactic treatment of<br />

Coxsackie virus A21 infection in humans and it will be tested therapeutically.<br />

The inability of R77975 to show a therapeutic effect is more likely due to poor pharmacodynamics<br />

rather than a fundamental inability for this compound to inhibit an established infection. The<br />

pharmacodynamic problem was witnessed in the R77975 study, which showed that administration<br />

intranasally 3 times a day did not yield prophylactic protection. Yet if administered 6 times daily the<br />

prophylaxis occurs. Therefore if more potent and metabolically stable compounds are found, more<br />

positive clinical results would be expected.<br />

VII. Future Directions<br />

The structures of the picornaviruses (native, with receptor bound, in the presence of acid, with a myriad<br />

of compounds bound, and of acid- and drug-resistant mutants) have yielded valuable information about<br />

possible molecular mechanisms for their uncoating. These same studies have suggested the mechanism<br />

<strong>by</strong> which these uncoating inhibitors work. A <strong>by</strong>-product of this research is the hypothesis that these<br />

compounds may mimic naturally occurring factors that occupy the VP1 pocket. The hunt for these<br />

natural compounds and their significance is underway.<br />

This understanding of the mechanism of action, as well as the structure-activity studies, have also<br />

yielded valuable information for future development of antipicornaviral drugs. The determination of<br />

compound's size, shape, and other physical requirements for activity will also be of assistance.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_518.html [4/9/2004 12:45:50 AM]


Document<br />

Page 519<br />

From the animal and human experiments it is clear that the cure for some common colds is within reach.<br />

These therapies are also quite likely to be efficacious in enteroviral diseases, for example in cardiogenic<br />

coxsackie virus infection [97].<br />

Acknowledgments<br />

We would like to acknowledge that there have probably been hundreds of people who have contributed<br />

to work reviewed here. We would like especially to thank the many people at ViroPharma, Sterling<br />

Winthrop, Purdue University, and the University of Wisconsin at Madison with whom we have worked<br />

over the years. We would also like to thank Dirksen Bussiere and Yvonne Martin for reading and editing<br />

this manuscript prior to submission.<br />

References<br />

1. Rueckert RR. Picornaviridae and their replication. In: Fields BN, Knipe DM, eds. Virology. New<br />

York: Raven Press, 1990:507–548.<br />

2. Sperber SJ, Hayden FG. Chemotherapy of rhinovirus colds. Antimicrob Agents Chemother 1988;<br />

32:409–419.<br />

3. Uncapher CR, DeWitt CM, Colonno RJ. The major and minor group receptor families contain all but<br />

one human rhinovirus serotype. Virology 1991; 180:814–817.<br />

4. Hamparian VV, Colonno RJ, Cooney MK, et al. A collaboration report: rhinoviruses—extension of<br />

the numbering system from 89 to 100. Virology 1987; 159:191–192.<br />

5. Muckelbauer JK, Kremer MK, Minor I, et al. The structure of coxsackievirus B3 at 3.5 Å resolution.<br />

<strong>Structure</strong> 1995; 3:653–667.<br />

6. Schiff GM, Sherwood JR, Young EC, Mason JL, Gamble JN. Prophylactic efficacy of WIN 54954 in<br />

prevention of experimental human coxsackie A21 infection and illness. Antivir Res 1992; 17 (Suppl<br />

1):92.<br />

7. Andries K, Dewindt B, Snoeks J, et al. Two groups of rhinoviruses revealed <strong>by</strong> a panel of antiviral<br />

compounds present sequence divergence and differential pathogenicity. J Virol 1990; 64(3):1117–1123.<br />

8. Flower RJ, Moncada S, Vane Jr. The pharmacologic basis of therapeutics. In: Gilman AG, Goodman<br />

LS, Gilman A, eds. Analgesic-antipyretics and anti-inflammatory agents; drugs employed in the<br />

treatment of gout. New York: Macmillan, 1080:682–728.<br />

9. Drake JW. Rates of spontaneous mutation among RNA viruses. Proc Natl Acad Sci USA 1993;<br />

90(9):4171–4175.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_519.html (1 of 2) [4/9/2004 12:45:52 AM]


Document<br />

10. Mast EE, Harmon MW, Gravenstein S, et al. Emergence and possible transmission of amantadineresistant<br />

viruses during nursing home outbreaks of influenza A (H3N2). Am J Epidem 1991;<br />

134(9):988–997.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_519.html (2 of 2) [4/9/2004 12:45:52 AM]


Document<br />

Page 520<br />

11. Acharya R, Fry E, Stuart D, Fox G, Rowlands D, Brown F. The three-dimensional structure of footand-mouth<br />

disease virus at 2.9 Å resolution. Nature 1989; 337:709–716.<br />

12. Hogle JM, Chow M, Filman DJ. Three-dimensional structure of poliovirus at 2.9 Å resolution.<br />

Science 1985; 229:1358–1365.<br />

13. Kim S, Smith TJ, Chapman MS, et al. The crystal structure of human rhinovirus serotype 1A<br />

(HRV1A). J Mol Biol 1989; 210:91–111.<br />

14. Luo M, Vriend G, Kamer G, et al. The atomic structure of Mengo virus at 3.0 Å resolution. Science<br />

1987; 235:182–191.<br />

15. Oliveira MA, Zhao R, Lee W-M, et al. The structure of human rhinovirus 16. <strong>Structure</strong> 1993;<br />

1:51–68.<br />

16. Rossmann MG, Arnold E, Erickson EW, et al. <strong>Structure</strong> of a human common cold virus and<br />

functional relationship to other picornaviruses. Nature 1985; 317:145–153.<br />

17. Smyth M, Tate J, Hoey E, Lyons C, Martin S, Stuart D. Implications for viral uncoating from the<br />

structure of bovine enterovirus. Structural Biology 1995; 2(3):224–231.<br />

18. Paul AV, Schultz A, Pincus SE, Oroszlan S, Wimmer E. Capsid protein VP4 of poliovirus is Nmyristoylated.<br />

Proc Natl Acad Sci USA 1987; 84:7827–7831.<br />

19. Chow M, Newman JFE, Filman D, Hogle JM, Rowlands DJ, Brown F. Myristoylation of<br />

picornavirus capsid protein VP4 and its structural significance. Nature 1987; 327:482–486.<br />

20. Greve JM, Davis G, Meyer AM, et al. The major human rhinovirus receptor is ICAM-1. Cell 1989;<br />

56:849–853.<br />

21. Hofer F, Gruenberger M, Kowalski H, et al. Members of the low density lipoprotein receptor family<br />

mediate cell entry of a minor-group common cold virus. Proc Natl Acad Sci USA 1994;<br />

91(5):1839–1842.<br />

22. Olson NH, Kolatkar PR, Oliveira MA, et al. <strong>Structure</strong> of the human rhinovirus complexed with its<br />

receptor molecule. Proc Natl Acad Sci USA 1993; 90:507–511.<br />

23. Giranda VL, Chapman MS, Rossmann MG. Modeling of the human intercellular adhesion<br />

molecule—1, the human rhinovirus major group receptor. Proteins 1990; 7:227–233.<br />

24. Smith TJ, Kremer MJ, Luo M, et al. The site of attachment in human rhinovirus 14 for antiviral<br />

agents that inhibit uncoating. Science 1986; 233:1286–1293.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_520.html (1 of 2) [4/9/2004 12:45:54 AM]


Document<br />

25. Arnold E, Rossmann MG. Analysis of the structure of the common cold virus, human rhinovirus 14,<br />

refined at a resolution of 3.0 Å. J Mol Biol 1990; 211:763–801.<br />

26. Towler DA, Gordon JI, Adams SP, Glaser L. The biology and<br />

enzymology of eukaryotic protein acylation. Annu Rev Biochem<br />

1988; 57:69–99.<br />

27. Hadfield AT, Oliveira Ma, Kim KH, et al. Structural studies on human rhinovirus 14 drug-resistant<br />

compensation mutants. J Mol Biol 1995; 253:61–73.<br />

28. Shepard DA, Heinz BA, Rueckert RR. WIN 52035-2 inhibits both attachment and eclipse of human<br />

rhinovirus 14. J Virol 1993; 67(4):2245–2254.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_520.html (2 of 2) [4/9/2004 12:45:54 AM]<br />

29. Fox MP,<br />

Otto MJ,<br />

McKinlay<br />

MA.<br />

Prevention<br />

of<br />

rhinovirus<br />

and<br />

poliovirus<br />

uncoating<br />

<strong>by</strong> WIN<br />

51711, a<br />

new<br />

antiviral<br />

drug.<br />

Antimicrob<br />

Agents<br />

Chemother<br />

1986;<br />

30:110–116.


Document<br />

30. Pevear DC, Fancher MJ, Felock PJ, et al. Conformational change in the floor of the human<br />

rhinovirus canyon blocks adsorption to HeLa cell receptors. J Virol 1989; 63:2002–2007.<br />

31. Lonberg-Holm K, Korant BD. Early interaction of rhinoviruses with host cells. J Virol 1972;<br />

9:29–40.<br />

32. Korant BD, Lonberg-Holm K, Yin FH, Noble-Harvey J. Fractionation of biologically active and<br />

inactive populations of human rhinovirus type 2. Virology 1975; 63:384–394.<br />

Page 521<br />

33. Everaert L, Vrusen R, Boeye A. Eclipse products of poliovirus after cold-synchronized infection of<br />

HeLa cells. Virology 1989; 171:76–82.<br />

34. Hooverlitty H, Greve JM. Formation of rhinovirus-soluble ICAM-1 complexes and conformational<br />

changes in the virion. J Virol 1993; 67(1):390–397.<br />

35. Neubauer C, Frasel L, Kuechler E, Blaas D. Mechanism of entry of human rhinovirus 2 into HeLa<br />

cells. Virology 1987; 158:255–258.<br />

36. Zeichhardt H, Wetz K, Willingmann P, Habermehl KO. Entry of poliovirus type 1 and mouse<br />

elberfeld (ME) virus into HEp-2 cells: receptor mediated endocytosis and endosomal or lysosomal<br />

uncoating. J Gen Virol 1985; 66:483–492.<br />

37. Gromier M, Wetz K. Kinetics of poliovirus uncoating in HeLa cells in a nonacidic environment. J<br />

Virol 1990; 68(8):3590–3597.<br />

38. Madshus IH, Olsnes S, Sandvig. Mechanism of entry into the cytosol of poliovirus type 1:<br />

requirement for low pH. J Cell Biol 1984; 98:1194–1200.<br />

39. Madshus IH, Olsnes S, Sandvig K. Different pH requirements for entry of the two picornaviruses,<br />

human rhinovirus 2 and murine encephalomyocarditis virus. Virology 1984; 139:346–357.<br />

40. Perez L, Carrasco L. Entry of poliovirus into cells does not require a low-pH step. J Virol 1993;<br />

67(8):4543–4548.<br />

41. Giranda VL, Heinz BA, Oliveira MA, et al. Acid-induced structural changes in human rhinovirus-<br />

14—possible role in uncoating. Proc Natl Acad Sci USA 1992; 89(21):10213–10217.<br />

42. Skern T, Torgersen H, Auer H, Kuechler E, Blaas D. Human rhinovirus mutants resistant to low pH.<br />

Virology 1991; 183:757–763.<br />

43. Alber T, Dao-pin S, Nye JA, Muchmore DC, Matthews BA. Temperature-sensitive mutations of<br />

bacteriophage T4 lysozyme occur at sites with low mobility and low solvent accessibility in the folded<br />

protein. Biochemistry 1987; 26:3754–3758.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_521.html (1 of 2) [4/9/2004 12:45:56 AM]


Document<br />

44. Argos P, Rossmann MG, Grau UM, Zuber H, Frank G, Tratschin JD. Thermal stability and protein<br />

structure. Biochemistry 1979; 18:5698–5703.<br />

45. Matthews BA. Genetic and structural analysis of the protein stability Problem. Biochemistry 1987;<br />

26:6885–6888.<br />

46. Perutz MF, Raidt H. Stereochemical basis of heat stability in bacterial ferredoxins and in<br />

haemoglobin A2. Nature 1975; 255:256–259.<br />

47. Gruenberger M, Pevear D, Diana GD, Kuechler E, Blaas D. Stabilization of human rhinovirus<br />

serotype 2 against pH-induced conformational change <strong>by</strong> antiviral compounds. J Gen Virol 1991;<br />

72:431–33.<br />

48. Fricks CE, Hogle JM. Cell-induced conformational change in poliovirus: Externalization of the<br />

amino terminus of VP1 is responsible for liposome binding. J Virol 1990; 64:1934–1945.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_521.html (2 of 2) [4/9/2004 12:45:56 AM]


Document<br />

Page 522<br />

49. Roivainen M, Piirainen L, Rysa T, Narvanen A, Hovi T. An immunodominant N-terminal region of<br />

VP1 protein of poliovirion that is buried in the crystal structure can be exposed in solution. Virology<br />

1993; 195:762–765.<br />

50. Yafal AG, Kaplan G, Racaniello VR, Hogle JM. Characterization of poliovirus conformational<br />

alteration mediated <strong>by</strong> soluble cell receptors. Virology 1993; 197:501–505.<br />

51. Stroud R. Ion channel forming colicins. Curr Opin Struc Biol 1995; 5:514–520.<br />

52. McSharry JJ, Caliguiri LA, Eggers HJ. Inhibition of uncoating of poliovirus <strong>by</strong> arildone, a new<br />

antiviral drug. Virology 1979; 97:307–315.<br />

53. Abel MD, Cameron AD, Ha CM, et al. Novel azoylalkyloxy compounds with antipicornaviral<br />

activity. Antivir Chem Chemother 1995; 6(4):245–254.<br />

54. Diana GD, Cutcliffe D, Volkots DL, et al. Antipicornavirus activity of tetrazole analogues related to<br />

disoxaril. J Med Chem 1993; 36(22):3240–3250.<br />

55. Giranda VL, Russo GR, Felock PJ, et al. The structures of four methyltetrazole-containing<br />

compounds in human rhinovirus serotype 14. Acta Cryst D 1995; D51:496–503.<br />

56. Kim KH, Willingman P, Gong ZX, et al. A comparison of the anti-rhinoviral drug binding pocket in<br />

HRV14 and HRV1A. J Mol Biol 1993; 230:206–227.<br />

57. Andries K, DeWindt B, Snoeks J, Willebrords R, VanEemeren K. Anthirhinovirus spectrum and<br />

mechanism of action of R77975. Antivir Res 1991; 1 Suppl:98.<br />

58. Chapman MS, Minor I, Rossmann MG, Diana GD, Andries K. Human rhinovirus 14 complexed<br />

with antiviral compound R 61837. J Mol Biol 1991; 217:455–463.<br />

59. Moeremans M, Raermaeker M, Daneels G, et al. Study of the binding of R 61837 to human<br />

rhinovirus 9 and immunobiochemical evidence of capsid-stabilizing activity of the compound.<br />

Antimicrob Agents Chemother 1992; 36(2):417–424.<br />

60. Tisdale M, Selway JWT. Effect of dichloroflavan (BW683C) on the stability and uncoating of<br />

rhinovirus type 1B. Antimicrob Chemother 1984; 14(Suppl A):97–105.<br />

61. Ash RJ, Parker RA, Hagan AC, Mayer GD. RMI 15,731 (1-[5-tetradecyloxy-2-furanyl]ethanone), a<br />

new antirhinovirus compound. Antimicrob Agents Chemother 1979; 16:301–305.<br />

62. Zhang AQ, Nanni RG, Li T, et al. <strong>Structure</strong> determination of antiviral compound SCH-38057<br />

complexed with human rhinovirus-14. J Mol Biol 1993; 230(3):857–867.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_522.html (1 of 2) [4/9/2004 12:46:14 AM]


Document<br />

63. Rozhon E, Cox S, Buontempo P, et al. SCH-38057—a picornavirus capsid-binding molecule with<br />

antiviral activity after the initial stage of viral uncoating. Antivir Res 1993; 21(1):15–35.<br />

64. Ishitsuka H, Ninomiya YT, Ohsawa C, Fujiu M, Suhara Y. Direct and specific inactivation of<br />

rhinovirus <strong>by</strong> chalcone Ro 09-0410. Antimicrob Agents Chemother 1982; 22(4):617–621.<br />

65. Yasin SR, al-Nakib W, Tyrrell AJ. Pathogenicity for humans of human rhinovirus type 2 mutants<br />

resistant to or dependent on chalcone Ro 09–0410. Antimicrob Agents Chemother 1990; 34:963–966.<br />

66. Ninomiya Y, Shimma N, Ishitsuka H. Comparative studies on the antirhinovirus activity and mode<br />

of action of the rhinovirus capsid binding agents, chalcone amides. Antivir Res 1990; 13:61–74.<br />

67. Kenny MT, Dulworth JK, Bargar TM, Torney HL, Graham MC, Manelli AM. In vitro<br />

antipicornavirus activity of the 6-substituted 2-(3',4'-dichlorophenoxy)-2H-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_522.html (2 of 2) [4/9/2004 12:46:14 AM]


Document<br />

pyrano[2,3-b] pyridines MDL 20,610, MDL 20,646, and MDL 20,957. Antimicrob Agents<br />

Chemother 1986; 30:516–518.<br />

Page 523<br />

68. Kelly JL, Linn JA, Selway JWT. Antirhinovirus activity of 6-anilino-9-benzyl-2-chloro-9H-purines.<br />

J Med Chem 1990; 33:1360–1363.<br />

69. Alarcon B, Zerial A, Dupiol C, Carrasco L. Antirhinovirus compound 44081 R. P. inhibits virus<br />

uncoating. Antimicrob Agents Chemother 1986; 30(1):31–34.<br />

70. Filman DJ, Syed R, Chow M, Mcadam AJ, Minor PD, Hogle JM. Structural factors that control<br />

conformational transitions and serotype specificity in type 3 poliovirus. EMBO 1989; 8(5):1567–1579.<br />

71. Yeates TO, Jacobson DH, Martin A, et al. Three-dimensional structure of a mouse adapted<br />

type2/type1 chimera. EMBO 1991; 10:2331–2341.<br />

72. Smyth M, Fry E, Stuart D, Lyons C, Hoey E, Martin SJ. Preliminary crystallographic analysis of<br />

bovine enterovirus. J Mol Biol 1993; 231(3):930–932.<br />

73. Grant RA, Hiremath CN, Filman DJ, Syed R, Andries K, Hogle JM. <strong>Structure</strong>s of poliovirus<br />

complexes with anti-viral drugs: Implications for viral stability and drug design. Curr Biol 1994;<br />

4:784–797.<br />

74. Eriksson AE, Baase WA, Wozniak JA, Matthews BA. A cavity-containing mutant of T4 lysozyme is<br />

stabilized <strong>by</strong> buried benzene. Nature 1992; 355:371–373.<br />

75. Sattar SA, Jacobsen H, Springthorpe VS, Cusack TM, Rubino JR. Chemical disinfection to interrupt<br />

transfer of rhinovirus type-14 from environmental surfaces to hands. Appl Environ Microbiol 1993;<br />

59(5):1579–1585.<br />

76. Hendley JO, Wenzel RP, Gwaltney JMJ. Transmission of rhinovirus colds <strong>by</strong> self-inoculation. N<br />

Engl J Med 1973; 288:1361–1364.<br />

77. Jaeger EP, Pevear DC, Felock PJ, Russo GR, Treasurywala AM. Genetic algorithm <strong>based</strong> method to<br />

design a primary screen for antirhinovirus agents. Am Chem Soc Symp Ser 1995; 589:139–155.<br />

78. Fox MP, McKinlay MA, Diana GD, Dutko FJ. The binding affinities of structurally related human<br />

rhinovirus capsid-binding compounds correlate with antiviral activity. Antimicrob Agents Chemother<br />

1991; 35:1040–1047.<br />

79. Bibler-Muckelbauer JK, Kremer MJ, Rossmann MG, et al. Human rhinovirus 14 complexed with<br />

fragments of active antiviral compounds. Virology 1994; 201:360–369.<br />

80. Diana GD, Cutcliffe D, Ogles<strong>by</strong> RC, et al. Synthesis and structure-activity studies of some<br />

disubstituted phenylisoxazoles against human picornaviruses. J Med Chem 1989; 32:450–455.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_523.html (1 of 2) [4/9/2004 12:46:42 AM]


Document<br />

81. Diana GD, McKinlay MA, Otto MJ, Akullian V, Ogels<strong>by</strong> C. [[(4,5-Dihydro-2oxazolyl)phenoxy]alkyl]isoxazoles.<br />

Inhibitors of picornavirus uncoating. J Med Chem 1985;<br />

28:1906–1910.<br />

82. Diana GD, Ogles<strong>by</strong> RC, Akullian V, et al. <strong>Structure</strong>-activity studies of 5-[[4-(4,5-Dihydro-2oxazolyl)phenoxy]alkyl]-3-methylisoxazoles:<br />

inhibitors of picornavirus uncoating. J Med Chem 1987;<br />

30:383–388.<br />

83. Diana GD, Otto MJ, Treasurywala AM, et al. Enantiomeric effects of homologues of disoxaril in the<br />

inhibitory activity against human rhinovirus 14. J Med Chem 1988; 31:540–544.<br />

84. Horsnell C, Gama RE, Hughes PJ, Stanway G. Molecular relationships between 21 human<br />

rhinovirus serotypes. J Gen Virol 1995; 76:2549–2555.<br />

85. Diana GD, Kowalczyk P, Treasurywala AM, Ogles<strong>by</strong> RC, Pevear DC, Dutko FJ. CoMFA analysis<br />

of the interactions of antipicornavirus compounds in the binding pocket of human rhinovirus-14. J Med<br />

Chem 1992; 35:1002–1008.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_523.html (2 of 2) [4/9/2004 12:46:42 AM]


Document<br />

Page 524<br />

86. Abraham MH, Duce PP, Prior DV, Barratt DG, Morris JJ, Taylor PJ. Hydrogen bonding. Part 9.<br />

Solute proton donor and proton acceptor scales for use in drug design. J Chem Soc Perkin Trans II 1989;<br />

1355–1375.<br />

87. Abraham M, Grellier P, Prior D, Morris J, Taylor P. Hydrogen bonding. Part 10. A scale of solute<br />

hydrogen-bond basicity using log K values for complexation in tetrachloromethane. J Chem Soc Perkin<br />

Trans II 1990; 521–529.<br />

88. Bailey TR, Diana GD, Kowalczyk PJ, et al. Antirhinoviral activity of heterocyclic analogs of WIN-<br />

54954. J Med Chem 1992; 35(24):4628–4633.<br />

89. Bailey TR, Diana GD, Mallamo JP, et al. An evaluation of the antirhinoviral activity of acylfuran<br />

replacements for 3-methylisoxazoles. Are 2-acetylfurans bioisosteres for 3-methylisoxazoles. J Med<br />

Chem 1994; 37:4177–4184.<br />

90. Mallamo JP, Diana GD, Pevear DC, et al. Conformationally restricted analogues of disoxaril—A<br />

comparison of the activity against human rhinovirus type-14 and type-1A. J Med Chem 1992;<br />

35(25):4690–4695.<br />

91. Badger J, Krishnaswamy S, Kremer MJ, et al. Three-dimensional structures of drug-resistant<br />

mutants of human rhinovirus 14. J Mol Biol 1989; 207:163–174.<br />

92. Heinz BA, Rueckert RR, Shepard DA, et al. Genetic and molecular analyses of spontaneous mutants<br />

of human rhinovirus 14 that are resistant to an antiviral compound. J Virol 1989; 63(6):2476–2485.<br />

93. Colonno RJ, Condra JH, Mizutani S, Callahan PL, Davies ME, Murcko MA. Evidence for direct<br />

involvement of the rhinovirus canyon in receptor binding. Proc Natl Acad Sci USA 1988;<br />

85:5449–5453.<br />

94. McKinlay MA, Frank JA, Steinberg BA. Use of WIN 51711 to prevent echovirus type 9-induced<br />

paralysis in suckling mice. J Infec Dis 1986; 154:676–681.<br />

95. Al-Nakib W, Higgins PG, Barrow Gl, et al. Suppression of colds in human volunteers challenged<br />

with rhinovirus <strong>by</strong> a new synthetic drug (R61837). Antimicrob Agents Chemother 1989; 33:522–525.<br />

96. Hayden FG, Andries K, Janssen PAJ. Safety and efficacy of intranasal Pirodavir (R77975) in<br />

experimental rhinovirus infection. Antimicrob Agents Chemother 1992; 36(4):727–732.<br />

97. Ilback N-G, Wesslen L, Pauksen K, Stalhandske T, Friman G, Fohlman J. Effects of the antiviral<br />

WIN 54954 and the immune modulator LS 2616 on cachectin/TNF and gamma-interferon responses<br />

during viral heart disease. Scand J Infect Disease Suppl 1993; 88:117–123.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_524.html [4/9/2004 12:47:17 AM]


Document<br />

20<br />

The Integration of <strong>Structure</strong>-Based <strong>Design</strong> and Directed Combinatorial<br />

Chemistry for New Pharmaceutical Discovery<br />

Roger Bone and F. Raymond Salemme<br />

3-Dimensional Pharmaceuticals, Inc., Exton, Pennsylvania<br />

I. New Challenges For <strong>Drug</strong> Discovery<br />

Page 525<br />

Rapid advances in cell and molecular biology, together with comprehensive genome sequencing efforts,<br />

are providing detailed correlations between specific pathological conditions and discrete molecular<br />

targets. The same tools of recombinant DNA technology that identify key gene targets also provide the<br />

means for target biosynthesis in quantities sufficient for both the high-throughput screening of<br />

compound libraries for leads and the structure-<strong>based</strong> refinement of leads using x-ray crystallography and<br />

NMR spectroscopy.<br />

The rapid expansion in genomics data makes it inevitable that targets will be identified whose functions<br />

are so poorly understood that the most rapid and efficient way to establish their involvement in disease<br />

will be through the development of prototype drugs. New approaches to drug discovery that are able to<br />

integrate many different types of information are needed to seize this opportunity and drive an optimally<br />

efficient discovery process.<br />

In what follows, we describe a practical integration of structure-<strong>based</strong> design and combinatorial<br />

chemistry aimed at enhancing the effectiveness of both approaches. Three-dimensional structures<br />

provide the information required to most efficiently direct the design and optimization of new lead<br />

compounds. Combinatorial chemistry technologies, which are <strong>based</strong> on high-throughput automated<br />

methods of chemical synthesis, produce new classes of lead compounds and permit the rapid generation<br />

of structure—activity relationships<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_525.html [4/9/2004 12:47:27 AM]


Document<br />

Figure 1<br />

An integrated technology for drug discovery combines the<br />

precision of structure-<strong>based</strong> design with the parallelism of<br />

combinatorial synthesis. Chemi-informatics systems track<br />

and integrate all data emerging from the discovery cycle.<br />

Page 526<br />

(SAR). Chemi-informatics plays a key role in this integration <strong>by</strong> assuring that properties important in<br />

drug development are both factored into compound design and cumulatively tracked throughout the<br />

discovery process (Figure 1). The product of this approach is a permanently useful set of drug-design<br />

parameters. The integration of these technologies promises to produce an increase in the efficiency of<br />

drug discovery and may ultimately offer a means for reducing the aggregate failure rate of compounds<br />

selected for development. One paradigm for achieving the integration of these technologies is described<br />

below.<br />

II. <strong>Structure</strong>-Based <strong>Design</strong><br />

<strong>Structure</strong>-<strong>based</strong> drug design has become a highly developed technology that is in active use in most<br />

major pharmaceutical companies. <strong>Structure</strong>-<strong>based</strong> design is an iterative process in which lead<br />

compounds identified <strong>by</strong> screening, de novo design, or mechanism-<strong>based</strong> features are systematically<br />

elaborated to improve potency and specificity [1,2]. The process involves successive rounds of structure<br />

determination of lead—target complexes, design of lead modifications using molecular modeling tools,<br />

synthesis of new drug leads, and measurement of the chemical and biological properties of the modified<br />

leads using screens for target<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_526.html (1 of 2) [4/9/2004 12:47:32 AM]


Document<br />

Page 527<br />

function. Iterative refinement and optimization of drug leads is an effective strategy for generating<br />

potent preclinical candidates. <strong>Structure</strong>-<strong>based</strong> design can also be used to design new chemical classes of<br />

compounds that present similar substituents to the target using a template or scaffold which is<br />

chemically distinct from previously characterized leads [3,4].<br />

<strong>Structure</strong> determination typically relies on x-ray crystallography or high-field nuclear magnetic<br />

resonance to directly visualize the 3-dimensional structure of a molecular target and the structures of<br />

complexes of the target with drug leads. Alternately, many targets fall into identifiable classes that<br />

frequently enable the development of homology models of the 3-dimensional target structure or a<br />

mechanism-<strong>based</strong> strategy for drug-lead generation. Ongoing genome sequencing efforts have led to the<br />

identification of hundreds of potential therapeutic targets, many of which represent possible sources of<br />

crossover pharmacology. Homology modeling is a key feature of an integrated drug discovery effort<br />

because it allows this genomics information to be utilized early in the development of target ligands or<br />

in the engineering of ligand specificity.<br />

Although structure-<strong>based</strong> design is an effective technology, current limitations center on the inability to<br />

quantitatively predict how specific modifications of the lead will actually affect ligand binding affinity<br />

[5,6]. This reflects the complexity of the drug-binding process and our inability to accurately predict the<br />

conformational response of macromolecular structures to ligand binding. In addition, we have only<br />

limited ability to accurately calculate molecular energy parameters or to accurately estimate the effects<br />

of factors such as polarizability, solvation, and entropy that may have an important influence on drugbinding<br />

energetics. Although computational methods will continue to improve, most design work (and<br />

algorithms) still relies heavily on heuristic rules (Figure 2) that have been developed through experience<br />

and that guide the structural and medicinal chemists in the systematic modification of lead compounds<br />

[7]. As a practical consequence, many cycles of serial lead modification are required in order to produce<br />

molecules of suitable potency and specificity to be considered preclinical drug candidates.<br />

Structural information can increase the efficiency with which pharmacokinetic or toxicological liabilities<br />

in lead compounds are eliminated <strong>by</strong> suggesting where compounds can be modified so as to alter drug<br />

properties without affecting target potency. Structural data can also be used to direct de novo design of<br />

alternate and distinct chemical classes of lead compounds, each of which might be expected to have a<br />

different pharmacological profile [3,4]. New chemical compound classes can also be designed from<br />

existing lead compounds <strong>by</strong> recombining substituents and core regions (scaffolds) from existing lead<br />

compounds. Chemically distinct lead series can then be optimized in parallel so that when a preclinical<br />

candidate is found to have inadequate drug properties, a backup is immediately available for preclinical<br />

evaluation. As<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_527.html [4/9/2004 12:47:45 AM]


Document<br />

Figure 2<br />

Some heuristic rules frequently used in structure-<strong>based</strong> drug<br />

design. The positions of “bound” water molecules are key<br />

indicators for lead modification sites.<br />

outlined below, both the scaffold modification and structural recombination strategies are key<br />

components of an integrated drug discovery technology that combines structure-<strong>based</strong> design and<br />

combinatorial chemistry.<br />

III. Combinatorial Chemistry<br />

Page 528<br />

Combinatorial chemical technology enables the parallel synthesis of organic compounds through the<br />

systematic addition of defined chemical components using highly reliable chemical reactions and robotic<br />

instrumentation [8–11]. Large libraries of compounds result from the combination of all possible<br />

reactions that can be done at one site with all the possible reactions that can be done at a second, third,<br />

or greater number of sites. Combinatorial chemical methods can potentially generate tens to hundreds of<br />

millions of new chemical compounds as mixtures, attached to a solid support, or as individual<br />

compounds. The first combinatorial libraries with millions of members were oligopeptide and<br />

oligonucleotide libraries for which reliable and versatile chemical reactions already existed [10,12,13].<br />

However, these libraries were not generally useful as a source of leads for small-molecule<br />

pharmaceuticals due to the relatively high molecular weight of the resulting leads and other limitations<br />

of oligonucleotides and oligopeptides as drugs. More recently, “drug-like” libraries have been produced<br />

that offer much greater utility as a source of drug leads (Figure 3). However, practical limitations of the<br />

versatility and reliability of the chemical reactions used to make these libraries have resulted in<br />

somewhat smaller sets of compounds [14–18].<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_528.html (1 of 2) [4/9/2004 12:48:37 AM]


Document<br />

Figure 3<br />

Generation of combinatorial libraries.<br />

Traditional combinatorial libraries <strong>based</strong> on<br />

polypeptides or polynucleotides were built up<br />

<strong>by</strong> oligomer condensation and achieved a<br />

high level of diversity through conformational<br />

flexibility. Small-molecule libraries of drug-like<br />

molecules can be built up using a variety of<br />

strategies that focus on maximizing chemical<br />

diversity while (generally) minimizing<br />

conformational flexibility.<br />

Page 529<br />

From the drug discovery perspective, large combinatorial libraries have the same utility as conventional<br />

pharmaceutical or natural-product compound libraries, i.e., as a source of leads for new drugs. The<br />

design of large combinatorial libraries is driven <strong>by</strong> the requirement that individual reactions be highly<br />

reliable and versatile, while producing libraries with the highest possible degree of chemical diversity.<br />

Individual steps must be optimized, the compatibility of building blocks must be examined thoroughly<br />

and the synthesis must be automated. As a consequence, a significant investment in time and resources<br />

must be made before a library can actually be produced. <strong>Design</strong>ed and used in this way, large<br />

combinatorial libraries provide a new source of screening leads. However, libraries with maximum<br />

diversity, which may be used for “blind” lead discovery, do not address the principle rate-limiting step<br />

in drug discovery, the elaboration of a suitable SAR around the lead compound after an active lead has<br />

been discovered.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_529.html (1 of 2) [4/9/2004 12:49:08 AM]


Document<br />

While initial combinatorial chemical strategies have focused on the exhaustive synthesis of all members<br />

of a given library, it is inevitable that advances in automated chemistry and equipment will soon make it<br />

possible to synthesize a vastly greater number of compounds than it will be practical to<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_529.html (2 of 2) [4/9/2004 12:49:08 AM]


Document<br />

Page 530<br />

screen. For example, for a simple library created using amine and acid condensations onto a given amino<br />

acid scaffold, over a million compounds can be produced from commercially available reagents. If this<br />

simple library is expanded to incorporate the hundreds of commercially available or easily synthesized<br />

amino acids scaffolds, the number of compounds that can possibly be made increases to greater than<br />

10 8. Although biological, chemical, and physical assays can be automated so that hundreds of thousands<br />

to millions of compounds can be screened annually, the process is expensive and the reliability of the<br />

screening process decreases when measurements are made on mixtures of multiple compounds. Various<br />

estimates of the number of drug-like molecules with molecular weight (MW)


Document<br />

Figure 4<br />

The information flow in a drug discovery process that<br />

connects elements of structure-<strong>based</strong> design and<br />

combinatorial chemistry for<br />

drug discovery.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_531.html (1 of 2) [4/9/2004 12:49:36 AM]<br />

Page 531


Document<br />

ciency of drug discovery (Figure 4). The process begins with the knowledge-<strong>based</strong> design of a library<br />

template or scaffold and involves the synthesis of small subsets of library members. As with structure<strong>based</strong><br />

design approaches, the process is an iterative one in which SAR data and structural data guide not<br />

only the iterative selection of library members for testing, but also the design of later generation library<br />

scaffolds. The integrated process is described in detail in the next few paragraphs.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_531.html (2 of 2) [4/9/2004 12:49:36 AM]


Document<br />

Figure 5<br />

Generation of a virtual combinatorial library <strong>by</strong> finding<br />

substituents of a custom scaffold that can be<br />

accommodated in the binding site of a molecular target<br />

or meet other 3D structural criteria. Once the virtual<br />

library is synthesized in the computer, individual<br />

members can be selected using structural or additional<br />

criteria and synthesized using automated equipment.<br />

Page 532<br />

The first step in the process is the generation of a chemical template or scaffold that can be derivatized<br />

at multiple sites using reliable chemical reactions to produce a large combinatorial library (Figure 5).<br />

This custom chemical template is designed <strong>based</strong> on the structure of the target using the same heuristic<br />

set of rules used for traditional structure-<strong>based</strong> drug design. Useful 3-dimensional pharmacophore<br />

models are best derived from crystallographic or nuclear magnetic resonance structures of the target, but<br />

can also be derived from homology models <strong>based</strong> on the structures of related targets or 3-dimensional<br />

quantitative structure-activity relationships (3D QSAR) derived from a previously discovered series of<br />

active compounds [20]. In addition, the mechanism of action of the target or any other information that<br />

exists regarding the target or the target class can be used in the design to maximize the chances of<br />

finding hits [21,22]. The combinatorial libraries are designed so that a few thousand to millions of<br />

discrete molecules can be produced <strong>by</strong> reaction of the custom-designed template with appropriate<br />

proprietary and commercially available chemical building blocks.<br />

The next step in the process involves the implementation of the automated synthesis and generation of<br />

the library. Significant lead time is anticipated before a library can be produced because even the most<br />

reliable chemical reactions require optimization if they are to be carried out <strong>by</strong> a robot, particularly if<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_532.html [4/9/2004 12:50:09 AM]


Document<br />

Page 533<br />

the reactions are to be implemented with the template attached to a solid support. When the synthesis is<br />

optimized and fully automated, thousands to millions of compounds are accessible.<br />

One of the key features of this process is that all of the compounds that can potentially be made <strong>by</strong><br />

elaboration of a custom scaffold are first “made” in virtual form in a computer. Each of the chemical<br />

reactions required to create the library is encoded in a program that then systematically combines the<br />

template and the building blocks to create a 2-dimensional representation of each member of the library.<br />

Next, 3-dimensional representations are created and molecular-property descriptors are calculated for<br />

each member of the library. Molecular property descriptors encompass molecular connectivity, dipole<br />

moments, calculated partition coefficients, and many other calculable molecular properties. The virtual<br />

library of compounds can then be computationally screened and the library members ranked according<br />

to their ability to interact with the target receptor or 3-dimensional pharmacophore model [23–25]. The<br />

compounds can also be ranked <strong>by</strong> their inability to interact with any number of alternative targets whose<br />

inhibition is undesirable, or their ability to meet any range of desired chemical or physical properties<br />

that may be important in drug pharmacology. Alternately or additionally, compounds can be ranked<br />

according to their ability to span or sample the physical-chemical property space to produce the most<br />

diverse set of compounds for initial screening [26,27].<br />

Small sublibraries of the large virtual library that best satisfy the selection criteria are chemically<br />

synthesized using automated methods and then the biological and/or chemical properties of each<br />

compound are measured using automated assays. The SAR data that emerge from the assays are stored<br />

in a central database and used in the selection process to drive additional rounds of sublibrary selection,<br />

synthesis, and assay. Multiple mathematical models are developed to correlate the computed structure<br />

and properties of each synthesized library member with the biological, chemical, or physical properties<br />

that are measured during each cycle of testing [28]. A key feature of this approach is that compounds<br />

can be selected not only on the basis of which are predicted to perform the best in the target assay but<br />

also on the basis of their ability to perform the best in the target assay but also on the basis of their<br />

ability to distinguish between or validate the SAR models that are generated. The observed and<br />

predicted properties of a given sublibrary are compared so that the set of assumptions upon which<br />

property refinement is <strong>based</strong> is constantly updated. In principle, this process can become completely<br />

automated so that leads are discovered and refined with very little manual intervention [28].<br />

To achieve the greatest improvements in drug discovery efficiency, empirical data of various kinds must<br />

be collected throughout the iterative refinement process. It is desirable to obtain more accurate<br />

dissociation constants rather than IC 50 or single-point percent-inhibition values. In addition, the 3dimensional<br />

structures of interesting target—inhibitor complexes are determined<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_533.html [4/9/2004 12:50:11 AM]


Document<br />

Figure 6<br />

The combination of structure-<strong>based</strong> design and combinatorial<br />

chemistry can facilitate the generation of recombined<br />

compounds to rapidly produce potent compounds with<br />

maximum chemical and structural diversity.<br />

to provide information regarding how substituents are interacting with the target and when binding<br />

modes change. This information may be essential if SAR models are to remain predictive over large<br />

numbers of compounds.<br />

Page 534<br />

The integrated drug discovery process utilizes as much information as is available to find and optimize<br />

initial lead compounds. Because the automated synthesis of large libraries of compounds requires<br />

reliable and versatile chemical reactions, initial libraries are designed to discover new chemical lead<br />

classes or to develop SAR models. Both library and substituent designs evolve as hits are encountered<br />

and the structures of target—inhibitor complexes are determined. Single modifications on both the<br />

template and substituents may be made during the process <strong>based</strong> on the structures of target—lead<br />

complexes. When sufficient SAR data has accumulated and the structures of target complexes with key<br />

templates and substituents have been determined, potent compounds with the desired “drug like”<br />

compositions are designed and synthesized using a structure-<strong>based</strong> recombination strategy (Figure 6).<br />

The integration of these recent advances in drug discovery offers the possibility to substantially decrease<br />

the time required to find initial leads and develop them into prototype drugs. Similarly, the interval<br />

required to produce preclinical development compounds and backup candidates is also expected to<br />

shrink. Perhaps most interesting is the potential of this integrated technology, which is ultimately <strong>based</strong><br />

on an abstract information model of the biological process, to dramatically increase the reliability of<br />

successful drug development<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_534.html (1 of 2) [4/9/2004 12:50:22 AM]


Document<br />

Page 535<br />

<strong>by</strong> factoring in and refining important drug properties concurrently with the optimization of drug affinity<br />

and specificity for a specific molecular receptor. This objective is achieved <strong>by</strong> integrating assays and<br />

computational models that relate to important drug development issues (e.g., oral absorption, optimal<br />

pharmacokinetics, minimal toxicity, etc.) directly into the iterative design process.<br />

V. Chemi-Informatics<br />

The development of methods that can optimize the parallel refinement of drug potency and<br />

pharmacological properties is a key objective in enhancing the efficiency and productivity of drug<br />

discovery. To address this problem and to handle the dramatic increase in experimental data generated<br />

using robotic synthesis and assay methodologies, advanced informatics systems are required to collect<br />

and exploit data relating to the properties of the chemicals being produced. These systems will amplify<br />

the synergy between structure-<strong>based</strong> design and combinatorial chemistry and provide a means for<br />

reducing the aggregate failure rate of development candidates.<br />

The poor success rate for preclinical development candidates (about 1 in 20) results from our limited<br />

ability to predict such drug properties as intestinal absorption, excretion, metabolism, toxicity, efficacy,<br />

and side effects. Rendering the prediction more difficult is the certainty that these drug properties are<br />

composite properties that result from the operation of many biological processes. Recently, progress has<br />

been made towards understanding the underlying molecular basis for some of the individual components<br />

that contribute to the observed drug properties. For instance, the absorption of compounds into Caco-2<br />

cells in culture may be predictive of intestinal absorption [29–31]. Retention times of compounds during<br />

artificial membrane chromatography has also been correlated with oral absorption [32]. Recently,<br />

molecular transporters have been identified, cloned, and characterized that are responsible for the<br />

absorption of dipeptides and the excretion of organic cations [33,34]. The enzymatic basis for the<br />

metabolism of xenobiotics (cytochrome P450, glutathione transferase, etc.) has been known for some<br />

time, at least in part. In fact, many of the individual components of the composite biological processes<br />

can be developed into high-throughput assays, providing the opportunity to collect SAR data and<br />

develop SAR models.<br />

These observations, together with the integrated structure-<strong>based</strong> design and combinatorial chemical<br />

technology described here, define a comprehensive new strategy for drug discovery that has the<br />

potential to reduce the aggregate failure rate for development candidates (Figure 4). The creation of<br />

virtual libraries of compounds that are readily accessible through the use of automated<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_535.html [4/9/2004 12:50:24 AM]


Document<br />

Page 536<br />

synthesis methods allows for the first time an extensive and systematic investigation of structure-activity<br />

relationships related to drug properties. Small sublibraries can be selected, synthesized, and screened in<br />

high-throughput assays with the goal of developing predictive SAR models for individual components<br />

of the biological processes responsible for pharmacokinetic and toxicological properties of drug<br />

candidates. Eventually, it will be possible to assemble the predictive SAR models for various component<br />

processes to produce predictive models for bioavailability and toxicology. At this point it will be<br />

possible to use these models to guide the selection of sublibraries directed against therapeutic targets so<br />

that library members with the most “drug like” properties and the fewest liabilities are chosen. Steering<br />

away from compounds with undesired properties will focus the selection process on molecules with an<br />

improved probability of successful development. This strategy permits the simultaneous refinement of<br />

multiple chemical properties in addition to target efficacy and will shorten the drug-discovery process.<br />

Developing a system capable of collecting multivariate SAR data and exploiting the data to produce<br />

predictive SAR models is a major systems integration task. However, recent advances in computers,<br />

operating systems, and computational chemical tools now enables the implementation of a system that<br />

can track compounds, store chemical property data in a comprehensive relational database, and operate<br />

on virtual libraries in an iterative fashion to develop SAR models and refine chemical properties [28].<br />

Tools for the production of virtual libraries have been developed <strong>by</strong> several groups and large virtual<br />

libraries can be produced within a few days using high-power computer workstations. A variety of tools<br />

also exist for selecting compounds <strong>based</strong> on criteria such as the ability to fit a target receptor [23],<br />

similarity or diversity [26,27], or any number of other properties that might be important for interaction<br />

with the target receptor. Statistical programs also exist that are up to the task of developing SAR models<br />

[20]. To obtain SAR models that have utility in the drug-discovery process it will be necessary to collect<br />

data on the properties of large numbers of compounds so that the models are predictive across diverse<br />

chemical classes of lead molecules. This requires implementation of a comprehensive relational<br />

database to collect and correlate SAR data.<br />

The challenge is to integrate these tools into a system that can collect and operate on vast arrays of<br />

chemical-property data in an intelligent fashion to direct and refine the properties of robotically<br />

synthesizable compounds. One approach that has been used successfully is to create a hierarchical clientserver<br />

system with a powerful central selection algorithm (Figure 4) that chooses compounds from<br />

virtual libraries <strong>based</strong> on their computed properties, supervises data analysis, and integrates results from<br />

experimental measurements [28,35]. This system, which has been termed DirectedDiversity in the<br />

author's laborato-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_536.html [4/9/2004 12:50:32 AM]


Document<br />

Page 537<br />

ries, holds the promise of a major reduction in the time required to produce new preclinical compounds<br />

with enhanced development potential.<br />

VI. Conclusion<br />

The preceding outlines a new drug-discovery paradigm that integrates structure-<strong>based</strong> design, directed<br />

strategies for combinatorial chemical synthesis, and a comprehensive chemi-informatics system for<br />

accumulating and analyzing information regarding chemical properties. Three-dimensional structures<br />

provide the information required to most efficiently direct the design and optimization of new lead<br />

compounds. High-throughput automated methods of chemical synthesis produce new classes of lead<br />

compounds and provide for the rapid generation of structure—activity data. Chemical informatics<br />

systems track chemical compounds, store chemical property data, develop predictive SAR models, and<br />

provide a means for intelligently directing the drug-discovery process. By applying this approach not<br />

only to the therapeutic target, but also to molecules involved in absorption, clearance, metabolism, or<br />

toxicology it will be possible to develop predictive models for bioavailability and toxicology. Ultimately<br />

this approach will greatly increase the cost effectiveness and efficiency of drug discovery <strong>by</strong> reducing<br />

the aggregate failure rate for development candidates.<br />

References<br />

1. Appelt K, Bacquet RJ, Bartlett CA, Booth CL, Freer ST, Fuhry MAM, Gehring MR, Herrmann SM,<br />

Howland EF, Janson CA, Jones TR, Ka C-C, Kathardekar V, Lewis KK, Marzoni GP, Matthews DA,<br />

Mohr C, Moomaw EW, Morse CA, Oatley SJ, Ogden RC, Reddy MR, Reich SH, Schoettlin WS, Smith<br />

WW, Varney MD, Villafranca JE, Ward RW, Webber S, Welsh KM, White J. <strong>Design</strong> of enzyme<br />

inhibitors using iterative protein crystallographic analysis. J Med Chem 1991; 34:1925–1934.<br />

2. Ealick SE, Babu YS, Bugg CE, Erion MD, Guida WC, Montgomery JA, Secrist III JA. Application of<br />

crystallographic and modeling methods in the design of purine nucleoside phosphorylase inhibitors.<br />

Proc Natl Acad Sci USA 1991; 88:11540–11544.<br />

3. Varney MD, Marzoni GP, Palmer CL, Deal JG, Webber S, Welsh KM, Bacquet RJ, Bartlett CA,<br />

Morse CA, Booth CLJ, Herrmann SM, Howland EF, Ward RW, White J. Crystal-structure-<strong>based</strong> design<br />

of Benz[cd]indole-containing inhibitors of thymidylate synthase. J Med Chem 1992; 35:663–676.<br />

4. Reich SH, Fuhry MAM, Nguyen D, Pino MJ, Welsh KM, Webber S, Janson CA, Jordan SR,<br />

Matthews DA, Smith WA, Bartlett CA, Booth CLJ, Herrmann SM, Howland EF, Morse CA, Ward RW,<br />

White J. <strong>Design</strong> and synthesis of novel 6,7-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_537.html [4/9/2004 12:50:39 AM]


Document<br />

Imidazotetrahydroquinoline inhibitors of thymidylate synthase using iterative protein crystal<br />

structure analysis. J Med Chem 1992; 35:847–858.<br />

5. Weber PC, Wendoloski JJ, Pantoliano MW, Salemme FR. Crystallographic and thermodynamic<br />

comparison of natural and synthetic ligands bound to streptavidin. J Amer Chem Soc 1992;<br />

114:3197–3200.<br />

6. Weber PC, Pantoliano MW, Simons DM, Salemme FR. <strong>Structure</strong>-<strong>based</strong> design of synthetic<br />

azobenzene ligands for streptavidin. J Amer Chem Soc 1994; 116:2717–2724.<br />

7. Andrews PR, Craik DJ, Martin JL. Functional group contributions to drug-receptor interactions. J<br />

Med Chem 1984; 27:1648–1657.<br />

Page 538<br />

8. Terrett NK, Gardner M, Gordon DW, Ko<strong>by</strong>lecki RJ, Steele J. Combinatorial synthesis—the design of<br />

compound libraries and their application to drug discovery. Tetrahedron 1995; 51:8135–8173.<br />

9. Baum RM. Combinatorial approaches provide fresh leads for medicinal chemistry. C and EN 1994;<br />

7:20–26.<br />

10. Gallop MA, Barrett RW, Dower WJ, Fodor SPA, Gordon EM. Applications of combinatorial<br />

technologies to drug discovery. 1. Background and peptide combinatorial libraries. J Med Chem 1994;<br />

37:1234–1251.<br />

11. Gordon EM, Barrett RW, Dower WJ, Fodor SPA, Gallop MA. Applications of combinatorial<br />

technologies to drug discovery. 2. Combinatorial organic synthesis, library screening strategies and<br />

future directions. J Med Chem 1994; 37:1385–1399.<br />

12. Dooley CT, Chung NN, Wolkes BC, Schiller PW, Bidlack JM, Pasternak GW, Houghten RA. An all<br />

D-amino acid peptide with central analgesic activity from a combinatorial library. Science 1994;<br />

266:2019–2022.<br />

13. Kerr JM, Banville SC, Zuckermann RN. Encoded combinatorial peptide libraries containing nonnatural<br />

amino acids. J Amer Chem Soc 1993; 115:2529–2531.<br />

14. Ohlmeyer MHJ, Swanson RN, Dillard LW, Reader JC, Asouline G, Kobayashi R, Wigler M, Still C.<br />

Complex synthetic chemical libraries indexed with molecular tags. Proc Natl Acad Sci USA 1993;<br />

90:10922–10926.<br />

15. Dankwardt SM, Newman SR, Krestenansky JL. Solid phase synthesis of aryl and benzylpiperidines<br />

and their application in combinatorial chemistry. Tetrahedron Letters 1995; 36:4923–4926.<br />

16. Baldwin JJ, Burbaum JJ, Henderson I, Ohlmeyer MHJ. Synthesis of a small molecule combinatorial<br />

library encoded with molecular tags. J Amer Chem Soc 1995; 117:5588–5589.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_538.html (1 of 2) [4/9/2004 12:50:42 AM]


Document<br />

17. Zuckermann RN, Martin EJ, Spellmeyer DC, Stauber GB, Shoemaker KR, Kerr JM, Figliozzi GM,<br />

Goff DA, Siani MA, Simon RJ, Banville SC, Brown EG, Wang L, Richter LS, Moos WH. Discovery of<br />

nanomolar ligands for 7-transmembrane G-protein coupled receptors from a diverse N-<br />

(substituted)glycine peptoid library. J Med Chem 1994; 37:2678–2685.<br />

18. Burbaum JJ, Ohlmeyer MHJ, Reader JC, Henderson I, Dillard LW, Li G, Randle TL, Sigal NH,<br />

Chelsky D, Baldwin JJ. A paradigm for drug discovery employing encoded combinatorial libraries. Proc<br />

Natl Acad Sci USA 1995; 92:6027–6031.<br />

19. Lam KS, Salmon SE, Hersh EM, Hru<strong>by</strong> VJ, Kazmierski WM, Knapp RJ. A new type of synthetic<br />

peptide library for identifying ligand-binding activity. Nature 1991; 354:82–84.<br />

20. Loew GH, Villar HO, Alkorta I. Strategies for indirect computer-aided drug design. Pharmaceutical<br />

Res 1993; 10:475–486.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_538.html (2 of 2) [4/9/2004 12:50:42 AM]


Document<br />

21. Kick EK, Ellman JA. Expedient method for the solid-phase synthesis of aspartic acid protease<br />

inhibitors directed towards the generation of libraries. J Med Chem 1995; 38:1427–1430.<br />

Page 539<br />

22. Campbell DA, Bermak JC, Burkoth TS, Patel DV. A transition state analog inhibitor combinatorial<br />

library. J Amer Chem Soc 1995; 117:5381–5382.<br />

23. Kuntz ID, Meng EC, Shoichet BK. <strong>Structure</strong>-<strong>based</strong> molecular design. Acc Chem Res 1994;<br />

27:117–123.<br />

24. Bohm H-J. LUDI: Rule-<strong>based</strong> automatic design<br />

of new substituents for enzyme inhibitor leads.<br />

Journal of Computer-Aided Molecular <strong>Design</strong> 1992;<br />

6:593–606.<br />

25. Bohm H-J. The computer program LUDI: A new method for the de novo design of enzyme<br />

inhibitors. Journal of Computer-Aided Molecular <strong>Design</strong> 1992; 6:61–78.<br />

26. Martin EJ, Blaney JM, Siani M, Spellmeyer DC, Wong AK, Moos WH. Measuring diversity:<br />

experimental design of combinatorial libraries for drug discovery. J Med Chem 1995; 38:1431–1436.<br />

27. Sullivan M. Mass Screening: a new approach to chemical discovery. Today's Chemist at Work<br />

1994;September:19–26.<br />

28. Agrafiotis DK, Bone RF, Salemme FR, Soll RM. A system and method of automatically generating<br />

chemical compounds with desired properties. US Patent 5463564, October 31, 1995.<br />

29. Hildalgo IJ, Raub TJ, Borchardt RT. Characterization of the human colon carcinoma cell line (Caco-<br />

2) as a model system for intestinal epithelial permeability. Gastroenterology 1989; 96:736–749.<br />

30. Gan L-S, Eads C, Niederer T, Bridgers A, Yanni S, Hsyu P-H, Pritchard FJ, Thakker D. Use of Caco-<br />

2 cells as an in vitro intestinal absorption and metabolism model. <strong>Drug</strong> Development and Industrial<br />

Pharmacy 1994; 20:615–631.<br />

31. Conradi RA, Hilgers AR, Ho NFH, Burton PS. The influence of peptide structure on transport across<br />

Caco-2 cells. Pharmaceutical Research 1991; 8:1453–1460.<br />

32. Pidgeon C, Ong S, Liu H, Qiu X, Pidgeon M, Dantzig AH, Munroe J, Hornback WJ, Kasher JS,<br />

Glunx L, Szczerba T. IAM chromatography: an in vitro screen for predicting drug membrane<br />

permeability. J Med Chem 1995; 38:590–594.<br />

33. Grundermann D, Gorboulev V, Gambaryan S, Veyhl M, Koepsell H. <strong>Drug</strong> excretion mediated <strong>by</strong> a<br />

new prototype of polyspecific transporter. Nature 1994; 372:549–552.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_539.html (1 of 2) [4/9/2004 12:50:46 AM]


Document<br />

34. Fei Y-J, Kanai Y, Nussberger S, Ganapathy V, Leibach FH, Romero MF, Singh SK, Boron WF,<br />

Hediger MA. Expression and cloning of a mammalian proton-coupled oligopeptide transporter. Nature<br />

1994; 368:563–566.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_539.html (2 of 2) [4/9/2004 12:50:46 AM]<br />

35. Agrafiotis<br />

DK, Bone R,<br />

Jaeger EP,<br />

Rhind AW,<br />

Salemme FR,<br />

Soll RM.<br />

Directed<br />

Diversity®: a<br />

new paradigm<br />

for drug<br />

discovery.<br />

1996; in<br />

preparation.


http://legacy.netlibrary.com/reader/message.asp?message=811&BookID=12640&FileName=Page_540.html<br />

The requested page could not be found.<br />

Return to previous page<br />

http://legacy.netlibrary.com/reader/message.asp?message=811&BookID=12640&FileName=Page_540.html [4/9/2004 12:50:57 AM]


Document<br />

21<br />

<strong>Structure</strong>-Based Combinatorial Ligand <strong>Design</strong><br />

Amedeo Caflisch<br />

University of Zürich, Zürich, Switzerland<br />

Claus Ehrhardt<br />

Novartis Pharma Inc., *Basel, Switzerland<br />

I. Introduction<br />

Page 541<br />

<strong>Structure</strong>-<strong>based</strong> ligand design is fascinating and challenging. Whenever it is possible to determine the<br />

three-dimensional structure of a pharmacologically relevant enzyme or receptor, computational<br />

approaches can be used to design novel high-affinity ligands. These methods can complement the broad<br />

screening efforts, which represent traditional lead discovery.<br />

In this chapter we focus on our approach to computer-aided ligand design. It is <strong>based</strong> on the docking of a<br />

diverse set of molecular fragments into the active site of a macromolecular target and on the use of a<br />

combinatorial strategy to connect them to form candidate ligands. The methodology is illustrated <strong>by</strong> an<br />

application to human thrombin, a trypsin-like serine protease fulfilling a central role in both hemostasis<br />

and thrombosis. The selective inhibition of thrombin is expected to prevent thrombotic diseases.<br />

Ligand-design programs are being developed at an ever increasing rate and some are related to various<br />

aspects of our ligand design approach. The LEGO software tool is <strong>based</strong> on the combination of multiple<br />

fragment docking, automatic connection <strong>by</strong> small linker units (one to four atom chains), and searching<br />

of three-dimensional databases for complementary molecules [1,2]. It has been implemented within the<br />

MOLOC molecular modeling system [3], which allows the visualization of the functionality maps and<br />

interactive model building of the growing ligands. Another related approach is that embodied in the<br />

program LUDI [4–8]. It makes use of statistical data from small-molecule<br />

* Formerly Sandoz Pharma Ltd.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_541.html [4/9/2004 12:51:01 AM]


Document<br />

Page 542<br />

crystal structures to determine binding sites of molecular fragments, i.e., discrete positions on the<br />

binding site surface suitable to form hydrogen bonds and/or to fill hydrophobic sites of the receptor.<br />

Alternatively, it uses simple rules or the output of the program GRID [9–12] to generate the interaction<br />

sites. Finally, the fragments fitted in the interaction sites are connected <strong>by</strong> linker groups. Other fragment<strong>based</strong><br />

programs are GROUPBUILD [13]; GROW [14], HOOK [15], NEWLEAD [16], SPROUT [17],<br />

and TORSION [18]. These and other strategies for computer-aided structure-<strong>based</strong> ligand design have<br />

been reviewed <strong>by</strong> several contributors [13,19,20].<br />

II. Docking Molecular Fragments<br />

A. Multiple Copy Simultaneous Search<br />

The present approach for ligand design is <strong>based</strong> on the combinatorial selection of molecular fragments<br />

optimally docked on the protein binding site to form a population of diverse candidate ligands. The<br />

multiple copy simultaneous search (MCSS) procedure combines the advantages of random distribution<br />

and simultaneous minimization of a set of replicas of a chemical fragment to obtain maps of<br />

energetically favorable positions and orientations (local energy minima) [21,22]. These maps, which<br />

contain all possible low-energy minima of a fragment-protein complex, are called functionality maps. A<br />

plethora of structural and thermodynamic data on inhibitor-enzyme complexes [23–26] suggest that the<br />

burial of nonpolar groups of the ligand in hydrophobic pockets of the protein is important for binding<br />

affinity and that intermolecular electrostatic interactions determine selectivity. For this reason, and<br />

because most of the known enzymes' binding sites have both hydrophilic and hydrophobic character,<br />

very diverse functional groups are used in MCSS. Representative examples include charged (e.g.,<br />

acetate, benzamidine, methylammonium, methylguanidinium, pyrrolidine); polar (e.g., methanol, 2propanone,<br />

N-methylacetamide); aromatic (e.g., benzene, pyrrole, imidazole, phenol); and aliphatic<br />

(e.g., propane, isobutane, cyclopentane, cyclohexane) groups. Although most of these fragments are<br />

rigid, MCSS can also generate the functionality maps of flexible medium-size fragments, e.g., the amino<br />

acid side chains. Additional functional groups and more complex heterocyclic systems are currently<br />

being introduced to increase the diversity of the resulting ligands and to better characterize the<br />

specificity of the binding pockets (A. Caflisch and C. Ehrhardt, unpublished results).<br />

As shown <strong>by</strong> a flowchart in Figure 1, the method is fully automated, although certain critical parameters<br />

(e.g., number of replicas, radius of the sphere for random distribution, CHARMM parameters for the<br />

minimization) can be adjusted <strong>by</strong> the user to optimize it for specific applications. Several thousand<br />

replicas of a given group are randomly distributed inside a sphere<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_542.html [4/9/2004 12:51:09 AM]


Document<br />

Figure 1<br />

Schematic representation of the MCSS procedure. Conditional<br />

statements are enclosed <strong>by</strong> diamonds.<br />

Page 543<br />

whose radius is chosen large enough to cover the entire region of interest. This can be a known binding<br />

site or the entire protein, if one wants to explore alternative binding pockets. The initial random<br />

distribution also can be performed inside a parallelepiped if the region of interest is elongated in one or<br />

two directions. A minimal distance can be given as input to avoid bad contacts between functional group<br />

atoms and protein atoms for the initial distribution.<br />

Subsets of between 500 and 3000 randomly distributed replicas of the same group are simultaneously<br />

minimized in the force field of the protein. The<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_543.html (1 of 2) [4/9/2004 12:51:17 AM]


Document<br />

Page 544<br />

interactions between the group replicas are omitted. The polar-hydrogen approximation (PARAM19) of<br />

the CHARMM force field is used [27]. In the application of the MCSS method to the sialic acid binding<br />

site of the influenza coat protein hemagglutinin [21], HIV-1 aspartic proteinase [22], and thrombin<br />

[20,28] the protein was kept fixed; hence, the forces on each replica consist of its internal forces and<br />

those due to the protein, which has unique conformation and, therefore, generates a unique field. The<br />

positions are compared every 1000 steps to eliminate replicas converging toward a common minimum.<br />

Further details concerning the methodology are given in References 21 and 22, while a critical<br />

assessment has been presented in Reference 20.<br />

B. Simple Approximations of Solvation Effects<br />

In previous applications of MCSS [21,22] the effects of the solvent were neglected, i.e., all proteinfragment<br />

interactions were calculated with a vacuum potential [27]. This choice was <strong>based</strong> on the<br />

principle that fast methods are necessary to perform effective searches of the binding site and that good<br />

candidate ligands subsequently can be ranked in terms of their binding free energy [20,28]. A possible<br />

difficulty with this approach is that minimized positions may be missed or misplaced due to the lack of a<br />

solvation correction during the MCSS minimization.<br />

Electrostatic Shielding<br />

In MCSS studies of thrombin [20,28], it was observed that minima of charged groups tend to cluster in<br />

the vicinity of charged side chains on the thrombin surface and in the S1 (basic groups) or S1' pocket<br />

(acidic groups). It is then necessary to estimate the electrostatic desolvation of both protein and fragment<br />

to obtain a realistic ranking of the minima [28]. As a simple test of the importance of electrostatic<br />

shielding, a distance-dependent dielectric function [29] was introduced instead of the unit dielectric<br />

constant in the vacuum potential. The overall shape of the acetate map did not change, but three more<br />

minima were found close to the Lys60F side chain in S1' [20]. It is difficult to find a physical meaning<br />

in favor of the distance-dependent dielectric function. Nevertheless, it is a simple and useful<br />

approximation, since it yields a smoother and more realistic potential surface than the vacuum<br />

Coulombic interaction.<br />

Hydrophobic Effect<br />

Aliphatic and aromatic fragments do not bear any partial charges in the polar-hydrogen approximation.<br />

Hence, MCSS determines their optimal position in the<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_544.html [4/9/2004 12:51:24 AM]


Document<br />

Page 545<br />

protein binding site exclusively <strong>by</strong> van der Waals interactions. Minima of nonpolar fragments may be<br />

found in hydrophilic pockets because of the lack of an energy penalty for protein desolvation. A<br />

representative example of a cluster of propane minima in a mainly hydrophilic region of the HIV-1<br />

aspartic proteinase binding site is shown in Reference 20. In a simple attempt to approximate<br />

desolvation of polar regions of the protein, the attractive contribution of the van der Waals interaction<br />

energy was switched off between atoms of nonpolar MCSS fragments and all protein polar hydrogen,<br />

nitrogen, oxygen, and sp 2 carbon atoms. In addition, the van der Waals radius of nitrogen and oxygen<br />

atoms was increased from the PARAM19 default value of 1.6 Å to 2.2 Å and the van der Waals radius<br />

of the aliphatic carbons was reduced <strong>by</strong> 0.1 Å to avoid the too large van der Waals distance between<br />

carbons often produced <strong>by</strong> PARAM19. The modified force field yields thrombin functionality maps of<br />

propane, cyclopentane, cyclohexane, and benzene in agreement with structural data of known inhibitors<br />

(see Section II. C). In addition, these nonpolar groups are prevented from occupying hydrophilic<br />

pockets.<br />

As a further test of the modified force field, the MCSS procedure was used to generate the propane<br />

functionality map on the surface of the A peptide chain of the leucine zipper, i.e., residues 249–281 from<br />

the yeast transcriptional activator protein GCN4. Leucine zippers are composed of amphipathic α<br />

helices containing heptad repeats (abcdefg) in which hydrophobic residues are frequent at a and d. As<br />

the x-ray structure indicates [30,31], the two amphipathic α helices are held together <strong>by</strong> hydrophobic<br />

interactions between residues in the a and d positions (Figure 2). The B peptide chain was removed and<br />

MCSS was run separately with the original and the modified force field starting from the same random<br />

distribution of 5000 propane replicas around helix A. The sixty most favorable minima obtained with the<br />

modified force field are distributed in twelve clusters, seven of which match the Val and Leu side chains<br />

of the B helix involved in the interhelical interactions (Figure 2). Of the remaining five clusters, labeled<br />

A to E in Figure 2, B has minima in contact with Val and Ala side chains, D with Leu and Tyr side<br />

chains, while A, C, and E with both a Val or Leu side chain and the alkyl part of Lys side chain. The<br />

sixty most favorable minima obtained with the original force field form sixteen clusters (not shown).<br />

Only four of these match Val and Leu side chains of the B helix, while eight clusters desolvate one (or<br />

more) polar group on the hydrophilic (exposed) surface of the A helix.<br />

The functionality maps of polar groups generated with a distance-dependent dielectric and those of<br />

nonpolar groups obtained with the modified force field are closer to those obtained <strong>by</strong> using a<br />

continuum dielectric model [32–35] to postprocess the MCSS minima and estimate solvation effects<br />

[28]. Hence, simple modifications of the force field may result in more accurate projections of the<br />

binding free energy. Since these modifications are used during the<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_545.html [4/9/2004 12:51:27 AM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_546.html (1 of 2) [4/9/2004 12:51:45 AM]<br />

Page 546


Document<br />

Figure 2<br />

Stereo view of the sixty propane minima (thick lines) obtained with the modified force<br />

field (see text) on the surface of the A peptide chain (medium lines) of the GCN4 leucine<br />

zipper (PDB code 2ZTA). Although the B peptide chain was removed during the MCSS<br />

procedure, its backbone and hydrophobic side chains are also drawn (thin lines) to show<br />

how the propane minima match the aliphatic groups of chain B. Hydrophobic residues<br />

are labeled at their C α atom.Five clusters of propane minima that do not match the<br />

hydrophobic side chain of the B helix involved in the interhelical interactions are<br />

labeled from A (top right) to E (bottom center) and discussed in the text.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_546.html (2 of 2) [4/9/2004 12:51:45 AM]


Document<br />

minimization phase, a more realistic distribution of functional group minima is generated.<br />

C. Thrombin Functionality Maps<br />

Page 547<br />

Human thrombin is one of the best characterized enzymes from a structural point of view (Figure 3). It<br />

binds a series of diverse inhibitors without major rearrangements in its conformation, as shown <strong>by</strong> a<br />

number of x-ray crystallography studies [26,36–39]. Its S3 and S2 precleavage subpockets have<br />

hydrophobic character, whereas at the bottom of the S1 or recognition pocket the carboxy group of<br />

Asp189 is a salt bridge partner for basic side chains (Figure 3). D-phenylalanyl-L-prolyl-L-arginine<br />

chloromethane, PPACK, (Figure 4a), and Nα-((2-naphthylsulfonyl)glycyl)-DL-pamidinophenylalanylpiperidine,<br />

NAPAP, (Figure 4b) are the archetypal active-site inhibitors of<br />

thrombin. The crystal structure of the thrombin-NAPAP complex is shown in Figure 3. The PPACK and<br />

NAPAP inhibitors bind to the thrombin active site <strong>by</strong> occupying the S3 and S2 pockets with their<br />

hydrophobic moieties and <strong>by</strong> positioning their basic group (guanidinium of PPACK, benzamidine of<br />

NAPAP) into S1 to form a salt bridge with Asp189.<br />

In continuation of a project aiming at the structure-<strong>based</strong> design of low molecular weight, active-site<br />

directed inhibitors of human thrombin [40], MCSS was used to generate a series of functionality maps<br />

of the thrombin S3 to S2' pockets [20,28]. A detailed description of the thrombin functionality maps and<br />

the continuum approximation used to postprocess the MCSS minima is given in Reference [28]. From<br />

the analysis of the results for the nonpolar groups it is evident that hydrophobic moieties prefer to bind<br />

to the S3 and S2 pockets (Figure 5). The solvent exposed face of the Trp60D indole is another favorable<br />

site, though the intermolecular van der Waals interactions are much smaller. Binding to the S2' region is<br />

favored <strong>by</strong> interactions with the Leu40 side chain but implies a desolvation penalty because of the burial<br />

of part of the Arg73 guanidinium and/or the Gln151 side chain. The latter might be an artifact of the<br />

rigid protein structure used in the minimization, since the side chains of Arg73 and Gln151 are flexible<br />

enough to displace their polar groups towards a more exposed region. Binding to the neighboring Leu41<br />

side chain in S1' is highly unfavorable because of the concomitant desolvation of Lys60F.<br />

For polar groups with zero net charge there are several hydrophylic groups on the thrombin main chain<br />

that might be involved in strong hydrogen<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_547.html [4/9/2004 12:51:56 AM]


Document<br />

Figure 3<br />

(a) Stereo view of the thrombin molecule in its complex with NAPAP. The side chains of<br />

thrombin involved in the binding of NAPAP and the disulfide bridges are shown in<br />

stick-and-ball representation with black sticks. NAPAP is shown in stick-and-ball<br />

representation with white sticks. (b) Zoom image of the active site region. Figures made<br />

with the program MOLSCRIPT [42].<br />

Page 548<br />

bonds. These are 214CO, 216NH, 216CO, and 219NH in S1; 193NH and 195NH in the oxyanion hole;<br />

41CO in S1'; 40CO in S2'; and 147NH and 148NH on the autolysis loop, whose exposure is dependent<br />

on crystallization conditions and inhibitor type.<br />

Two main conclusions can be drawn from the analysis of the minimized positions of the charged<br />

functional groups. First, the minima with the lowest binding free energy have optimal hydrogen bonds<br />

with the Asp189 side chain in the S1 pocket. Representative examples are the lowest energy minimum<br />

of benzamidine (Figure 5) and the lowest energy minima of methylguanidinium and methylammonium<br />

(Figure 4 of Reference 28). Since the Asp189 side chain<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_548.html (1 of 2) [4/9/2004 12:52:27 AM]


Document<br />

Figure 4<br />

Chemical structure of PPACK (a) and NAPAP (b).<br />

Page 549<br />

is more buried than the side chain of Lys60F, the minima of positively charged groups interacting with<br />

the former have a more favorable binding free energy than those of the negatively charged groups close<br />

to the latter. This is due to reduced shielding of the charge—charge interaction and the smaller<br />

desolvation of the carboxylate oxygens of Asp189 compared to the amino group in Lys60F [28].<br />

Second, polar groups on the protein surface are not ideal partners for a charged functional group because<br />

the high desolvation penalty for these groups might not be completely compensated for <strong>by</strong> the favorable<br />

electrostatic interaction energy. This finding is analogous to the results for the polar functional groups,<br />

i.e., their binding to a partially exposed charged side chain of the protein may result in an unfavorable<br />

total binding free energy.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_549.html (1 of 2) [4/9/2004 12:52:31 AM]


Document<br />

Figure 5<br />

Stereo view of the three lowest energy minima of benzene obtained with the<br />

modified force field and the lowest energy minimum of benzamidine (thick lines for<br />

heavy atoms and thin lines for polar hydrogens) in the thrombin active site (thin lines).<br />

The inhibitor PPACK is also shown (medium lines), though it was removed during the<br />

MCSS procedure. Some C α atoms of thrombin are labeled.<br />

III. Connecting Molecular Fragments<br />

A. Computational Combinatorial Ligand <strong>Design</strong><br />

Overview<br />

Page 550<br />

The recently developed program for computational combinatorial ligand design (CCLD) requires as<br />

input atomic coordinates and partial charges of the protein atoms, as well as the coordinates of the<br />

MCSS minima and the individual contributions to the free energy of binding [28]. An additional file<br />

contains a number of control parameters and, for each functional group used for MCSS, a list of atoms<br />

which can be used for connection (linkage atoms). The following procedures are performed during a<br />

regular execution of CCLD (Figure 6): The MCSS minima are first sorted according to their<br />

approximated binding free energies [28]; then a list of bonding fragment pairs and a list of overlapping<br />

fragment pairs are generated (see below). This is followed <strong>by</strong> the combinatorial generation of putative<br />

ligands (see below).<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_550.html (1 of 2) [4/9/2004 12:53:02 AM]


Document<br />

Figure 6<br />

Schematic representation of the CCLD program. Variable assignments are symbolized<br />

<strong>by</strong> :=. Conditional statements are enclosed <strong>by</strong> diamonds. [From Reference 28.]<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_551.html (1 of 2) [4/9/2004 12:53:13 AM]<br />

Page 551


Document<br />

Lists of Bonding Fragment Pairs and Overlapping Fragment Pairs<br />

Page 552<br />

The user has to specify for each functional group type which atoms are to be used for connection to<br />

other fragments. For each linkage atom CCLD generates a set of possible linkage points, i.e., points that<br />

will be used to determine the position and orientation of the link. All possible pairs of minimized<br />

positions are then analyzed and added to the list of bonding fragment pairs if they can be linked;<br />

otherwise, if two fragments have bad contacts they are added to the list of overlapping fragment pairs. A<br />

pair of bonding fragments may be connected <strong>by</strong> a linker unit, <strong>by</strong> a single covalent bond (1-bond), or <strong>by</strong><br />

fusing two overlapping atoms belonging to different fragments (0-bond). The linker units are small since<br />

their function is to optimally connect two fragments without adding considerably to the molecular<br />

weight. The following linker elements have been implemented so far: Keto and methylene (2-bond),<br />

amide and ethylene (3-bond). The user is free to choose minimal and maximal values for the distance (d)<br />

between linkage atoms for each connection type. In the application to thrombin the following values in<br />

Ångstroms were used: d


Document<br />

B. Candidate Ligands of Thrombin<br />

Page 553<br />

To test if CCLD is able to reproduce the known thrombin inhibitors, the functional groups of PPACK<br />

and NAPAP were given as input molecular fragments in the orientation derived from the crystal<br />

structures of the complexes. The CCLD program generated a set of candidate ligands that not only<br />

contained the PPACK and NAPAP structures but also a number of interesting hybrid molecules<br />

consisting of fragments from both inhibitors. A representative example is shown in Figure 7. This<br />

putative ligand consists of the C-terminal part of NAPAP, whose piperidine ring is connected at the 3position<br />

to the PPACK D-Phe <strong>by</strong> an amide linker. The latter has its carbonyl oxygen involved in a<br />

hydrogen bond with the Gly216 NH.<br />

In another run, the MCSS minima of benzamidine (Figure 5), benzene (Figure 5), cyclopentane, and<br />

cyclohexane [28] were used as starting molecular fragments. In a few seconds of CPU time of an SGI<br />

Indigo2 (R4400 processor), CCLD generated a series of molecules showing the same interaction<br />

patterns as those of known thrombin inhibitors, i.e., hydrophobic moieties in S3 and S2, hydrogen bonds<br />

with the polar groups of Gly216, and benzamidine in S1. One of these putative ligands is shown in<br />

Figures 8 and 9. It is involved in the same interactions as in the NAPAP-thrombin complex except for<br />

the hydrogen bond with the CO of Gly216. Its cyclohexane ring in S3 is connected to the<br />

Figure 7<br />

PPACK-NAPAP hybrid ligand (thick lines) generated <strong>by</strong> CCLD starting<br />

from the functional groups of PPACK and NAPAP (thin lines). The amide linker<br />

connecting the piperidine in 3-position to the D-Phe was created <strong>by</strong> CCLD.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_553.html (1 of 2) [4/9/2004 12:53:22 AM]


Document<br />

Figure 8<br />

Minimized structure of a putative ligand suggested <strong>by</strong> CCLD (thick lines).<br />

The CCLD run used MCSS minima of benzamidine, benzene, cyclopentane, and<br />

cyclohexane. The putative ligand was minimized in the thrombin active site, whose residues<br />

within 8 Å of any atom of the ligand were allowed to move. The remaining residues of<br />

thrombin were kept rigid. The NAPAP structure is also shown (thin lines) as a basis of<br />

comparison.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_554.html (1 of 2) [4/9/2004 12:53:32 AM]<br />

Page 554


Document<br />

Figure 9<br />

Stereo view of the minimized complex between the putative ligand shown in Figure 8<br />

and thrombin (thin lines). Intermolecular hydrogen bonds are drawn <strong>by</strong> dashed lines.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_554.html (2 of 2) [4/9/2004 12:53:32 AM]


Document<br />

Page 555<br />

N-acylpyrrolidine ring in S2 <strong>by</strong> a single methylene linker. This is a novel design and the candidate<br />

ligand appears to be more rigid than NAPAP, since it has a smaller number of rotatable bonds. Hence,<br />

the penalty paid for the loss in entropy upon binding should be smaller for this CCLD hit than for<br />

NAPAP.<br />

In a study with a more diverse set of starting molecular fragments, ligands containing both a “core”<br />

similar to known inhibitors and additional intermolecular hydrogen bonds and/or van der Waals<br />

interactions were generated [28].<br />

IV. Conclusion<br />

A combinatorial approach for the computer-aided design of putative ligands of proteins or receptors of<br />

known three-dimensional structure has been presented. Diversity of these candidate ligands is provided<br />

<strong>by</strong> first docking a set of diverse molecular fragments. For aliphatic functional groups a modified force<br />

field (switching off the attractive part of the van der Waals interaction with polar atoms of the protein)<br />

was introduced to better approximate solvation effects, there<strong>by</strong> avoiding the docking of apolar fragments<br />

into hydrophilic cavities of the macromolecular target. The second part of the present ligand design<br />

approach consists of a combinatorial strategy for the connection of optimally docked molecular<br />

fragments <strong>by</strong> small linkage elements having optimal interactions with the target molecule. An<br />

application was presented in which candidate inhibitors of thrombin were found.<br />

It is important to note that the most difficult part of ligand design is the prediction of binding affinity. A<br />

method to estimate relative binding constants has recently been applied to a series of similar inhibitors<br />

of HIV-1 aspartic proteinase [41]. Our own efforts are concentrated on the development of an approach<br />

that will predict relative binding affinities and will be general enough to be used for any enzyme or<br />

receptor of known structure (A. Caflisch, D. Arosio, and C. Ehrhardt, work in progress).<br />

One of the purposes of this chapter was to show that structure-<strong>based</strong> ligand design is a fascinating and<br />

progressing research field. It is fascinating not only for its ultimate goal, i.e., the discovery of ethical<br />

drugs, but also because it is <strong>based</strong> on, and there<strong>by</strong> increases, our understanding of molecular interactions<br />

and recognition phenomena on an atomic level. Another reason is its multidisciplinary character, which<br />

requires skills in different branches of science, from theoretical physics and chemistry to computer<br />

science and statistics. That structure-<strong>based</strong> computer-aided ligand design is a progressing field is evident<br />

in many chapters of this book. This is mainly a consequence of the methodological developments and<br />

the ever-increasing performance of computers.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_555.html [4/9/2004 12:53:33 AM]


Document<br />

Acknowledgments<br />

Page 556<br />

We thank J. Apostolakis and Professor A. Plückthun for helpful discussions. The calculations were<br />

performed on an SGI Indigo2 and an eight-processor SGI Challenge (R4400 processors). The<br />

CHARMM program within the version 4.0 of the QUANTA software package (Biosym-MSI Inc) was<br />

used for some of the minimization performed in this work. The CCLD program is available from A.<br />

Caflisch.<br />

This work was supported <strong>by</strong> the Swiss National Science Foundation (Schweizerischer Nationalfonds<br />

grant nr. 3100-043423.95) and <strong>by</strong> Novartis Pharma Inc.<br />

References<br />

1. Gubernator K, Broger C, Bur D, Doran DM, Gerber PR, Müller K, Schaumann TM. <strong>Structure</strong>-<strong>based</strong><br />

ligand design. In: Hermann EC, Frankle R, eds. Computer-Aided <strong>Drug</strong> <strong>Design</strong> in Industrial Research<br />

1995:61–77.<br />

2. Müller K. Paradigms of rational molecular design. In: Schwartz TW, Hjorth SA, Kastrup JS, eds.<br />

<strong>Structure</strong> and Function of 7TM Receptors, in press.<br />

3. Gerber PR, Gubernator K, Müller K. Generic shapes for the conformational analysis of macrocyclic<br />

structures. Helv Chim Acta 1988; 71:1429–1441.<br />

4. Böhm HJ. The computer program LUDI: a new method for de novo design of enzyme inhibitors. J<br />

Comput-Aided Mol <strong>Design</strong> 1992; 6:61–78.<br />

5. Böhm HJ. LUDI: rule-<strong>based</strong> automatic design of new substituents for enzyme inhibitor leads. J<br />

Comput-Aided Mol <strong>Design</strong> 1992; 6:593–606.<br />

6. Böhm HJ. The development of a simple empirical scoring function to estimate the binding constant<br />

for a protein-ligand complex of known three-dimensional structure. J Comput-Aided Mol <strong>Design</strong> 1994;<br />

8:243–256.<br />

7. Böhm HJ. On the use of LUDI to search the fine chemicals directory for ligands of proteins of known<br />

three-dimensional structure. J Comput-Aided Mol <strong>Design</strong> 1994; 8:623–632.<br />

8. Böhm HJ. Site-directed structure generation <strong>by</strong> fragment-joining. Perspectives in <strong>Drug</strong> Discovery and<br />

<strong>Design</strong> 1995; 3:21–33.<br />

9. Goodford PJ. A computational procedure for determining energetically favorable binding sites on<br />

biologically important macromolecules. J Med Chem 1985; 28:849–857.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_556.html (1 of 3) [4/9/2004 12:53:38 AM]


Document<br />

10. Bob<strong>by</strong>er DNA, Goodford PJ, McWhinnie PM, Wade RC. New hydrogen-bond potentials for use in<br />

determining energetically favorable binding sites on molecules of known structure. J Med Chem 1989;<br />

32:1083–1094.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_556.html (2 of 3) [4/9/2004 12:53:38 AM]<br />

11. Wade<br />

RC, Clark<br />

KJ,<br />

Goodford<br />

PJ. Further<br />

development<br />

of hydrogen<br />

bond<br />

functions for<br />

use in<br />

determining<br />

energetically<br />

favorable<br />

binding sites<br />

on<br />

molecules of<br />

known<br />

structure. 1.<br />

Ligand<br />

probe<br />

groups with<br />

the ability to<br />

form two<br />

hydrogen<br />

bonds. J<br />

Med Chem<br />

1993;<br />

36:140–147.


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_556.html (3 of 3) [4/9/2004 12:53:38 AM]<br />

12. Wade<br />

RC,<br />

Goodford<br />

PJ. Further<br />

development<br />

of hydrogen<br />

bond<br />

functions for<br />

use in<br />

determining<br />

energetically<br />

favorable<br />

binding sites<br />

on<br />

molecules of<br />

known


Document<br />

structure. 2. Ligand probe groups with the ability to form more than two hydrogen bonds. J Med<br />

Chem 1993; 36:148–156.<br />

Page 557<br />

13. Rotstein SH, Murcko MA. GroupBuild: A fragment-<strong>based</strong> method for de novo drug design. J Med<br />

Chem 1993; 36:1700–1710.<br />

14. Moon JB, Howe WJ. Computer design of bioactive molecules: a method for receptor-<strong>based</strong> de novo<br />

ligand design. Proteins: <strong>Structure</strong>, Function and Genetics 1991; 11:314–328.<br />

15. Eisen MB, Wiley DC, Karplus M, Hubbard RE. HOOK: A program for finding novel molecular<br />

architectures that satisfy the chemical and steric requirements of a macromolecule binding site. Proteins:<br />

<strong>Structure</strong>, Function and Genetics 1994; 19:199–221.<br />

16. Tschinke V, Cohen NC. The NEWLEAD program: a new method for the design of candidate<br />

structures from pharmacophoric hypotheses. J Med Chem 1993; 14:3863–3870.<br />

17. Gillet VJ, Myatt G, Zsoldos Z, Johnson AP. SPROUT, HIPPO and CAESA: Tools for de novo<br />

structure generation and estimation of synthetic accessibility. Perspectives in <strong>Drug</strong> Discovery and<br />

<strong>Design</strong> 1995; 3:34–50.<br />

18. Lewis RA, Roe DC, Huang C, Ferrin TE, Langridge R, Kuntz ID. Automated site-directed drug<br />

design using molecular lattices. J Mol Graphics 1992; 10:66–78.<br />

19. Kuntz ID. <strong>Structure</strong>-<strong>based</strong> strategies for drug design and discovery. Science 1992; 257:1078–1082.<br />

20. Caflisch A, Karplus M. Computational combinatorial chemistry for de novo ligand design: Review<br />

and assessment. Perspectives in <strong>Drug</strong> Discovery and <strong>Design</strong> 1995; 3:51–84.<br />

21. Miranker A, Karplus M. Functionality maps of binding sites: a multiple copy simultaneous search<br />

method. Proteins: <strong>Structure</strong>, Function, and Genetics 1991; 11:29–34.<br />

22. Caflisch A, Miranker A, Karplus M. Multiple copy simultaneous search and construction of ligands<br />

in binding sites: application to inhibitors of HIV-1 aspartic proteinase. J Med Chem 1993;<br />

36:2142–2167.<br />

23. Appelt K. Crystal structures of HIV-1 protease-inhibitor complexes. Perspectives in <strong>Drug</strong> Discovery<br />

and <strong>Design</strong> 1993; 1:23–48.<br />

24. Wlodawer A, Erickson JW. <strong>Structure</strong>-<strong>based</strong> inhibitors of HIV-1 protease. Annu Rev Biochem 1993;<br />

62:543–585.<br />

25. Stubbs MT, Bode W. Crystal structures of thrombin and thrombin complexes as a framework for<br />

antithrombotic drug design. Perspectives in <strong>Drug</strong> Discovery and <strong>Design</strong> 1993; 1:431–452.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_557.html (1 of 2) [4/9/2004 12:53:41 AM]


Document<br />

26. Hilpert K, Ackermann J, Banner DW, Gast A, Gubernator K, Hadvary P, Labler L, Müller K,<br />

Schmid G, Tschopp T, van de Waterbeemd H. <strong>Design</strong> and synthesis of potent and highly selective<br />

thrombin inhibitors. J Med Chem 1994; 37:3889–3901.<br />

27. Brooks BR, Bruccoleri RE, Olafson BD, States DJ, Swaminathan S, Karplus M. CHARMM: A<br />

program for macromolecular energy, minimization, and dynamics calculations. J Comput Chem 1983;<br />

4:187–217.<br />

28. Caflisch A. Computational combinatorial ligand design: Application to human α-thrombin. J<br />

Computer-Aided Molec <strong>Design</strong> 1996; 10:372–396.<br />

29. Gelin BR, Karplus M. Sidechain torsional potentials and motion of amino acids in proteins: bovine<br />

pancreatic trypsin inhibitor. Proc Natl Acad Sci USA 1975; 72:2002–2006.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_557.html (2 of 2) [4/9/2004 12:53:41 AM]


Document<br />

30. O'Shea EK, Rutkowski R, Kim PS. Evidence that the leucine zipper is a coiled coil. Science 1989;<br />

243:538–542.<br />

31. O'Shea EK, Klemm JD, Kim PS, Alber T. X-ray structure of the GCN4 leucine zipper, a twostranded,<br />

parallel coiled coil. Science 1991; 254:539–544.<br />

Page 558<br />

32. Warwicker J, Watson HC. Calculation of the electric potential in the active site cleft due to α-helix<br />

dipoles. J Mol Biol 1982; 157:671–679.<br />

33. Gilson MK, Honig BH. Calculation of the total electrostatic energy of a macromolecular system:<br />

solvation energies, binding energies, and conformational analysis. Proteins: <strong>Structure</strong>, Function, and<br />

Genetics 1988; 4:7–18.<br />

34. Bashford D, Karplus M. pK as of ionizable groups in proteins: atomic detail from a continuum<br />

electrostatic model. Biochemistry 1990; 29:10219–10225.<br />

35. Davis ME, Madura JD, Luty BA, McCammon JA. Electrostatics and diffusion of molecules in<br />

solution: simulations with the University of Houston Brownian dynamics program. Comp Phys Comm<br />

1991; 62:187–197.<br />

36. Bode W, Mayr I, Baumann U, Huber R, Stone SR, Hofsteenge J. The refined 1.9-Å crystal structure<br />

of human α-thrombin: interaction with D-Phe-Pro-Arg chloromethylketone and significance of the Tyr-<br />

Pro-Pro-Trp insertion segment. EMBO J 1989; 8:3467–3475.<br />

37. Banner DW, Hadvary P. Crystallographic analysis at 3.0-Å resolution of the binding to human<br />

thrombin of four active site-directed inhibitors. J Biol Chem 1991; 266:20085–20093.<br />

38. Obst U, Gramlich V, Diederich F, Weber L, Banner DW. <strong>Design</strong> neuartiger, nichtpeptidischer<br />

Thrombin-Inhibitoren und Struktur eines Thrombin-Inhibitor-Komplexes. Angew Chem 1995;<br />

107:1874–1877.<br />

39. Håkansson K, Tulinsky A, Abelman MM, Miller TA, Vlasuk GP, Bergum PW, Lim-Wil<strong>by</strong> MSL,<br />

Brunck TK. Crystallographic structure of a peptidyl keto acid inhibitor and human α-thrombin.<br />

Bioorganic and Medicinal Chemistry 1995; 3:1009–1017.<br />

40. Tapparelli C, Metternich R, Ehrhardt C, Cook NS. Synthetic low-molecular weight thrombin<br />

inhibitors: molecular design and pharmacological profile. TIPS 1993; 14:366–376.<br />

41. Holloway MK, Wai JM, Halgren TA, Fitzgerald PMD, Vacca JP, Dorsey BD, Levi RB, Thompson<br />

WJ, Chen LJ, deSolms SJ, Gaffin N, Ghosh AK, Giuliani EA, Graham SL, Guare JP, Hungate RW, Lyle<br />

TA, Sanders WM, Tucker TJ, Wiggins M, Wiscount CM, Woltersdorf OW, Young SD, Darke PL,<br />

Zugay JA. A priori prediction of activity for HIV-1 protease inhibitors employing energy minimization<br />

in the active site. J Med Chem 1995; 38:305–317.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_558.html (1 of 2) [4/9/2004 12:53:43 AM]


Document<br />

42. Kraulis P, Molscript, a program to produce both detailed and schematic plots of protein structures. J<br />

Appl Crystallogr 1991; 24:946–950.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_558.html (2 of 2) [4/9/2004 12:53:43 AM]


Document<br />

22<br />

Peptidomimetic and Nonpeptide <strong>Drug</strong> Discovery: Impact of <strong>Structure</strong>-<br />

Based <strong>Drug</strong> <strong>Design</strong><br />

Tomi K. Sawyer<br />

Parke-Davis Pharmaceutical Research, Ann Arbor, Michigan<br />

I. Introduction<br />

Page 559<br />

Peptidomimetic and nonpeptide drug discovery has evolved to become an extraordinarily intriguing area<br />

of interdisciplinary research. It challenges synthetic, computational, and biophysical chemists,<br />

biochemists, pharmacologists, and drug delivery scientists to collaboratively discover lead compounds<br />

that exhibit sufficient potency, selectivity, metabolic stability, and in vivo pharmacological efficacy to<br />

warrant further development as drug candidates. Over the past two decades a highly focused effort in<br />

both industry and academia has advanced the rational transformation of “first generation” peptide lead<br />

compounds to significantly modified analogs having minimal peptide-like chemical structure [1–18].<br />

Such work is typically illustrated <strong>by</strong> systematic backbone and/or side chain modifications,<br />

transformation into macrocycles, structure-conformation analysis (x-ray, NMR, CD), and computerassisted<br />

molecular modeling. These extensive structure-activity and structure-conformation studies have<br />

enabled the creation of prototype peptidomimetic “second-generation” analogs from initial peptide<br />

leads. Over recent years the emergence of 3D structural information on target proteins has significantly<br />

impacted peptidomimetic drug discovery strategies. Particularly noteworthy has been the advancement<br />

of the de novo design of chemically novel compounds that possess very limited peptide-like<br />

substructure, but include structure-<strong>based</strong> functional group modifications which may make new<br />

intermolecular interactions with the target protein (versus a “first generation” peptide lead compound).<br />

Furthermore, the integration of such structure-<strong>based</strong> drug design efforts with high-throughput<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_559.html [4/9/2004 12:53:45 AM]


Document<br />

Page 560<br />

screening and synthetic chemical library generation is markedly reshaping peptide, peptidomimetic, and<br />

nonpeptide drug discovery research. In this chapter a few examples of peptidomimetic and nonpeptide<br />

drug discovery are detailed to highlight the scope of such work relative to a few specific targets (e.g.,<br />

receptors, proteases, and signal transduction proteins) in which structure-<strong>based</strong> drug design has<br />

contributed in significant ways.<br />

A. Peptides: Molecular Diversity and φ-Ψ-χ Space<br />

Peptides exhibit extraordinary molecular diversity <strong>by</strong> virtue of their varying primary structures (Figure<br />

1). For many peptide hormones, neurotransmitters, and releasing factors the substructure of amino acids<br />

that contribute to molecular recognition (binding) and biological activity (signal transduction) at their<br />

target receptors have been determined [2]. Such work has led to the generation of pharmacophore<br />

models of either agonist or antagonist analogs and, in some cases, the design of peptidomimetics. Yet,<br />

for most peptide growth factors, cytokines, and large-sized (>50 amino acids) peptide hormones, the<br />

identification of the amino acid substructure which accounts for molecular recognition and signal<br />

transduction has been a difficult task, and proposals for pharmacophore models remain significant<br />

challenges. In this regard the term “pharmacophore” is defined as the collection of relevant groups<br />

(substructure) of a ligand which are arranged in three-dimensional space in a manner complementary to<br />

the target protein and are responsible for the biological property of the ligand as a result of binding of<br />

the ligand to its target protein [8b].<br />

The three-dimensional structural properties of peptides (Figure 2) are defined in terms of torsion angles<br />

(Ψ, φ, ω, χ) between the backbone amine nitrogen (Nα), backbone carbonyl carbon (C'), backbone<br />

methine carbon (Cα), and side chain hydrocarbon functionalization (eg., Cβ, Cγ, Cδ, Cε of Lys) derived<br />

from the amino acid sequence. A Ramachandran plot (Ψ versus φ) may define the preferred<br />

combinations of torsion angles for ordered secondary structures (conformations), such as α helix, β turn,<br />

γ turn, or β sheet. With respect to the amide bond torsion angle (ω) the trans geometry is more<br />

energetically favored for most typical dipeptide substructures, however, when the C-terminal partner is<br />

Pro or other N-alkylated (including cyclic) amino acids the cis geometry is possible and may further<br />

stabilize β-turn or γ-turn conformations. Molecular flexibility is directly related to covalent and/or<br />

noncovalent bonding interactions within a particular peptide. Even modest chemical modifications <strong>by</strong><br />

Nα-methyl, Cα-methyl or Cβ-methyl can have significant consequences on the resultant conformation<br />

[6; also, see Phe analogs in Figure 2].<br />

The Nα-Cα-C' scaffold may be further transformed <strong>by</strong> introduction of olefin substitution (e.g., Cα-<br />

Cβ rarrow.gif C=C or dehydroamino acid [19]) or insertion (e.g., Cα-C' rarrow.gif Cα-C=C-C' or<br />

vinylogous amino acid [20]). Also the Cβ carbon may be substituted to advance the design of so-called<br />

“chimeric” amino acids<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_560.html [4/9/2004 12:53:53 AM]


Document<br />

Figure 1<br />

Examples of native peptide hormones, neurotransmitters, and releasing factors<br />

Page 561<br />

[9]. Finally, with respect to N-substituted amides it is also noteworthy to mention the intriguing<br />

approach [21] of replacing the traditional peptide scaffold <strong>by</strong> achiral N-substituted glycine building<br />

blocks. Overall, such Nα-Cα-C scaffold or Cα-Cβ side chain modifications provide significant<br />

opportunities for expanding peptide-<strong>based</strong> molecular diversity (i.e., so-called “peptoid” libraries) as well<br />

as to extend our 3D structural knowledge of traditional φ-Ψ-χ space.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_561.html (1 of 2) [4/9/2004 12:57:49 AM]


Document<br />

Figure 2<br />

Three-dimensional structural properties of peptides: backbone and side chain<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_562.html [4/9/2004 12:58:32 AM]<br />

Page 562


Document<br />

B. Peptidomimetic <strong>Drug</strong>s: Concepts, Strategies, and Technologies<br />

Page 563<br />

Historically, the major focus of peptidomimetic design has evolved from receptor-targeted drug<br />

discovery research and has not been directly impacted <strong>by</strong> an experimentally-determined 3D structure of<br />

the target protein. Nevertheless, a hierarchial approach of peptide rarrow.gif peptidomimetic molecular<br />

design and chemical modification has evolved over the past two decades, <strong>based</strong> on systematic<br />

transformation of a peptide ligand and iterative analysis of the structure-activity and structureconformation<br />

relationships of “second generation” analogs (Figure 3). Such work has typically<br />

integrated biophysical techniques (x-ray crystallography and/or NMR spectroscopy) and computerassisted<br />

molecular modeling with biological testing to advance peptidomimetic drug design.<br />

A plethora of sophisticated synthetic chemistry approaches have entered into the arena of peptide-<strong>based</strong><br />

molecular design, including well-established applications of unusual amino acids and dipeptide<br />

surrogates, among other types of chemical modifications. Such backbone or side chain modifications<br />

may afford stability of the parent peptide to peptidases and have provided conceptual impetus for yet<br />

more sophisticated molecular design and peptidomimetic chemistry studies [1-18]. For example, a few<br />

of the known amide bond replacements (Figure 4) include: aminomethylene or Ψ [CH 2NH], 1;<br />

ketomethylene or Ψ[COCH 2], 2; ethylene or Ψ[CH 2CH 2], 3; olefin or Ψ[CH=CH], 4; ether or<br />

Figure 3<br />

Hierarchial approach in peptide rarrow.gif peptidomimetic structure-<strong>based</strong> drug design<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_563.html [4/9/2004 1:00:07 AM]


Document<br />

Page 564<br />

Ψ[CH 2O], 5; thioether or Ψ[CH 2S], 6; tetrazole or Ψ[CN 4], 7; thiazole or Ψ[thz], 8; retroamide or<br />

Ψ[NHCO], 9; thioamide or Ψ[CSNH], 10; and ester or Ψ[CO 2], 11. These amide bond surrogates<br />

provide insight into the conformational and H-bonding properties that may be requisite for peptide<br />

molecular recognition and/or biological activity at receptor targets. Furthermore, such backbone<br />

replacements can impart metabolic stability towards peptidase cleavage relative to the parent peptide.<br />

The discovery of yet other nonhydrolyzable amide bond isosteres has particularly impacted the design of<br />

protease inhibitors, and these include: hydroxymethylene or Ψ[CH(OH)], 12; hydroxyethylene or<br />

Ψ[CH(OH)CH 2] and Ψ[CH 2CH(OH)], 13 and 14, respectively; dihydroxyethylene or<br />

(Ψ[CH(OH)CH(OH)], 15, hydroxyethylamine or Ψ[CH(OH)CH 2N], 16, dihydroxyethylene 17 and C 2symmetric<br />

hydroxymethylene 18. In the specific case of aspartyl protease inhibitor design (see below)<br />

such backbone modifications have been extremely effective, as they may represent transition state<br />

mimics or bioisosteres of the hypothetical tetrahedral intermediate (e.g., Ψ[C(OH) 2NH] for this class of<br />

proteolytic enzymes.<br />

Both peptide backbone and side chain modifications may provide prototypic leads for the design of<br />

secondary structure mimicry [11, 22–31] as typically suggested <strong>by</strong> the fact that substitution of D-amino<br />

acids, Nα-Me-amino acids, Cα-Me-amino acids, and/or dehydroamino acids within a peptide lead may<br />

induce or stabilize regiospecific β-turn, γ-turn, β-sheet, or α-helix conformations. To date, a variety of<br />

secondary structure mimetics have been designed and incorporated in peptides or peptidomimetics<br />

(Figure 5). The β-turn has been of particular interest to the area of receptor-targeted peptidomimetic<br />

drug discovery. This secondary structural motif exists within a tetrapeptide sequence in which the first<br />

and fourth Cα atoms are < 7 Å separated, and they are further characterized as to occur in a nonhelical<br />

region of the peptide sequence and to possess a ten-membered intramolecular H-bond between the i and<br />

i+4 amino acid residues. On of the initial approaches of significance to the design of β-turn mimetics<br />

was the monocyclic dipeptide-<strong>based</strong> template 19 [22] which employs side chain to backbone constraint<br />

at the i+1 and i+2 sites. Over the past decade a variety of other monocyclic or bicyclic templates have<br />

been developed as β-turn mimetics, and specific examples include 20 [23], 21 [24], 22 [25], 23 [26] and<br />

24 [27]. Most recently, the monocyclic β-turn mimetic 25 has been described [28] and illustrates the<br />

potential opportunity to design scaffolds that may incorporate each of the side chains (i, i+1, i+2 and i+3<br />

positions), as well as five of the eight NH or C=O functionalities, within the parent tetrapeptide<br />

sequence. tetrapeptide sequence modeled in type I–IV β-turn conformations. Similarly, the<br />

benzodiazepine template 26 has shown [29, 30] utility as a β-turn mimetic scaffold which also may be<br />

multisubstituted to simulate side chain functionalization, particularly at the i and i+3 positions of the<br />

corresponding tetrapeptide sequence modeled in type I–VI β-turn conformations. A recently reported<br />

[31]<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_564.html [4/9/2004 1:00:25 AM]


Document<br />

Figure 4<br />

Backbone amide bond surrogates: Ψ[CONH] replacements<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_565.html (1 of 2) [4/9/2004 1:00:45 AM]<br />

Page 565


Document<br />

Figure 5<br />

Secondary structure modifications using β-turn α-turn, and β-sheet scaffolds.<br />

Page 566<br />

γ-turn mimetic, 27, illustrates an innovative approach to incorporate a retroamide surrogate between the<br />

i and i+1 amino acid residues with an ethylene bridge between the N' (i.e., nitrogen replacing the<br />

carbonyl C') and N atoms of the i and i+2 positions, and this template allows the possibility for all three<br />

side chains of the parent tripeptide sequence. Finally, the design of a β-sheet mimetic, 28, provides an<br />

attractive template to constrain the backbone of a peptide to that simulating an extended conformation<br />

[32]. The β-sheet is of particular interest to the area of protease-targeted peptidomimetic drug discovery.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_566.html (1 of 2) [4/9/2004 1:01:15 AM]


Document<br />

12640-0567a.gif<br />

Figure 6<br />

Contemporary drug discovery: integration of structure-<strong>based</strong> drug design,<br />

synthetic chemical libraries, and high-throughput mass screening technologies<br />

Finally, the convergence of structure-<strong>based</strong> drug design (biophysical and computational chemistry),<br />

synthetic chemical libraries, and high-throughput screening technologies have established a new<br />

paradigm for drug discovery (Figure 6). This powerful alliance of scientific disciplines is accelerating<br />

the identification of lead compounds and/or the optimization of drug candidates. The impact to<br />

academic, biotechnological and pharmaceutical research will be immense.<br />

C. Receptor, Protease, and Signal Transduction-Protein Targets<br />

Page 567<br />

<strong>Structure</strong>-<strong>based</strong> drug design and peptidomimetic drug discovery has emerged as a powerful approach in<br />

many areas of pharmaceutical research, including receptor agonists and antagonists, protease inhibitors,<br />

and, more recently, in the rapidly developing area of signal transduction, in which the protein targets<br />

include catalytic and noncatalytic domains of a particular intracellular macro<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_567.html [4/9/2004 1:04:46 AM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_568.html (1 of 2) [4/9/2004 1:07:02 AM]<br />

Page 568


Document<br />

Page 569<br />

molecule. A noncomprehensive listing of such receptor, protease, and signal-transduction protein targets<br />

is shown in Table 1.<br />

With respect to receptor-targeted peptidomimetic drug discovery, the most noteworthy success has been<br />

attained for G-protein-coupled receptor agonists and antagonists as well as, more recently, cell-adhesion<br />

integrin receptor antagonists (see below). However, it is important to state that the impact of highthroughput<br />

screening in the discovery of nonpeptide ligands (typically antagonists) at G-protein-coupled<br />

receptors has yielded extraordinary success. Although screening-<strong>based</strong> nonpeptide drug discovery will<br />

not be extensively reviewed here, the possibility of common pharmacophores between peptide and<br />

nonpeptide ligands may exist (limited cases) in relation to receptor binding. Nevertheless, in most cases<br />

receptor mutagenesis studies suggest the existence of different binding pockets for peptide and<br />

nonpeptide ligands, regardless of whether they both are functionally similar as related to agonism or<br />

antagonism [33]. With respect to both protease-and signal-transduction protein-targeted peptidomimetic<br />

drug discovery, the emergence of 3D structural information to provide high resolution molecular “maps”<br />

of the catalytic (or noncatalytic) domain has provided incredible opportunities for structure-<strong>based</strong> drug<br />

design.<br />

II. Peptide Ligand Lead Discovery and <strong>Structure</strong>-Based <strong>Drug</strong> <strong>Design</strong><br />

As stated previously, peptidomimetic drug discovery was first advanced <strong>by</strong> molecular design concepts<br />

and chemical modification strategies focused on<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_569.html [4/9/2004 1:07:26 AM]


Document<br />

Figure 7<br />

Convergent pathways in peptide, peptidomimetic, and nonpeptide drug<br />

discovery<br />

Page 570<br />

using receptor-targeted peptide ligands or “second generation” agonist or antagonist analogs to develop<br />

pharmacophore models (see Figure 3). However, this represents only a part of the sophisticated<br />

convergent pathways that exist currently to advance both peptidomimetic and nonpeptide drug discovery<br />

(Figure 7). In this scenario both native and foreign (biological or chemical origin) peptides may provide<br />

the opportunity for ligand structure-<strong>based</strong> drug design. From pharmacophore models of key peptide<br />

leads the iterative transformation to peptidomimetic “second generation” analogs may proceed through<br />

either peptide scaffold-or nonpeptide template-<strong>based</strong> approaches. Examples of such work are detailed<br />

below.<br />

Independent of the hierarchial approach to peptide ligand structure-<strong>based</strong> design of peptidomimetics is<br />

the work that encompasses screening-<strong>based</strong> nonpeptide lead discovery. Such nonpeptides may be of<br />

chemical or biological origin (see below), and historically have been identified <strong>by</strong> targeted or random<br />

screening approaches. Although they are not designed to be peptidomimetics in the chemical structure<br />

sense, many appear to be biological mimics (e.g., receptor agonists or antagonists, protease inhibitors,<br />

signal-transduction protein inhibitors, or antagonists) of a known native or foreign peptide. And, in few<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_570.html [4/9/2004 1:08:04 AM]


Document<br />

cases, a pharmacophore model of a nonpeptide lead (or series) may show similarity to that of a<br />

pharmacophore model of a peptide ligand. Examples of such work are detailed below.<br />

A. Peptidomimetics: Receptor Agonists and Antagonists<br />

Page 571<br />

Specific examples that illustrate peptide scaffold-and nonpeptide template-directed drug design<br />

strategies are shown in Figure 8 and include μ-opioid endorphin (END) agonist, 29 [34]; thyrotropinreleasing<br />

hormone (TRH) agonist, 30 [13]; fibrinogen (GPIIa/IIIb) antagonists, 31 [35] and 32 [36];<br />

CCK A antagonist, 33 [37]; CCK B/gastrin antagonist, 34 [38]; endothelin antagonist, 35 [39]; growth<br />

hormone secretagogue (GHRP), 36 [40]; somatostatin agonist (partial), 37 [41], substance-P (NK 1)<br />

antagonists, 38 [42]; neurokinin-A (NK 2) antagonist, 39 [43]; and neurokinin-B (NK 3) antagonist, 40<br />

[44]. For the most part, the above compounds have been advanced as the result of extensive structureactivity<br />

studies and, typically, more focused structural studies (NMR) on a conformationally<br />

constrained, linear or cyclic peptide lead or series. Such structure-conformation activity studies have led<br />

to the development of pharmacophore models to guide iterative structure-<strong>based</strong> design strategies.<br />

One example is that of integrin receptor gpIIb/IIIa antagonists that are structurally derived from the<br />

tripeptide sequence Arg-Gly-Asp, which is common to gpIIb/IIIa protein ligands such as fibrinogen,<br />

vitronectin, fibronectin, von Willebrand factor, osteopontin, thrombospondin, and the collagens [45]. As<br />

shown in Figure 9, transformations of the linear peptide ligand Arg Gly-Asp-Phe <strong>by</strong> both peptide<br />

scaffold (at the Arg-Gly backbone) modification and substitution of the Arg side chain <strong>by</strong> a benzamidine<br />

moiety provided the peptidomimetic lead 31 that is active in vivo as an antiplatelet agent [35]. On the<br />

other hand peptidomimetics such as 32 illustrate nonpeptide template-<strong>based</strong> design strategies derived<br />

from iterative transformations of a cyclic peptide lead in which a<br />

γ-turn about the Asp residue was implicated in a pharmacophore model for the bioactive conformation<br />

[36]. Specifically, the benzodiazepinone substructure of 32 may effectively replace this predicted γ-turn<br />

conformation about the Asp, and the N-Me-Arg replacement with piperidine moiety was also compatible<br />

to high-affinity receptor binding.<br />

A second example is that involving the use of a glucopyranoside nonpeptide template <strong>by</strong> Hirschmann<br />

and coworkers [41, 46] for systematic functionalization to create novel peptidomimetics for the<br />

somatostain (SRIF) and substance-P (NK 1) receptors. As illustrated in Figure 10, the cyclic hexapeptide<br />

SRIF agonist provided a macrocyclic lead structure that was transformed to a glucopyranoside template<br />

designed to substitute for a postulated β turn about the Tyr-D-Trp-Lys-Thr substructure of the parent<br />

peptide ligand. The prototype<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_571.html [4/9/2004 1:08:24 AM]


Document<br />

Figure 8<br />

Receptor-targeted peptidomimetics exemplifying ligand structure <strong>based</strong> drug design<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_572.html [4/9/2004 1:08:54 AM]<br />

Page 572


Document<br />

Figure 9<br />

Peptide scaffold- and nonpeptide template-<strong>based</strong> design strategies:<br />

gpIIb/IIIa antagonists<br />

Page 573<br />

peptidomimetic 37 was found to be a moderately potent SRIF-like agonist (partial) in cellular assays<br />

[41]. This discovery extends previous studies on TRH (see peptidomimetic 30, Figure 8) which utilized<br />

a cyclohexane ring system as a nonpeptide template to functionalize with the pyroglutamic acod and His<br />

side chains as well as the C-terminal carboxamide group of the parent peptide ligand [13]. However, in<br />

the comparative analysis of analogs of the SRIF-mimetic 37 it was also found that N-acetylation of the<br />

Lys side chain moiety yielded a potent antagonist of substance-P (NK 1 receptor). This indicated that<br />

slightly different functionalization of the nonpeptidic glucopyranoside template was quite compatible<br />

with NK 1 receptor molecular recognition. Intuitively, a “reversed design” strategy to convert the latter<br />

glucopyranoside-<strong>based</strong> NK 1 ligand to a cyclic peptide was next investigated (Figure 10), and<br />

successfully led to the discovery of a novel cyclic peptide ligand also exhibiting potent NK 1 receptor<br />

binding and antagonism [46].<br />

The above examples of peptide scaffold- or nonpeptide template-<strong>based</strong> peptidomimetic agonists or<br />

antagonists illustrate various strategies to elaborate bioactive conformation and/or pharmacophore<br />

models of peptide ligands at their receptors. In many cases, receptor subtype selectivity has also been<br />

achieved <strong>by</strong> systematic structural modifications of prototypic leads of peptidomimetics. Thus, although<br />

the 3D structures of G-protein-coupled receptors (GPCRs) remain as elusive (except for models<br />

constructed from homology-<strong>based</strong> low-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_573.html (1 of 2) [4/9/2004 1:09:25 AM]


Document<br />

Figure 10<br />

Peptide scaffold- and nonpeptide template-<strong>based</strong> design strategies: somatostatin<br />

agonist and substance-P (NK 1) antagonists<br />

resolution 3D structures of bacteriorhodopsin or rhodopsin, see below) the development of<br />

Page 574<br />

pharmacophore models using the hierarchial approach in peptide rarrow.gif peptidomimetic structure<strong>based</strong><br />

drug design (see above, Figure 3) remains quite promising to creative science. Indeed, targetedreceptor<br />

screening of synthetic chemical libraries of highly modified peptide molecules (e.g., Nsubstituted<br />

Gly “peptoids” [21] is rapidly expanding our database of structurally diverse ligands. Such<br />

structurally unique lead compounds will provide the opportunity for further pharmacophore modeling<br />

strategies to be used in the discovery of novel agonists or antagonists at GPCR and other receptor types.<br />

B. Peptidomimetics: Protease Inhibitors<br />

Specific examples of peptidomimetics which illustrate peptide scaffold- and nonpeptide templatedirected<br />

drug-design strategies as applied to protease in-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_574.html (1 of 2) [4/9/2004 1:09:58 AM]


Document<br />

Figure 11<br />

Protease-targeted peptidomimetics derived <strong>by</strong> ligand structure-<strong>based</strong> drug<br />

design<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_575.html [4/9/2004 1:10:57 AM]<br />

Page 575


Document<br />

Page 576<br />

hibitors (Figure 11) include renin inhibitors, 41 [47] and 42 [48]; HIV protease inhibitors 43 [49] and 44<br />

[50]; angiotensin-converting enzyme inhibitors 45 [51] and 46 [52]; collagenase inhibitor, 47 [53];<br />

gelatinase inhibitor, 48 [54], stromelysin inhibitor, 49 [55]; elastase inhibitor, 50 [56]; thrombin<br />

inhibitors, 51 [57] and 52 [58]; and interleukin-converting enzyme inhibitor, 53 [59]. In general, the<br />

design of protease inhibitors has focused on both the natural substrate structure and the mechanism of<br />

substrate cleavage to provide “first-generation” inhibitors. Also, these initial leads are typically peptide<br />

scaffold-<strong>based</strong> to provide the possibility for β-sheet conformation, which may permit “extensive” Hbonding<br />

between the backbone amide groups of the inhibitor and complementary H-bond donor or<br />

acceptor groups of the enzyme active site. Furthermore, the traditional approach to designing protease<br />

inhibitors includes the substitution of nonhydrolyzable amide surrogates (see below Figure 4) at the P1- P1' cleavage site. Specificity to a particular protease may sometimes be extrapolated directly from the<br />

primary structure of the substrate (e.g., human renin substrate specificity is conferred from the<br />

angiotensinogen N-terminal octapeptide sequence ~His6-Pro-Phe-His-Leu Val-Ile-His13~ in which<br />

refers to the cleavage site). In many cases substitution of the scissile amide (substrate) <strong>by</strong> a “transition<br />

state” bioisosteres or an electrophilic ketomethylene moiety have provided tight-binding “first<br />

generation” pseudopeptide inhibitor leads.<br />

One example of protease inhibitor design that illustrates the peptide scaffold-<strong>based</strong> approach is that for<br />

HIV protease inhibitors. Albeit over the past several years HIV protease inhibitor research has become a<br />

highly advanced example of iterative structure-<strong>based</strong> drug design (see below), the first discoveries of<br />

pseudopeptide and peptidomimetic inhibitors of this aspartyl protease were not made with knowledge of<br />

the 3D structure of the target enzyme. Specifically, the natural product pseudopeptide pepstatin (Figure<br />

12), a typical inhibitor of the aspartyl protease family of enzymes, was determined to be weakly potent<br />

against HIV-1 protease [60]. Relative to pepstatin, the central P 1-P 1' statine (i.e., Sta or<br />

LeuΨ[CH(OH)]Gly) moiety was further evaluated within the context of an “optimized” N- and Cterminal<br />

amino acid sequence using a chemical-library strategy [61]. As shown in Figure 12, a resultant<br />

pseudo-tetrapeptide Ac-Trp-Val-Sta-D-Leu-NH 2 was found to be a relatively potent HIV protease<br />

inhibitor. In another approach, a designed renin inhibitor (55; Figure 12) was determined to be a highly<br />

potent HIV protease inhibitor [62]. Further optimization studies led to the discovery [49] of the first<br />

bonafide peptidomimetic inhibitor of HIV protease (Tba-ChaΨ[CH(OH)CH 2] Val-Ile-Amp; 43) of HIV<br />

protease. This compound (43) provided the first evidence of cellular anti-HIV activity and, therefore,<br />

supported proof-of-concept studies related to the therapeutic significance of targeting HIV-1 protease.<br />

Replacement of the peptide scaffold <strong>by</strong> the pyrrolidinone-type β-sheet mimetic 28 in a<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_576.html [4/9/2004 1:11:16 AM]


Document<br />

Figure 12<br />

Peptide scaffold-<strong>based</strong> design strategies: HIV protease inhibitors<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_577.html (1 of 2) [4/9/2004 1:11:42 AM]<br />

Page 577


Document<br />

Page 578<br />

chemically-related P 1-P 1' PheΨ[CH(OH)CH 2] Phe-modified lead has been reported [63] to yield<br />

effective peptidomimetic inhibitors of the HIV-1 protease (56; Figure 12). The pyrrolidinone-type lead<br />

has shown enhanced cellular permeability relative to its peptide backbone-type counterparts. In a third<br />

approach guided <strong>by</strong> HIV substrate-<strong>based</strong> design, the cleavage site dipeptide Phe-Pro was substituted <strong>by</strong><br />

the “transition state” bioisostere to provide the highly potent and selective HIV protease inhibitor 57, a<br />

P 1-P 1' PheΨ[CH(OH)CH 2N] Pro-modified heptapeptide [64]. As compared to this pseudopeptide, a<br />

pioneering effort focused on peptide ligand structure-<strong>based</strong> design provided a second series of highly<br />

potent, selective, and cellularly-active HIV protease inhibitors [50] as represented <strong>by</strong> the recently FDAapproved<br />

anti-HIV drug 44 (Saquinavir). The design of still more HIV protease inhibitors having novel<br />

chemical structures (e.g., C 2-symmetric scaffolding, P 1-P 1' “transition state” bioisostere cyclization, and<br />

achiral nonpeptide template replacement) has progressed at an extraordinary pace (for reviews see<br />

Reference 65), and in a majority of cases such work has been strongly impacted <strong>by</strong> knowledge of the 3D<br />

structure of the target enzyme and/or inhibitor complexes of it (see below).<br />

A second example of protease inhibitor design that properly illustrates the peptide scaffold-<strong>based</strong><br />

approach is that of thrombin inhibitors. Work with these compounds led to the identification of highly<br />

potent, selective, and in vivo-effective lead compounds. A member of the serine protease family,<br />

thrombin cleaves a number of substrates (e.g., fibrinogen) and activates its platelet receptor (a G-proteincoupled<br />

receptor) <strong>by</strong> proteolysis of the extracellular N-terminal domain which results in self-activation<br />

(for a review see Reference 66). Initial lead inhibitors of thrombin were substrate-<strong>based</strong>, including the<br />

fibrinogen P 3-P 1 Phe-Pro-Arg sequence [67] and simple Arg derivatives such as Tos-Arg-OMe [68].<br />

Also, the natural product cyclothreonide-A, a macrocyclic peptide containing a Pro-Arg ketoamide<br />

sequence, provided an inhibitory peptide ligand lead [69]. As shown collectively in Figure 13,<br />

compounds 52 and 58–60 provided the opportunity to try different strategies to advance the design of<br />

thrombin inhibitors. Particularly noteworthy from these early peptidomimetic lead discovery studies was<br />

the design effort [70] that led to the highly potent thrombin inhibitor 52 (Agatroban), a sulfonamidemodified<br />

Arg derivative, which incorporated an unusual cyclic amino acid substituent C-terminal to the<br />

P 1 moiety as opposed to reactive electrophilic groups (e.g., ketone, aldehyde, or boronic acid).<br />

Interestingly, replacement of the Arg side chain moiety within a structurally similar analog 60 <strong>by</strong> a<br />

amidinobenzyl group was shown [71] to be optimal when the stereochemistry at the P 1 α-carbon has a Dconfiguration<br />

suggesting that the mode of binding may be different for 52 versus 60. In this regard, xray<br />

crystallographic analysis of thrombin-inhibitor complexes (see below) have provided insight in the<br />

interpretation of the structure-activity relationships of the aforementioned lead compounds.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_578.html [4/9/2004 1:11:53 AM]


Document<br />

Figure 13<br />

Peptide scaffold-<strong>based</strong> design strategies: thrombin and tripeptidyl peptidase II<br />

inhibitors<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_579.html (1 of 2) [4/9/2004 1:12:20 AM]<br />

Page 579


Document<br />

Page 580<br />

The recent discovery of peptidomimetic inhibitors of the serine protease TTP-II (tripeptidyl peptidase-II)<br />

further illustrates the peptide scaffold-<strong>based</strong> design approach [72]. Specifically, relative to a known TTP-<br />

II cleavage site on the endogenous neuropeptide CCK-8 (i.e., Asp 1-Tyr[SO 3H]-Met arrowd.gif Gly-Trp-<br />

Met-Asp-Phe 8-NH 2) the design of a highly potent inhibitor 61 (Figure 13) was successfully achieved <strong>by</strong><br />

iterative structure-<strong>based</strong> optimization of the P3-P1 sequence. Noteworthy of the relatively simple<br />

structure of the TTP-II inhibitor 61 was that it contains no functional group C-terminal to the P 1 αcarbon.<br />

Such absence of an electrophilic moiety, “transition state” bioisostere, or other type of<br />

nonhydrolyzable amide substituent is rather unique relative to most examples of substrate-<strong>based</strong><br />

protease inhibitors.<br />

C. Peptidomimetics: Signal-Transduction Protein Inhibitors and Antagonists<br />

Beyond receptors and proteases exists the rapidly emerging area of signal-transduction protein-targeted<br />

drug discovery research. To date, a multitude of catalytic and noncatalytic proteins have been identified<br />

which are critical components of intracellular signal-transduction pathways. These signal-transduction<br />

proteins provide the molecular basis for communication from extracellular “effectors” (e.g., hormones,<br />

neurotransmitters, growth factors, and cytokines) to stimulate cells in specific and regulated manner.<br />

Signal-transduction pathways often involve protein-protein interactions, including examples of enzymesubstrate<br />

(e.g., kinases, phosphatases, transferases, and isomerases) as well nonenzymatic complex<br />

formation (e.g., “adapter” proteins, exchange factors, and transcription factors). As compared to receptor-<br />

or protease-targeted peptidomimetic drug discovery, there are significantly fewer examples reported in<br />

the field of signal-transduction research. Thus, in several cases peptide ligands (as “prototype<br />

peptidomimetic leads”) will be described to illustrate opportunities for structure-<strong>based</strong> drug design.<br />

Specific examples which illustrate peptide scaffold- and nonpeptide template-directed drug-design<br />

strategies are shown in Figure 14 and include: Ras farnesyl transferase inhibitors, 62 [73], 63 [74], 64<br />

[75], and 65 [76], Src SH2 domain antagonists, 66 [77], 67 [78], and 68 [79]; and the protein tyrosine<br />

phosphatase PTP1B inhibitor 69 [80].<br />

An example of the signal-transduction protein-targeted inhibitor design which illustrates both peptide<br />

scaffold- and nonpeptide template-<strong>based</strong> approaches is that for the Ras farnesyl transferase inhibitor<br />

discovery. Such compounds show potential as new therapeutic agents for Ras-related carcinogenesis<br />

[81]. Substrate sequences for farnesyl transferase have the consensus ~Cys-AA 1-AA 2-Met motif (AA<br />

refers to Val or Ile). Both substrate-<strong>based</strong><br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_580.html [4/9/2004 1:13:34 AM]


Document<br />

Figure 14<br />

Signal-transduction protein-targeted peptidomimetics derived <strong>by</strong> structure-<strong>based</strong> drug design<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_581.html [4/9/2004 1:13:54 AM]<br />

Page 581


Document<br />

Page 582<br />

inhibitors [73–75,82,83] and, more recently, a novel non-Cys containing peptide inhibitors [76,84] have<br />

led to potent and cellularly active compounds. As illustrated in Figure 15, the “collected” substrate<strong>based</strong><br />

inhibitor 70 was designed to covalently attach farnesyl to a peptide via a phosphinic acid linker<br />

replacement for [82], and this compound has been determined to be both potent against the target<br />

enzyme and cellularly effective. Relative to peptide substrate structure-<strong>based</strong> design efforts,<br />

peptidomimetics incorporating Ψ[CH 2NH]-substitutions (e.g., 62, [73]) or a benzodiazepinone<br />

replacement of the central dipeptide moiety (e.g., 63, [74]) have yielded high affinity inhibitors. Another<br />

series of very potent Ras farnesyl transferase inhibitors have been designed in which the central<br />

dipeptide has been substituted <strong>by</strong> various isomeric and/or homologated derivatives of amino benzoic<br />

acid (e.g., 64 [75]), including a particularly effective analog biphenyl derivative 71 [83]. The above<br />

studies indicated that both conformationally flexible or constrained peptide scaffolds as well as<br />

nonpeptide template replacements can be used to “link” the Cys and Met substructures. It is also<br />

important to point out that although compounds such as 62–64 have “free” sulfhydryl groups (Cys) there<br />

is no evidence that they become farnesylated, and therefore the binding mode and effect on catalytic<br />

function of the target enzyme are unique relative to their peptide substrate counterparts. Recently, a<br />

novel peptide inhibitor series, as exemplified <strong>by</strong> Cbz-His-Try(O-benzyl)-Ser(O-benzyl)-Trp-D-Ala-NH 2,<br />

has been discovered [76]. These inhibitors do not contain a Cys residue and structure-<strong>based</strong> design<br />

efforts have successfully led to a series of peptidomimetics (e.g., 65) having only one chiral center.<br />

Interestingly, this novel inhibitor series has been determined to competitively inhibit farnesyl<br />

pyrophosphate binding rather than the binding of peptide substrate to the target enzyme. Another Hissubstituted<br />

peptidomimetic inhibitor of Ras farnesyl transferase has been recently reported [84] as<br />

exemplified <strong>by</strong> 73, which was designed relative to a peptide substrate-<strong>based</strong> parent analog (72, Figure<br />

15). Although the 3D structure of Ras farnesyl transferase is not known, biochemical studies suggest<br />

that a divalent metal ion (e.g., Zn 2+) may coordinate with the above inhibitor sulfhydryl or imidazole<br />

groups at their corresponding binding sites on the target enzyme.<br />

Another example of the signal-transduction protein-targeted drug design that illustrates peptide scaffold<strong>based</strong><br />

approaches is that for Src SH2 domain antagonist discovery. Such compounds show promise as<br />

new therapeutic agents for Src-related carcinogenesis, osteoporosis, and immune diseases [85]. The Src<br />

SH2 domain is a prototype example of a superfamily of intracellular signal-transduction proteins that<br />

possess structurally homologous SH2 domains that specifically recognize cognate phosphoproteins in a<br />

sequence-dependent manner relative to a critical phosphotyrosine (pTyr) residue (i.e., ~pTyr-AA 1-AA 2-<br />

AA 3-AA 4~ Furthermore, recent x-ray crystallographic studies of a several SH2 protein targets (e.g.,<br />

phosphopeptide complexes) is greatly impacting the<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_582.html [4/9/2004 1:14:40 AM]


Document<br />

Figure 15<br />

Peptide scaffold- and nonpeptide template-<strong>based</strong> design strategies: farnesyl transferase<br />

inhibitors<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_583.html [4/9/2004 1:15:15 AM]<br />

Page 583


Document<br />

Page 584<br />

opportunity for iterative structure-<strong>based</strong> drug design in this field or research (see below [86]). Relative<br />

to Src SH2 domain antagonist lead discovery, peptide library studies [87] have shown the ~pTyr-Glu-<br />

Glu-Ile~ as a preferred consensus sequence. Peptide scaffold-<strong>based</strong> approaches to replace the internal<br />

dipeptide, Glu-Glu, <strong>by</strong> both flexible and rigid linkers have been explored [88] but were unsuccessful in<br />

yielding potent analogs. As shown in Figure 16, prototype peptidomimetics 66 [77] and 67 [78] illustrate<br />

a successful approach in which stereoinversion at the second residue (P+2 relative to the pTyr) to the Dconfiguration<br />

and side chain substitution to hydrophobic functionalities (e.g., cyclohexyl and naphthyl)<br />

which provides accessibility to the known hydrophobic binding pocket for the P+3 Ile side chain.<br />

Indeed, such compounds showed binding affinities essentially identical to that of the N- and Cterminally<br />

extended phosphopeptides containing the pTyr-Glu-Glu-Ile sequence. Substitution of the<br />

pTyr residue of 66 <strong>by</strong> the difluoromethyl-phosphonate modified analog F 2Pmp provides a more<br />

metabolically stable derivative 74 (Figure 16) and a prototype lead to advance the design of cellularly<br />

active second-generation compounds. More recently, structure-<strong>based</strong> drug design studies of compound<br />

65 have led to the discovery of potent and Src SH2-selective peptidomimetic lead compounds [89], and<br />

this is further detailed below as related x-ray crystallographic structures of Src SH2-phosphopeptide<br />

complexes.<br />

III. Nonpeptide Ligand Lead Discovery and <strong>Structure</strong>-Based <strong>Drug</strong> <strong>Design</strong><br />

As previously illustrated in Figure 7, the convergent pathways to design drugs which act as mimics<br />

(agonists), antagonists, or inhibitors of native peptide (protein) ligands at their target receptors,<br />

proteases, signal-transduction proteins, and so forth, include “foreign” nonpeptides. The origin of such<br />

nonpeptides may be either biological (e.g., natural product) or chemical (synthetic compound collection<br />

or libraries) that have been subject to biochemical screening to identify leads for further molecular<br />

design and structure-activity studies. Over recent years the success of screening-derived nonpeptide lead<br />

discovery and iterative transformation to drug candidates has been quite impressive, and many aspects<br />

of this area of research have been reviewed [12, 15]. Nevertheless, it is intriguing to explore the<br />

potential relationship between peptide ligands (including peptidomimetic derivatives) and such<br />

screening-derived nonpeptide ligands as related to pharmacophore modeling and structure-<strong>based</strong> drugdesign<br />

studies.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_584.html [4/9/2004 1:15:50 AM]


Document<br />

Figure 16<br />

Peptide scaffold-<strong>based</strong> design strategies: Src SH2 domain antagonists<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_585.html [4/9/2004 1:16:28 AM]<br />

Page 585


Document<br />

A. Molecular Diversity and Screening-Based Identification of Nonpeptide Ligands<br />

Page 586<br />

Precedence for the success of nonpeptide drug discovery can be traced to the identification of morphine<br />

(75; Figure 17) as nonpeptide natural product agonist at μ-opioid peptide receptors [90]. To date, the<br />

robust momentum of nonpeptide drug discovery continues to be accelerated <strong>by</strong> sophisticated<br />

biochemical screening technologies. The scope of molecular diversity as well as therapeutic targets for<br />

such screening-<strong>based</strong> nonpeptide ligand lead compounds includes the following examples (Figure 17):<br />

substance P (NK 1 antagonist, 76 [91]; angiotensin AT 1 antagonist, 77 [92]; growth hormone-releasing<br />

peptide (GHRP) agonist, 78 [93]; cholecystokinin CCK A antagonist, 79 [94]; CCK B/gastrin antagonist,<br />

80 [95]; CCK A agonist, 81 [96]; endothelin antagonist, 82 [97]; gonadotropin-releasing hormone<br />

(GnRH) antagonist, 83 [98]; vasopressin V 1 antagonist, 84 [99]; gastrin-releasing peptide antagonist, 85<br />

[100]; glucagon antagonist, 86 [101], neurotensin antagonist, 87 [102]; angiotensin AT 1 agonist, 88<br />

[103]; oxytocin antagonist, 89 [104]; and HIV protease inhibitor, 90 [105]. A significant number of<br />

screening-derived nonpeptide leads have been identified for G-protein-coupled receptors (GPCRs), and<br />

in a majority of cases these compounds have been determined to be competitive antagonists. Albeit the<br />

3D structures of this receptor superfamily have not been directly determined, homology model-building<br />

and site-directed mutagenesis studies are impacting structure-activity analysis of agonist and antagonist<br />

ligands (peptide, peptidomimetic, and nonpeptide) for several GPCR targets (see below).<br />

B. Nonpeptides: Exploring Pharmacophore Relationships to Peptide Ligands<br />

Relative to a number of GPCR targets, there exists significant opportunity to compare chemical<br />

structures and 3D pharmacophore models of both peptide and nonpeptide ligands. Such comparative<br />

analyses can explore the possibility of similar 3D substructural elements that may account for their<br />

molecular recognition at the binding site(s) of the receptor. The fact that a vast number of screeningderived<br />

nonpeptide leads are multifunctionalized 5- to 7-membered ring heterocycles (e.g., alkaloid and<br />

benzodiazepine) and contain conformationally rigid substructural elements (e.g., biphenyl, spiro-bicyclic<br />

rings, and N-substituted amide or amine linkages) suggests the likelihood that such compounds are<br />

binding with highly favorable entropic driving forces as compared to the more conformationally flexible<br />

peptide ligands. In this regard, efforts to “rigidify” peptide-<strong>based</strong> scaffolding or replace it <strong>by</strong> nonpeptide<br />

templates has been the underlying theme of peptidomimetic design strategies, and concepts for the latter<br />

approach date back to the proposed used of cycloaliphatic ring systems that might be<br />

multifunctionalized to create “topographi-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_586.html [4/9/2004 1:16:39 AM]


Document<br />

Figure 17<br />

Nonpeptide drug discovery: examples from screening-<strong>based</strong> approaches<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_587.html [4/9/2004 1:17:33 AM]<br />

Page 587


Document<br />

Page 588<br />

cally-designed” peptidomimetics (or, as also defined, “peptoids” [106]). Nevertheless, screening-<strong>based</strong><br />

nonpeptide drug discovery has advanced a treasure of structure-function information to provide insight<br />

into both structure-<strong>based</strong> design and molecular recognition [12,15,107]. In a few (limited) cases, there<br />

exists a likely possibility of similar pharmacophoric features or substructural elements between<br />

nonpeptides and their peptide-ligand counterparts (Figure 18).<br />

Historically, drug discovery research on opioid GPCR receptor targets (e.g., μ, δ, κ) has provided insight<br />

to explore the pharmacophores of both agonist and antagonists derived from endogenous peptides (e.g.,<br />

endorphin, endorphin, and dynorphin) versus nonpeptides (e.g., the μ-receptor selective agonist<br />

morphine and its N-allyl-substituted antagonist derivative naloxone). Relative to the N-terminal Tyr<br />

moiety (side chain and α-amino functionalities) of the endogenous opioid peptides [108], the N-methyltyramine<br />

substructure of morphine represents a likely common pharmacophore for agonist ligand<br />

binding to the μ-receptor (Figure 18). In the case of angiotensin II receptor antagonist drug discovery, it<br />

has been proposed [109] that a common pharmacophore may exist relative to the C-terminal His-Pro-<br />

Phe-OH sequence of angiotensin II and nonpeptide 91 (Figure 18). In fact, these studies provided design<br />

insight leading to the discovery of the drug candidate 77 (Lorsartan). A third example in which<br />

correlation between peptide and nonpeptide pharmacophore models becomes apparent is that of<br />

neuropeptide-Y (NPY) versus the benextramine-<strong>based</strong> derivative 92 [110] or arpromidine-<strong>based</strong><br />

derivative 93 [111] as illustrated in Figure 18. In both cases, the C-terminal Arg-Gln-Arg-Tyr-NH 2<br />

sequence of NPY was modeled relative to the nonpeptide structures such that the guanido functionalities<br />

were superimposed upon the corresponding basic (i.e., guanido or imidazole) substructural elements of<br />

either 92 or 93.<br />

In some cases, the availability of x-ray crystallographic information of both the peptide and nonpeptide<br />

ligands may provide insight into the pharmacophore modeling studies. An example of this exists for<br />

oxytocin antagonist structure-<strong>based</strong> drug design studies [112]. As shown in Figure 19, pharmacophore<br />

models of both a cyclic hexapeptide oxytocin antagonists and conformationally-constrained,<br />

tolylpiperazine camphorsulfonamide nonpeptide antagonist (89) suggest the likelihood of common<br />

substructural elements that were key for molecular recognition at the oxytocin receptor, and led to the<br />

design of a highly potent derivative 94. Nevertheless, such comparative pharmacophore “mapping”<br />

studies are very simplistic since the 3D structures of the target receptors are not known. Furthermore,<br />

site-directed mutagenesis studies<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_588.html [4/9/2004 1:17:54 AM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_589.html (1 of 2) [4/9/2004 1:18:08 AM]<br />

Page 589


Document<br />

Figure 18<br />

Comparative substructural elements of peptide and nonpeptide ligands: μ-opioid receptor<br />

agonists, angiotensin, and NPY receptor antagonists<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_589.html (2 of 2) [4/9/2004 1:18:08 AM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_590.html (1 of 2) [4/9/2004 1:18:20 AM]<br />

Page 590


Document<br />

Figure 19<br />

Comparative substructural elements of peptide and nonpeptide ligands: examples<br />

of oxytocin receptor antagonists and HIV protease inhibitors<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_590.html (2 of 2) [4/9/2004 1:18:20 AM]


Document<br />

Page 591<br />

often suggest that peptide and nonpeptide ligands have different modes of binding to their receptors (see<br />

below). It is also important to point out that the discovery of nonpeptide agonists will likely provide<br />

important structural information to advance our understanding of ligand binding and activation of<br />

receptors as well as insight for pharmacophore modeling. In this regard, nonpeptide agonists 78 (growth<br />

hormone-releasing peptide receptor) and 81 (cholecystokinin receptor CCK A subtype) are noteworthy<br />

exceptions to the rule that nonpeptide screening-<strong>based</strong> leads are antagonists (see above, Figure 17).<br />

Finally, beyond receptor targets the advantages of high-throughput screening of chemical files, natural<br />

products, and synthetic libraries are increasing for proteases as well as other enzyme and noncatalytic<br />

targets. An recent example of such efforts is the nonpeptide HIV protease inhibitor 90 [105], which was<br />

originally identified from screening a chemical file. Through iterative structure-<strong>based</strong> drug design<br />

studies, including x-ray crystallographic analysis of both ligand and inhibitor-enzyme complexes, the<br />

pyrone template has led to the discovery of highly potent, selective, and cellularly active lead<br />

compounds (see below). In retrospect, the original concept of superimposing the key substructural<br />

elements of the nonpeptide ligand 90 to a known peptidomimetic inhibitor of HIV protease is illustrated<br />

in Figure 19. A more detailed account of this effort and successful elaboration of the nonpeptide lead<br />

structure is described below.<br />

IV. Protein Target 3D Structural Models and <strong>Structure</strong>-Based <strong>Drug</strong> <strong>Design</strong><br />

A significant impact in both peptidomimetic and nonpeptide drug discovery has emerged over recent<br />

years as the result of the determination of the 3D structures of protein targets <strong>by</strong> x-ray crystallography or<br />

NMR spectroscopy [113–119]. In addition, computational methodologies such as QSAR, 3D<br />

QSAR/CoMFA, homology-modeling, ligand docking, molecular dynamics, and mechanics, and solventaccessible<br />

surface visualization have greatly impacted such research. Furthermore, programs such as<br />

GRID, GROW, GrowMol, LEGEND, BUILDER, LUDI, FOUNDATION-SPLICE, and CONCERTS<br />

have enabled 2D and 3D database searching and de novo design [115,120,121]. Overall, the iterative<br />

cycle of structure-<strong>based</strong> drug design (Figure 20) has evolved to be an “engine of invention” for several<br />

examples of peptidomimetic and nonpeptide drug discovery.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_591.html [4/9/2004 1:18:32 AM]


Document<br />

A. Receptor Targets<br />

Figure 20<br />

Iterative cycle(s) for structure-<strong>based</strong> drug design<br />

Page 592<br />

Amongst the known superfamilies of cell membrane anchored receptors, significant research has been<br />

focused on GPCR targets (Table 1). The GPCR superfamily of receptors consist of seven<br />

transmembrane-spanning (TM) helices. Sequence homology among them varies from 25–70% (for<br />

reviews see References 33, 122–132). The initial development of 3D structural models of GPCR targets<br />

has been developed from homology-building methodologies <strong>based</strong> on a low-resolution structure of<br />

bacteriorhodopsin [133]. Representative examples of recent studies that provide insight to<br />

pharmacophore modeling and structure-<strong>based</strong> drug design of peptide, peptidomimetic, and nonpeptide<br />

ligands are discussed below.<br />

3 G-Protein-Coupled Receptors<br />

Recent studies have explored several GPCR targets with respect to ligand-receptor binding interactions<br />

using site-directed mutagenesis to explore agonist<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_592.html (1 of 2) [4/9/2004 1:18:59 AM]


Document<br />

Page 593<br />

versus antagonist ligands, and such work provides insight to pharmacophore modeling. Specific<br />

examples of such work as focused on peptide, peptidomimetic, and/or nonpeptide ligands, and working<br />

3D structural models of their GPCR targets include angiotensin II AT 1 and AT 2 subtypes [134],<br />

neurokinin NK 1 and NK 2 subtypes [135], cholecystokinin/gastrin CCK A and CCK B subtypes [136],<br />

opioid μ-, δ-, and κ-subtypes [137], vasopressin V 1A subtype [138], bradykinin B 2 subtype [139],<br />

neurotensin [140] and α-melanotropin MC 1 subtype [141]. From such work it has been inferred that<br />

different binding-site interactions may exist for peptide versus nonpeptide ligands as <strong>based</strong> on their<br />

differential sensitivities to site-directed mutants of the native GPCR. The recent development of 3D<br />

structural models of the neurotensin [140] and α-melanotropin MC 1 subtype [141] GPCRs provide<br />

interesting case studies. Both examples provide the correlation of significant structure-activity databases<br />

and experimentally determined (NMR) structures of key peptide analogs with predicted molecular<br />

contacts at their respective target receptors. As illustrated in Figure 21, the proposed peptide agonist<br />

binding interactions for neurotensin and α-melanotropin analogs at their human GPCR targets may be<br />

used to further guide the molecular design and synthesis of “second-generation” peptidomimetic<br />

derivatives.<br />

In the first example, the neurotensin C-terminal octapeptide was subject to conformational searching<br />

(~Arg-Pro-Tyr~ sequences from the Brookhaven Protein Databank), manual docking to the homologybuilt<br />

neurotensin GPCR receptor model, and constrained molecular dynamics simulation to provide a 3D<br />

structure of the ligand-receptor complex [140]. A compact structure of the peptide in its complexed<br />

conformation was consistent with a Type-1 β-turn as previously determined <strong>by</strong> structural and structureactivity<br />

studies. Key molecular contacts predicted from this neurotensin GPCR model include<br />

hydrophobic interactions with the C-terminal Ile and Leu side chains, π-cation interactions with each<br />

Arg residue side chain, and a “cluster” of aromatic-aromatic interactions with the Tyr side chain. No<br />

electrostatic interactions were predicted, and the primary contact residues on the neurotensin GPCR<br />

model were those comprising the third extracellular loop.<br />

In the second example, the α-melanotropin (MC1) GPCR model was constructed [141] <strong>by</strong> homologybuilding<br />

methods relative to both bacteriorhodopsin and rhodopsin fingerprint maps, and the MSH<br />

superagonist peptides [Nle 4, D-Phe 7]-MSH and Ac-cyclo[Nle 4, Asp 5, D-Phe 7, Lys 10]-MSH 4–10-NH 2 were<br />

modeled in conformations derived from previous experimental studies (i.e., a Type-II β-turn at the<br />

common tetrapeptide sequence ~His-D-Phe-Arg-Trp~). Of the alternative binding modes that were<br />

described for the above the two MSH peptide ligands, one predicts the possibility of a network of<br />

aromatic-aromatic and hydrophobic interactions between the MC1 receptor and the D-Phe and Trp side<br />

chains of the MSH ligand (Figure 21). In addition, this MC1 receptor model predicts multiple<br />

electrostatic and π-cation interactions between<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_593.html [4/9/2004 1:19:20 AM]


Document<br />

Figure 21<br />

GPCR 3D structural models for neurotensin and α-melanotropin agonists<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_594.html (1 of 2) [4/9/2004 1:20:08 AM]<br />

Page 594


Document<br />

Page 595<br />

the MC1 receptor and the Arg side chain of the peptide ligand. The primary contact residues of this<br />

particular MC1 receptor model were all transmembrane domain derived and lie within 4–7.5 angstrom<br />

(centroid to centroid). In conclusion, the development and refinement of 3D structural models of GPCR<br />

targets, iterative site-directed mutagenesis studies, and systematic testing of key agonists and/or<br />

antagonists will undoubtedly make a significant impact in the structure-<strong>based</strong> drug design of<br />

peptidomimetic and nonpeptide ligands at these receptors. Such work may be expected to synergize well<br />

with ligand-<strong>based</strong> pharmacophore modeling strategies which have become quite sophisticated in recent<br />

years as the results of advanced computational chemistry methodologies.<br />

B. Protease Targets<br />

Protease-inhibitor drug discovery illustrates significant success in both mechanistic and 3D structure<strong>based</strong><br />

drug discovery for each of the representative classes (i.e., aspartyl, serinyl, metallo, and cysteinyl)<br />

as exemplified in Table 2 (for a review see Reference 142a). In retrospect, pioneering achievements in<br />

the design of peptidomimetic inhibitors of angiotensin-converting enzyme (for reviews see References<br />

142b,142c) to have led to 45 (Captopril [51]) and 46 (Enalapril [52]). Such work has provided great<br />

impetus to the area of proteasetargeted drug discovery. Over the past two decades a pervasive effort<br />

integrating substrate-<strong>based</strong> inhibitor design, x-ray crystallography or NMR spectroscopy of inhibitorprotease<br />

complexes, high-throughput mass screening, and combinatorial chemical technologies has<br />

evolved to further advance this area of research.<br />

Aspartyl Proteases<br />

The aspartyl proteases include pepsin, renin, cathepsin-D, chymosin, and gastricsin as well as microbial<br />

enzymes (e.g., penicillopepsin, rhizopuspepsin, and endothiapepsin) and retroviral proteases (e.g., HIV-<br />

1 protease). The first high-resolution x-ray crystallographic structures of this protease family were<br />

determined for penicillopepsin [143], rhizopuspepsin [144], endothiapepsin [145], pepsinogen [146],<br />

and pepsin [147]. Based on homology-building strategies, 3D structural models of renin were<br />

subsequently constructed [148] to first guide the structure-<strong>based</strong> design of peptidomimetic inhibitors (for<br />

a review see Reference 149). Furthermore, in several cases the x-ray crystallographic structures of renin<br />

inhibitors were determined [150] as ligand-enzyme complexes with rhizopuspepsin, endothiapepsin, or<br />

pepsin. Eventually, the x-ray crystallographic structures of renin (apo/complexes) were achieved to<br />

provide high-resolution molecular maps of the target enzyme [151]. As illustrated in Figure 22, substrate<strong>based</strong><br />

inhibitors such as the highly potent peptidomimetic 95 [151a] show well-defined hydrophobic<br />

pockets for the P 3-P 1' side chains as well<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_595.html [4/9/2004 1:20:17 AM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_596.html (1 of 2) [4/9/2004 1:22:40 AM]<br />

Page 596


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_597.html (1 of 2) [4/9/2004 1:24:05 AM]<br />

Page 597


Document<br />

Page 598<br />

as hydrogen-bonding to the backbone of the inhibitor that exists in β-sheet type extended conformation.<br />

<strong>Structure</strong>-<strong>based</strong> design strategies of renin inhibitors have focused on systematic transformation of its<br />

substrate (angiotensinogen). Noteworthy examples which illustrate topographical designed include<br />

inhibitors 96 [152], 97 [153], 98 [154], and 99 [155], of which the latter macrocyclic inhibitor is quite<br />

novel in terms of having two D-aromatic amino acids and lacking a “transition state” bioisostere<br />

replacement at the P 1-P 1' site.<br />

In contrast to renin, the discovery of HIV protease inhibitors provides a high degree of synchronization<br />

of x-ray crystallography studies with iterative structure-<strong>based</strong> drug design efforts as well as the<br />

identification of nonpeptide ligands from mass screening (e.g., coumarins and pyrones) or 3D<br />

computational searching (e.g., haloperidol) to advance what has become a milestone achievement in<br />

rational drug design [113]; and for reviews on HIV protease see Reference 156 and the first chapter of<br />

this book). Representative of the scope of the many contributions to the discovery of both<br />

peptidomimetic and nonpeptide inhibitors of HIV protease (Figure 23) are the de novo designed C 2symmetric<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_598.html (1 of 2) [4/9/2004 1:27:18 AM]


Document<br />

Figure 22<br />

Protease 3D structural models: renin-inhibitor complex and drug design<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_599.html [4/9/2004 1:28:20 AM]<br />

Page 599


Document<br />

Figure 23<br />

Protease structure-<strong>based</strong> drug design: HIV protease-targeted peptidomimetics and nonpeptides<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_600.html [4/9/2004 1:28:56 AM]<br />

Page 600


Document<br />

Page 601<br />

inhibitors 100 [157] and 101 [158]; the nonsymmetric peptidomimetics 44 [50] and 102–106 [159–163,<br />

respectively]; and a series of nonpeptides derived originally from either 3-D computational searching<br />

107 [164] or high-throughput sceening 108–110 [165–168], respectively. Of these compounds, FDA<br />

approval has been recently granted to 102 (Indinivar), 103 (Ritonavir), and 44 (Saquinavir).<br />

Currently, it is believed that there exists well over 150 x-ray crystallographic structures of HIV proteaseinhibitor<br />

complexes, not including mutated forms of the target enzyme that have also been determined to<br />

be important to develop inhibitors which will be effective in so-called HIV resistant strains. The initial<br />

series of x-ray crystallographic structures of HIV protease included the apoprotein [169] and enzymeinhibitor<br />

complexes derived from substrate-<strong>based</strong> analogues having P 1-P 1' substitutions <strong>by</strong><br />

N1eΨ[CH 2NH]N1e [170], LeuΨ[CH(OH)CH 2]Val [171], PheΨ[CH(OH)CH 2N]Pro [172], and<br />

LeuΨ[CH(OH)]Gly or statine [173]. As illustrated in Figure 24, the first reported HIV protease-inhibitor<br />

complex [170] with the pseudopeptide 111 provided a high-resolution map of the active site of the<br />

enzyme as formed in a C 2-symmetric fashion <strong>by</strong> the homodimer, and the “flaps” of each monomeric<br />

subunit (i.e., residues 35–57) were shown to make intermolecular interactions with the backbone of the<br />

inhibitor <strong>by</strong> both direct hydrogen-bonding and through a structural water molecule (W301). Relative to<br />

the C 2-symmetry of the target enzyme, the discovery of C 2-symmetric inhibitors was successfully<br />

achieved <strong>by</strong> the design of PheΨ[CH(OH)]gPhe- and PheΨ[CH(OH)CH(OH)]gPhe-modified<br />

peptidomimetics (gPhe refers to gem-diamino-Phe in which the Cα-CO 2H moiety is replaced <strong>by</strong> Cα-<br />

NH 2) as exemplified <strong>by</strong> 101 [157] Among the plethora of other structure-<strong>based</strong> drug design strategies<br />

focused on HIV protease inhibitor discovery it is also noteworthy to highlight the nonpeptide leads 100<br />

and 108–111 as they displace a key structural water (i.e., W301) and, as opposed to all previously<br />

discovered substrate-<strong>based</strong> inhibitors, are capable of direct hydrogen-bonding interactions to the HIV<br />

protease flap regions (Figure 24). These discoveries provide impetus for molecular design strategies to<br />

consider tightly bound water molecules as possible secondary “ligands” in either de novo design or<br />

iterative structure-<strong>based</strong> design of novel peptidomimetic or nonpeptide lead compounds. Previous<br />

studies [174] on biotinstrepavidin and an x-ray crystallographic structure of the complex showed that<br />

structural water molecule displacement (relative to the apoprotein) <strong>by</strong> key functional groups of the<br />

ligand (biotin) was possible. It is noted, however, that HIV protease is unique from other members of the<br />

aspartyl protease family with respect to the role of structural water W301 role in substrate/ inhibitor<br />

binding. The catalytic water, which is critical to the mechanism of substrate cleavage for all aspartyl<br />

proteases, has been a key feature in the design of a plethora of so-called “transition state” modified<br />

inhibitors that incorporate a tetrahedral<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_601.html [4/9/2004 1:29:03 AM]


Document<br />

Figure 24<br />

Protease 3D structural models: HIV protease-inhibitor complex and drug design<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_602.html [4/9/2004 1:29:47 AM]<br />

Page 602


Document<br />

Page 603<br />

hydroxymethyl substituent within various types of nonhydrolyzable surrogates of the scissile amide<br />

bond (e.g., Ψ[CH(OH)], Ψ[CH(OH)CH 2], and Ψ[CH(OH) CH 2N]; see above). Nevertheless, there exist<br />

examples of highly potent inhibitors of renin e.g., the Ψ[CH 2NH]-modified 41 [47] and macrocylic<br />

peptide 99 [155] and HIV protease e.g., the pyrone-<strong>based</strong> series 108–111 [165–168] respectively) that<br />

do not possess a tetrahedral CH(OH) moiety per se.<br />

Serinyl Proteases<br />

The serinyl proteases include trypsin, chymotrypsin-A, elastase, thrombin, kallikrein, cathepsins-A, G,<br />

and R, Factor VII, Factors IXa-XIIa, and tissue plasminogen activator. High-resolution x-ray<br />

crystallographic structures of this protease family have been determined for thrombin (for a review see<br />

[175]; also refer to Table II [176–182]), Factor Xa [183], trypsin [184], kallikrein-A [185], and elastase<br />

[186–190]. As illustrated in Figure 25, a substrate-<strong>based</strong> inhibitor of thrombin having a boronic acid,<br />

B(OH) 2, substitution for the scissile amide was determine <strong>by</strong> x-ray crystallography to form a covalent<br />

bond to the active site Ser-195 residue [176]. The N-terminal Ac-D-Phe-Pro moiety of this inhibitor<br />

binds in a β-sheet type extended conformation that involves hydrogen-bonding contacts to the enzyme<br />

and well-defined hydrophobic and aromatic-aromatic (edge-to-face) stacking interactions. The inhibitor<br />

Arg side chain binds in an extended conformation and the guanidino moiety forms bidentate hydrogenbonding<br />

interactions with an Asp189 residue at the base of the S 1 “specificity” pocket as well as<br />

additional hydrogen-bonds to the enzyme, one of which is mediated through a structural water. Relative<br />

to the substrate-<strong>based</strong> peptidomimetic inhibitors of thrombin having C-terminal electrophilic groups<br />

(e.g., aldehyde, ketone, and boronic acid), the discovery and structure-<strong>based</strong> design of nonpeptide<br />

inhibitors not having P 1 electrophilic functionalization has also been extremely successful as represented<br />

<strong>by</strong> 52 [70], 60 [71], and 113 [182]. As shown in Figure 25, the design of a highly potent and selective<br />

amidinopiperidine-<strong>based</strong> thrombin inhibitor 113 was derived from analysis of the x-ray crystallographic<br />

structures of thrombin complexed with inhibitors 52 and 60. The latter two compounds showed different<br />

trajectories of their P 1 side chains (i.e., guanidinoalkyl and amidinophenyl, respectively) into the S 1<br />

pocket to account for the observed opposite chirality preferences at the Cα-position of the P 1 amino acid<br />

residues. Also, the C-terminal cycloalkyl moieties of both 52 and 60 were observed to bind to the socalled<br />

inhibitor “P-pocket” (i.e., the P 2 substrate pocket), thus explaining that these compounds were not<br />

binding in a substrate-like conformation such as the peptidomimetic inhibitors Ac-D-Phe-Pro-boroArg-<br />

OH as described above. Thus, the design of the novel amidinopiperidine-<strong>based</strong> inhibitor 113 illustrates a<br />

“transposition” of the P-pocket binding group to an N-substituted Gly-β-Asp scaffold.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_603.html [4/9/2004 1:30:10 AM]


Document<br />

Figure 25<br />

Protease 3D structural models: thrombin-inhibitor complex and drug design<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_604.html (1 of 2) [4/9/2004 1:31:02 AM]<br />

Page 604


Document<br />

Page 605<br />

For other serinyl proteases, such as Factor Xa, kallikrein and elastase, the availability of x-ray<br />

crystallographic structure of the enzymes (apo/complexes) provides further examples in which structure<strong>based</strong><br />

drug design is being advanced (Table 2). In particular, elastase-targeted drug discovery is<br />

highlighted here as it illustrates substrate-<strong>based</strong> peptidomimetic inhibitor design strategies that have<br />

focused on key P 2–P 3 side chain and backbone hydrogen-bonding interactions with the enzyme (for<br />

reviews see Reference [191]). Several x-ray crystallographic structures have been determined for<br />

pancreatic elastase [187,188,189b,190,191] and leukocyte elastase [186,187,189a], including complexes<br />

with peptide substrate-<strong>based</strong> inhibitors having P 1 electrophilic functionalities such as benzoxazole, [188]<br />

trifluoromethyl ketone [56,187,189], and α-,α-difluoro-β-ketoamide [190]. Recently, the design of<br />

peptidomimetic inhibitors incorporating nonpeptidyl P 2-P 3 replacements has resulted in the discovery of<br />

highly potent compounds [56,187]. Specifically, a lead series of highly potent trifluoromethylketone<strong>based</strong><br />

inhibitors of human leukocyte elastase which incorporate a N-carboxymethyl-3-amino-6-arylpyridone<br />

template (50, 116; see Figure 26) was developed and shown <strong>by</strong> x-ray crystallography to<br />

provide backbone hydrogen-bonding and a novel trajectory of a P 2 group from the pyridone ring to the<br />

enzyme. Further modification of the pyridone ring to give the bicyclic pyridopyrimidine derivative 117<br />

was predicted from molecular modeling studies to provide additional hydrogen-bonding to the enzyme<br />

as well as another site on the bicyclic heteroaromatic ring system for tethering various hydrophobic or<br />

hydrophilic groups. Finally, a series of novel dipeptide-<strong>based</strong> inhibitors (e.g., trifluoromethylacetyl-Leu-<br />

Phe-p-isoproylanilide and a peptidomimetic derivative) are particularly intriguing because they bind<br />

“backwards” as <strong>based</strong> on analysis of the x-ray crystallography structures of their complexes with<br />

pancreatic elastase [191]. In this binding mode the trifluoromethylacetyl moiety is proximate to the<br />

active site Ser residue and the ligand backbone and side chain substructures make hydrogen bonding and<br />

hydrophobic contacts with the enzyme, respectively.<br />

Cysteinyl Proteases<br />

The cysteinyl proteases include papain; calpains I and II; cathepsins B, H, and L; proline endopeptidase;<br />

and interleukin-converting enzyme (ICE) and its homologs. The most well-studied cysteinyl protease is<br />

likely papain, and the first x-ray crystallographic structures of papain [193] and a peptide<br />

chloromethylketone inhibitor-papain complex [194] provided the first high resolution molecular maps of<br />

the active site. Pioneering studies in the discovery of papain substrate peptide-<strong>based</strong> inhibitors having P 1<br />

electrophilic moieties such as aldehydes [195], ketones (e.g., fluoromethylketone, which has been<br />

determined [196] to exhibit selectivity for cysteinyl proteases versus serinyl proteases), semicarbazones,<br />

and nitriles are noteworthy since 13C-NMR spectro-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_605.html [4/9/2004 1:31:09 AM]


Document<br />

Figure 26<br />

Protease 3D structural models: elastase-inhibitor complex and drug design<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_606.html (1 of 2) [4/9/2004 1:32:04 AM]<br />

Page 606


Document<br />

Page 607<br />

scopic studies provided direct evidence that the active site Cys25 of papain forms a reversible covalent<br />

bond with such electrophiles (see review Reference 142). The x-ray crystallographic structures of<br />

cathepsin-B [197] and picornaviral 3C protease [198] have also been determined, but only in the<br />

apoprotein form. Recently, high-resolution x-ray crystallographic structures of ICE-inhibitor complexes<br />

have been determined [199,200]. This enzyme is structurally unique relative to other cysteinyl proteases<br />

(for a review see Reference 201) in that it has a heterodimeric architecture in which two subunits form<br />

the catalytically active enzyme site (actually, two p10/p20 heterodimers apparently create a tetrameric<br />

form of the competent protease). As shown in Figure 27, an x-ray crystallographic structure of ICE<br />

complexed with a substrate peptide-<strong>based</strong> chloromethylketone inhibitor (118 [200]) that is irreversibly<br />

bonded to the active site Cys285 shows the P 1 Asp specificity pocket to be comprised of two Arg<br />

residues that lie at the base of the S 1 binding pocket. Relative to other side chain binding pockets, a<br />

hydrophobic “channel” type S 4 site exists for the P 4 Tyr of the inhibitor, whereas the P 3-P 2 Val-Ala side<br />

chains are well exposed to solvent. With respect to the peptide backbone of the inhibitor, hydrogen<br />

bonding interactions between the P 3 Val (both NH and CO) and the P 1 Asp (NH) and the P 10 monomer<br />

are predicted.<br />

The x-ray crystallographic structure of ICE complexed with the inhibitor Ac-Tyr-Val-Ala-Asp-aldehyde<br />

(119 [202]), supports structure-activity studies [202] that employed a systematic analysis of N-methylamino<br />

acid substitutions in which hydrogen bonding interactions between inhibitor and enzyme tolerated<br />

only N-Me-Ala replacement at the P 2 site. Furthermore, C-terminal modification of P 1 Asp <strong>by</strong><br />

irreversible alkylating groups, such as the aryloxymethyl ketone analog 120 [203], have led to the first<br />

reported peptidomimetic inhibitor of ICE (121 [204]). Noteworthy in the structure of this<br />

peptidomimetic inhibitor is that a pyridone template provided an effective P 2-P 3 replacement, similar to<br />

that for the elastase inhibitor design (see above). It is likely that the backbone hydrogen bonding<br />

network between 121 and ICE is conserved as compared to the x-ray crystallographic structure of the<br />

substrate peptide-<strong>based</strong> inhibitor 118 complexed to the target enzyme. Finally, the x-ray crystallographic<br />

structure of the ICE homolog referred to as apopain, or CPP32, as a complex with a peptide-aldehyde<br />

inhibitor has been recently determined [205] and provides additional insight as the specificity of<br />

substrate recognition at both the S 1 (P 1 Asp) and S 4 (P 4 Asp) subsites. Thus, the opportunity for iterative<br />

structure-<strong>based</strong> drug design exists to advance novel peptidomimetic and nonpeptide inhibitors of ICE<br />

and/or its homologs.<br />

Metalloproteases<br />

The metalloproteases include both exopeptidases (e.g., angiotensin-converting enzyme, aminopeptidase-<br />

M, and carboxypeptidase-A) and endopeptidases (e.g.,<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_607.html [4/9/2004 1:32:09 AM]


Document<br />

Figure 27<br />

Protease 3D structural models: ICE-inhibitor complex and drug design<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_608.html (1 of 2) [4/9/2004 1:32:32 AM]<br />

Page 608


Document<br />

Page 609<br />

(e.g., thermolysin, endopeptidase 24.11 or NEP, collagenase, gelatinase and stromelysin). Historically,<br />

the most well-studied metalloprotease is thermolysin (for a review see [206]), and the first x-ray<br />

crystallographic structures of thermolysin [207] and several structurally distinct peptide inhibitors (see<br />

Table 2 [208–212]) provided the first high-resolution molecular maps of the active site and insight into<br />

the mechanistic roles of the metal ion for substrate hydrolysis. Specifically, the binding interactions of<br />

P 1–P 1' “transition state” amide bond isosteres (e.g., ψ[P=O(OH)NH]) as well as metal-chelating<br />

functionalities (e.g., hydroxamates, carboxylates, and sulfhydryls), introduced as N-substitutions on<br />

P 1–'P 2' peptide scaffolds [208–212] have been determined. Moreover, the structural and mechanistic<br />

information derived from studies on thermolysin have provided insight into the design of inhibitors of<br />

the therapeutically relevant target, angiotensin-converting enzyme or ACE (for reviews see Reference<br />

142). This has been particularly significant since the 3D structure of ACE has not yet been determined.<br />

However, it is noted that x-ray crystallographic structures of carboxypeptidase, a related metalloprotease<br />

of the exopeptidase group, for both the apoprotein and a Glyψ[P=O(OH)NH]Phe-modified inhibitor<br />

complex have been reported [213]. As illustrated in Figure 28, the x-ray crystallographic structure of the<br />

Pheψ[P=O(OH)NH]Leu-modified peptidomimetic inhibitor 122 complexed with thermolysin show the<br />

Zn 2+ coordination and P 1'-P 2' (or P 1-P 2' “collected product”) mode of binding. Relative to this structure<br />

the predicted molecular interactions between ACE and its inhibitors (e.g. Captopril [51], Enalaprilat<br />

[52], and Fosfinoprilat [214] as well as an emerging class of “dual specific” inhibitors of ACE and NEP<br />

(e.g., 123 [215]) may be envisaged.<br />

The ability to design specific inhibitors of several matrix metalloproteases (MMPs) is rapidly<br />

developing (for reviews see Reference 216). Such MMP targets include fibroblast collagenase (MMP-1),<br />

gelatinase-A (MMP-2), and stromelysin (MMP-3). Both x-ray crystallography and NMR spectroscopy<br />

have provided 3D structural information for several MMPs as well as MMP-inhibitor complexes (see<br />

Table 2; [217–226]). In the example of collagenase, “first generation” substrate-<strong>based</strong> inhibitor design<br />

strategies have focused on modifying the P 1–P 1' cleavage site (e.g., Gly-Phe, Ala-Tyr, and Ala-Phe) <strong>by</strong><br />

N-terminal functionalities capable of Zn 2+ coordination (e.g., sulfhydryl, carboxyl, phosphonoalkyl, and<br />

hydroxamate [227–229]). As shown in Figure 29, an x-ray crystallographic structure of the Nhydroxamate-modified<br />

peptide 124 complexed to fibroblast collagenase provides a molecular map of<br />

both its hydrogen-bonding interactions to the enzyme active site and binding to bound Zn 2+ [219].<br />

Potent MMP-1 inhibitors have been designed that tether the P 2' side chain to the inhibitor's C-terminus<br />

as macrocylic rings [53,230]. In the example of gelatinase-A, potent inhibitors have been designed (e.g.,<br />

48, Figure 11 [54]) <strong>by</strong> N-terminal hydroxamate and P 1' extended aromatic side chain modifications<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_609.html [4/9/2004 1:32:48 AM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_610.html (1 of 2) [4/9/2004 1:33:29 AM]<br />

Page 610


Document<br />

Figure 28<br />

Protease 3D structural models: thermolysin-inhibitor complex ACE inhibitor drug design<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_610.html (2 of 2) [4/9/2004 1:33:29 AM]


Document<br />

Page 611<br />

of a P 1'–P 3' peptide scaffold [54]. To date, no 3D structural information is available for MMP-2 with<br />

respect to either the apoprotein catalytic domain or inhibitor complexes thereof. Finally, in the example<br />

of stromelysin-1, potent inhibitors have been designed (e.g., 49, Figure 11 [55]) <strong>by</strong> N-terminal<br />

carboxyalkylamino functionalization that includes a P 1 substituent. An x-ray crystallographic structure<br />

of a related MMP-3 inhibitor 125 shows (Figure 29) the hydrogen-bonding interactions at the active site<br />

and carboxylate coordination with the Zn 2+ [223]. Finally, a recently determined x-ray crystallographic<br />

structure of an Pheψ[P=O(OH)CH 2]Ala-modified peptide inhibitor 126 complexed with astacin (Figure<br />

29) shows the extensive hydrogen-bonding network between inhibitor, enzyme, Zn 2+, and a structural<br />

water [226]. It is expected that iterative structure-<strong>based</strong> design of inhibitors of the MMP family will<br />

enable the discovery of novel compounds with superior binding affinities and/or selectivities.<br />

C. Signal-Transduction Protein Targets<br />

Beyond proteases the opportunity for structure-<strong>based</strong> drug design is being realized in the rapidly<br />

developing area of signal-transduction research (e.g., intracellular protein and nucleic acid targets). Both<br />

x-ray crystallography and/or NMR spectroscopy have significantly contributed to a wide-scope database<br />

of 3D structural information for various catalytic and noncatalytic signal-transduction protein targets<br />

(see Table 3). These include tyrosine kinases (e.g., growth factor receptor kinases and Src family<br />

kinases; for reviews see Reference 231), serine/threonine and “dual specificity” kinases (e.g., mitogenactivated<br />

protein kinases and CDK2 and cAMP-dependent protein kinases; for reviews see Reference<br />

232), phosphotyrosine phosphatases (e.g., PTP1B and Syp; for reviews see Reference 233),<br />

phosphoserine/phosphothreonine and “dual specificity” phosphatases (e.g., VH1 and CDC25; for<br />

reviews see Reference 234), noncatalytic “adapter” proteins (e.g., Crk, Grb2, Shc, and IRS-1; for<br />

reviews see Reference 85), transferases (e.g., Ras farnesyl transferase; for reviews see Reference 81),<br />

proline cis-trans isomerases (e.g., FKBP-12 and cyclophilin A; for reviews see Reference 235), and GTPbinding<br />

proteins (e.g., p21 Ras and α-β/γ heterotrimeric G-protein for GPCR superfamily; for reviews<br />

see References 236 and 237, respectively). The diversity of targets, mechanistic relationships (e.g.,<br />

enzyme–substrate or regulatory protein–protein interaction), and potential therapeutic opportunities has<br />

created great impetus for focused research in the area of signal transduction (for reviews see References<br />

265–267).<br />

Src Homology-2 and Homology-3 Domains<br />

The identification of noncatalytic regulatory domains referred to as Src homology (SH) domains has<br />

been rapidly advanced over recent years as a critical link<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_611.html [4/9/2004 1:38:43 AM]


Document<br />

Figure 29<br />

Protease 3D structural models: matrix metallo protease-inhibitor complexes<br />

and drug design<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_612.html (1 of 2) [4/9/2004 1:39:38 AM]<br />

Page 612


Document<br />

Table 3 Some Known 3D <strong>Structure</strong>s of Signal-Transduction Proteins (Apo/Complexes)<br />

Protein Target Apo/Complex Resolution Reference<br />

Src homology domains<br />

Abl SH2 —(Apo) NMR 238<br />

Src SH2 —(Apo) 2.5 Å 239<br />

phosphopeptide 2.7 Å 239<br />

phosphopeptide NMR 240<br />

PLC (C) SH2 phosphopeptide NMR 241<br />

Shc SH2 phosphopeptide NMR 242<br />

Lck SH2 phosphopeptide 2.2 Å 243a<br />

p85 (N) SH2 phosphopeptide 2.0 Å 244<br />

Syp SH2 phosphopeptide 2.0 Å 245<br />

Grb2 SH2 phosphopeptide 2.1 Å 246<br />

Syk SH2 phosphopeptide NMR 247<br />

Zap70 SH2-SH2 phosphopeptide (tandem) 1.9 Å 248<br />

Src SH3 —(Apo) NMR 249a<br />

peptide NMR 249b<br />

Abl SH3 peptide 2.0 Å 250<br />

Crk SH3 peptide 1.5 Å 251<br />

Grb2 SH3 peptide NMR 252<br />

Grb2 SH3-SH2-SH3 —(Apo) 3.1 Å 253<br />

Tyr and Ser/Thr kinases<br />

Insulin receptor Tyr kinase —(Apo) 2.1 Å 254<br />

cAMP-dep. protein kinase —(Apo) 3.9 Å 255<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_613.html (1 of 2) [4/9/2004 1:39:56 AM]<br />

Page 613


Document<br />

(Ser/Thr) peptide 2.9 Å 255<br />

Cell cycle-dep. protein —(Apo) 2.4 Å 256a<br />

kinase (CDK2) protein (p27 Kip1 inhibitory domain) 2.3 Å 256b<br />

Mitogen-activated protein kinase<br />

(MAPK)<br />

Tyr and Ser/Thr phosphatases<br />

—(Apo) 2.3 Å 257<br />

BH-PTP (Tyr) —(Apo) 2.2 Å 258a<br />

PTP1B (Tyr) —(Apo) 2.8 Å 258b<br />

peptide (C215S mutant enzyme) 2.6 Å 258c<br />

VH1 (Tyr and Ser/Thr”) —(Apo) 2.1 Å 259<br />

PP-1 (Ser/Thr) —(Apo) 2.1 Å 260<br />

pTyr binding domains<br />

IRS-1 PTB —(Apo) 1.95 Å 261<br />

peptide 1.8 Å 261<br />

IRS-1 PTB peptide NMR 262<br />

Pro cis-trans isomerases<br />

Cyclophilin peptide (cyclosporin-A) 2.8 Å 263<br />

FKBP-12 FK-506 (macrolide antibiotic) 1.7 Å 264<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_613.html (2 of 2) [4/9/2004 1:39:56 AM]


Document<br />

Page 614<br />

in deconvoluting both enzyme-substrate and regulatory protein-protein interactions for a number of<br />

signal-transduction pathways (for reviews see Reference 268). This emerging “superfamily” of proteins<br />

includes SH2 and SH3 domains, the so-called “choreographers of multiple signalling pathways,” and<br />

include very intriguing new therapeutic targets [85]. The SH2 domains have been determined to bind<br />

cognate phosphotyrosine (pTyr) containing proteins in a sequence-dependent manner relative to the<br />

amino acids contiguous to the C-terminal side of the pTyr residue (e.g., for Src SH2 a preferred<br />

sequence is ~pTyr-Glu-Glu-Ile~ versus ~pTyr-Tyr-Asn-Tyr for Grb2 [87,269]. The SH3 domains have<br />

been determined to specifically bind Pro-rich sequences of cognate proteins. Interestingly, as a result of<br />

the pseudosymmetrical nature of the SH3 domains there is the possibility of binding both N rarrow.gif<br />

C and C rarrow.gif N directions (e.g., for Src SH3 preferred sequences are ~Arg-Ala-Leu-Pro-Pro-Leu-<br />

Pro-Arg-Tyr and Ala-Phe-Ala-Pro-Pro-Leu-Pro-Arg-Arg, wherein Arg binds to a site 3 pocket [249b]).<br />

With respect to SH2 domain structure-<strong>based</strong> drug design, the first x-ray crystallographic structures of<br />

pTyr-containing peptide ligands complexed with Src SH2 domain [239] have been utilized to design the<br />

first peptidomimetic antagonists [89]. As illustrated in Figure 30, a molecular map of the tetrapeptide<br />

sequence ~pTyr-Glu-Glu-Ile~ complexed with Src SH2 [239] shows the pTyr binding pocket and a<br />

second binding site for the P +3 Ile residue. As previously described, a prototypic peptidomimetic AcpTyr-Glu-D-Hcy-NH<br />

2 (66) was first discovered <strong>by</strong> a peptide scaffold design strategy (see Figure 16;<br />

[77]) that took into account the x-ray crystallographic structure of Src SH2-phosphopeptide (Glu-Pro-<br />

Gln-pTyr-Glu-Glu-Ile-Pro-Ile-Tyr-Leu, 127) complex. Further structure-<strong>based</strong> drug design<br />

modifications have led to the discovery a series of potent peptidomimetics having novel C-terminal<br />

functionalization (e.g., “transposed” side chain of the P +1 Glu or conformational constraint using a<br />

pyrrolidine ring; see Figure 30) as represented <strong>by</strong> 128–130 [89,270]. Studies focused on Src SH2 [88]<br />

have shown that the phosphate ester of pTyr is particularly critical for molecular recognition, and that<br />

significant loss in binding occurs <strong>by</strong> replacement with sulfate, carboxylate, nitrosyl, hydroxy, and<br />

amino. However, backbone modifications of pTyr which replace its acylated amino functionality with<br />

aromatic rings designed to form π-cation type interactions with the Arg-αA2 were effective substitutions<br />

[271].<br />

Recently, high-resolution 3D structures have been described for the noncatalytic “adapter” protein Grb2<br />

with respect to the apoprotein (SH3-SH2-SH3 [153]) as well as the individual SH2 and SH3 domains<br />

[246 and 252, respectively]. In the case of the SH2 domain of Grb2 an x-ray crystallographic structure<br />

of a phosphopeptide complex provided insight to the molecular basis of the specificity of Grb2 SH2<br />

binding of ~pTyr-Xxx-Asn-Yyy~ sequences. As illustrated in Figure 31, the binding interactions of Lys-<br />

Pro-Phe-pTyr-Val-Asn-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_614.html [4/9/2004 1:40:12 AM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_615.html (1 of 2) [4/9/2004 1:41:29 AM]<br />

Page 615


Document<br />

Figure 30<br />

SH2 domain 3D structural models: pp60Src SH2 domain antagonists<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_615.html (2 of 2) [4/9/2004 1:41:29 AM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_616.html (1 of 2) [4/9/2004 1:42:26 AM]<br />

Page 616


Document<br />

Figure 31<br />

SH2 domain 3D structural models: Grb2 SH2-SH3 domain antagonists<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_616.html (2 of 2) [4/9/2004 1:42:26 AM]


Document<br />

Page 617<br />

Val (131) showed that the phophopeptide adopts a β-turn conformation about the P-P +3 residues and that<br />

the P +2 Asn side chain carboxamide moiety is extensively hydrogen bonded to the protein. In contrast to<br />

the well-defined binding pocket for the P +3 Ile of 127 to bind Src SH2, the P +3 Val of 131 engages in<br />

limited surface hydrophobic interactions because the Trp121 residue of Grb2 SH2 sterically blocks the<br />

phosphopeptide from attaining a similar binding mode. In the case of the SH3 domain, the binding of a<br />

cognate Pro-rich peptide sequence ~Pro-Pro-Pro-Val-Pro-Pro-Arg-Arg~ shows distinct pockets which<br />

recognize the Pro, Val, and Arg residues as illustrated in Figure 31. The peptide adopts a left-handed<br />

polyproline type-II helical conformation which projects the three aforementioned residues to their<br />

complementary binding pockets in the Grb2 SH3 domain.<br />

Tyrosine Phosphatases and Phosphotyrosine Binding Domains<br />

Two other types of signal-transduction proteins that recognize phosphotyrosine-containing sequences<br />

are tyrosine phosphatases (e.g., PTP1B, Syp, and CD45) and proteins that contain a noncatalytic motif<br />

referred to as a phosphotyrosine binding (PTB) domain. Although tyrosine phosphatases and PTB<br />

domains are structurally quite different from each other they both are similar with respect to the binding<br />

of pTyr-containing sequences with preference to the amino acids N-terminal to the pTyr residue. In<br />

other words, the tyrosine phosphatase PTP1 binds EGF receptor (pTyr 992)-<strong>based</strong> phosphopeptide<br />

substrates Asp-Ala-Asp-Glu-pTyr-Leu-NH 2 and Asp-Ala-Asp-Glu-pTyr-Leu-Ile-Pro-Gln-Gln-Gly<br />

equally [272], and the substitution of pTyr <strong>by</strong> nonhydrolyzable F 2Pmp in the hexapeptide derivative has<br />

been reported [273] to give a highly potent inhibitor (see compound 69, Figure 14). In the case of PTB<br />

domains, ~Asn-Pro-Xxx-pTyr~ (where Xxx is variable) has been determined [274] as the cognate<br />

sequence for several PTB-domain-containing proteins such as Shc and the insulin receptor substrate-1<br />

(IRS-1). The preference for amino acids N-terminal to the pTyr residue in binding to either tyrosine<br />

phosphatases or PTB domains is, therefore, opposite of that known for SH2 domains [275].<br />

Among the tyrosine phosphatase superfamily, PTP1B was the first to be discovered and structurallydetermined<br />

<strong>by</strong> x-ray crystallography, as the apoprotein catalytic domain [258b]. The first x-ray<br />

crystallographic structure of PTP1B complexed with a phosphopeptide has also been very recently<br />

determined [258c] using a catalytically inactive Cys215 rarrow.gif Ser PTP1B mutant and the<br />

phosphopeptide Asp-Ala-Asp-Glu-pTyr-Leu-NH 2 (133). As illustrated in Figure 32, the molecular<br />

interactions between the tyrosine phosphatase and the phosphopeptide are dominated <strong>by</strong> electrostatic<br />

(i.e., the pTyr, P -1 Glu, and P -2 Asp residues) and hydrogen bonding contacts to key amide functionalities<br />

of the backbone of 133. The P +1 Leu side chain forms hydrophobic contacts with<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_617.html [4/9/2004 1:42:40 AM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_618.html (1 of 2) [4/9/2004 1:43:11 AM]<br />

Page 618


Document<br />

Figure 32<br />

PTP and PTB 3D structural models: PTP1B inhibitor and PTB antagonists<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_618.html (2 of 2) [4/9/2004 1:43:11 AM]


Document<br />

several PTP1B residues at the surface proximate to the well-defined pTyr binding pocket. Such 3D<br />

structural information provides insight for the design of peptidomimetic inhibitors of PTP1B.<br />

Page 619<br />

In the case of the IRS-1 PTB domain, x-ray crystallographic studies of PTB complexed with a pTyrcontaining<br />

peptide (134) complex have shown (Figure 32) that the phosphopeptide forms a type-I β-turn<br />

within the Asn-Pro-Ala-pTyr sequence, and the peptide backbone is extensively hydrogen bonded to the<br />

PTB domain from the P +1-P -7 residues of 134 [261]. The pTyr binding pocket provides both electrostatic<br />

and multiple hydrogen bonding contacts to the phosphate ester moiety, and hydrophobic interactions<br />

exist for the P -1-P -3 side chains of the peptide. As in the case of PTP1B, such 3D structural information<br />

provides the opportunity for structure-<strong>based</strong> drug design to discover potent inhibitors which may be used<br />

for further exploration in cellular studies.<br />

V. Future Perspectives<br />

The impact of structure-<strong>based</strong> drug design on both peptidomimetic and nonpeptide drug discovery has<br />

been significant over the past few years. The integration of sophisticated computational chemical<br />

technologies, structural biology (x-ray crystallography and NMR spectroscopy), molecular diversity and<br />

high-through-put screening, and targeted biological testing are expected to provide invaluable guidance<br />

to drug discovery. These “technological tools” are contributing to an emerging “3D structure-activity<br />

database” that is the essence of rational drug design. Certainly, this intriguing area of peptidomimetic<br />

and nonpeptide drug discovery and design is providing tremendous insight to our understanding of<br />

molecular recognition and biochemical mechanisms. With particular regard to molecular diversity, the<br />

generation of new leads from combinatorial chemistry focused on synthetic peptide, peptidomimetic, or<br />

nonpeptide-type libraries will provide new opportunities (and challenges) for structure-<strong>based</strong> drug<br />

design strategies (for reviews see Reference 276). Novel scaffolds and templates will continue to be<br />

advanced and iteratively modified using “randomized” or “targeted” substructural replacements, and<br />

such work is exemplified in this review. Of particular significance to the field of peptidomimetic and<br />

nonpeptide drug discovery will be the rational use of D-amino acids, Nα-alkyl or Cα-alkyl amino acids,<br />

N-substituted Gly, XxxΨ[Z]Yyy (dipeptide isosteres), benzodiazepines, and amino-benzoic acids in<br />

structure-<strong>based</strong> drug design. Without question, structure-<strong>based</strong> drug design will be a decisive factor in<br />

drug discovery efforts ranging from de novo design to iterative structural optimization of<br />

peptidomimetic and nonpeptide lead compounds.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_619.html [4/9/2004 1:43:17 AM]


Document<br />

Acknowledgments<br />

Page 620<br />

I wish to acknowledge my colleagues at Parke-Davis for their critical review of this manuscript as well<br />

as for their collaborative contributions to structure-<strong>based</strong> drug design research in the several areas,<br />

including HIV protease inhibitor discovery, Ras farnesyl transferase inhibitor discovery, Src homology-<br />

2 domain antagonist discovery, interleukin-converting enzyme inhibitor discovery, and melanocortin<br />

receptor modeling development to probe MSH agonist and antagonists binding. I especially thank Mark<br />

Plummer, Elizabeth Lunney, Charles Stankovic, Kim Para, Aurash Shahripour, Vara Prasad, Carrie<br />

Haskell-Luevano, Daniel Ortwine, Daniele Leonard, Wayne Cody, and Christine Humblet in this regard.<br />

I also very much thank Pandi Veerapandian for the opportunity to contribute a chapter to this book, and<br />

for his critical review of this manuscript and excellent editorial suggestions.<br />

References<br />

1. (a) Hru<strong>by</strong> VJ, Al-Obeidi F, Kazmierski W, Biochem J 1990; 268:249–262; (b) Hru<strong>by</strong> VJ, Pettit BM.<br />

In: Perun TJ, Propst CL, Eds. Computer-Aided drug design, Method and Application. New York:<br />

Marcel Dekker, 1989:405–461.<br />

2. Fauchere, J-L In: Advances in <strong>Drug</strong> Research. Vol. 15. Testa B Ed. London: Academic Press,<br />

1986:29–69.<br />

3. Ward DJ. Peptide Pharmaceuticals-Approaches to the <strong>Design</strong> of Novel <strong>Drug</strong>s. Buckingham, England:<br />

Open University Press, (1991).<br />

5. Toniolo C. Int J Peptide Protein Res 1990; 35:287–300.<br />

4. DeGrado WF. Adv Protein Chem 1988; 39:51–124.<br />

6. Goodman M, Ro S. In: Medical Chemistry and <strong>Drug</strong> <strong>Design</strong>. Vol. I. Principles of <strong>Drug</strong> Discovery.<br />

5th ed. Wolff ME, Ed. 1994:803–861.<br />

7. Kessler H, Haupt A, Will M, In: Computer-Aided <strong>Drug</strong> <strong>Design</strong>, Method and Application (Perun TJ,<br />

Propst CL, Eds. New York: Marcel Dekker, 1989:461–484.<br />

8. (a) Marshall GR. Tetrahedron 1993; 49:3547–3558; (b) Humblet C, Marshall GR. Ann Report Med<br />

Chem 1980; 267–276.<br />

9. Kahn M. Tetrahedron Symposium-In-Print Peptide Secondary <strong>Structure</strong> Mimetics. 1993:50.<br />

10. McDowell RS, Artis DR. Ann Rep Med Chem 1995; 30:265–274.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_620.html (1 of 2) [4/9/2004 1:43:44 AM]


Document<br />

11. (a)<br />

Freidinger<br />

RM.<br />

Progress<br />

<strong>Drug</strong> Res<br />

1993;<br />

40:33–98.<br />

(b)<br />

Freidinger,<br />

RM. Trends<br />

Pharm Sci<br />

1989;<br />

10:270–274.<br />

12. Morgan BA, Gainor JA. Ann Rep Med Chem 1989; 24:243–252.<br />

13. Olson GL, Bolin DR, Bonner MP, Bos M, Cook CM, Fry DC, Graves BJ, Hatada M, Hill DE, Kahn<br />

M, Madison VS, Rusiecki VK, Sarubu R, Sepinwall J, Vincent GP, Voss ME. J Med Chem 1993;<br />

36:3039–3049.<br />

14. Fairlie DP, Abbenante G, March DR. Current Med Chem 1995; 2:654–686.<br />

15. Wiley RA, Rich DH. Med Res Rev 1993; 13:327–384.<br />

16. Giannis A, Kolter T. Angew Chem Int Ed Engl 1993; 32:1244–1267.<br />

17. Gante J. Angew Chem Int Ed Engl 1994; 33:1699–1720.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_620.html (2 of 2) [4/9/2004 1:43:44 AM]


Document<br />

18. Sawyer TK. In: Peptide-Based <strong>Drug</strong> <strong>Design</strong>: Controlling Transport and Metabolism Taylor MD,<br />

Amidon GL, Eds. Washington D.C.: ACS Professional Books, American Chemical Society,<br />

1995:387–422.<br />

Page 621<br />

19. (a) Shimohigashi Y, Stammer CH. Int J Peptide Protein Res 1982; 20:199–206; (b) Kaur P, Uma K,<br />

Balaram P, Chauhan VS. Int J Peptide Protein Res 1989; 33:103–109; (c) Gupta A, Chauhan VS. Int J<br />

Peptide Protein Res 1993; 41:421–426.<br />

20. Haghihara M, Anthony NJ, Stout TJ, Clardy J, Schreiber SL. J Amer Chem Soc 1992;<br />

114:6568–6570.<br />

21. Simon RJ, Kania RS, Zuckermann RN, Huebner VD, Jewell DA, Banville S, Ng S, Wang L,<br />

Rosenberg S, Marlowe CK, Spellmeyer DC, Tan R, Frankel AD, Santi DV, Cohen FE, Bartlett PA. Proc<br />

Natl Acad Sci USA 1992; 89:9367–9371.<br />

22. Freidinger RM, Veber DF, Perlow DS, Brooks JR, Saperstein R. Science 1980; 210:656–658.<br />

23. Nagai U, Sato K. Tet Lett 1985; 26:647–650.<br />

24. Feigl M. J Amer Chem Soc 1986; 108:181–182.<br />

25. Kahn M, Wilke S, Chen B, Fujita K. J Amer Chem Soc 1988; 110:1638–1639.<br />

26. Kemp DS, Stites WE. Tet Lett 1988; 29:5057–5060.<br />

27. Genin MJ, Johnson RL. J Amer Chem Soc 1992; 114:8316–8318.<br />

28. Kahn M, Wilke S, Chen B, Fujita K, Lee YH, Johnson ME. J Mol Recognition 1988; 1(2):75–79.<br />

29. Ripka WC, DeLucca GV, Bach II AC, Pottorf RS, Blaney JM. Tetrahedron 1993; 49:3593–3608.<br />

30. Ripka WC, DeLucca GV, Bach II AC, Pottorf RS, Blaney JM. Tetrahedron 1993; 49:3609–3628.<br />

31. Callahan JF, Newlander KA, Burgess JL, Eggleston DS, Nichols A, Wong A, Huffman WF.<br />

Tetrahedron 1993; 49:3479–3488.<br />

32. (a) Smith AB, Keenan TP, Holcomb RC, Sprengeler PA, Guzman MC, Wood JL, Caroll PJ,<br />

Hirschmann RS. J Amer Chem Soc 1992; 114:10672–10674; (b) Smith AB, Knight SD, Sprengler PA,<br />

Hirschmann Bioorg Med Chem 1996; 4:1021–1034.<br />

33. Cascieri MA, Fond TM, Strader CD. <strong>Drug</strong>s of the Future 1996; 21:521–527.<br />

34. Pitzele BS, Hamilton RW, Kudla KD, Taymbalov S, Stapefield A, Savage MA, Clare M, Hammond<br />

DL, Hansen R DW. J Med Chem 1994; 37:888–896.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_621.html (1 of 2) [4/9/2004 1:44:01 AM]


Document<br />

35. Zablocki JA, Miyano M, Garland B, Pireh D, Schretzman L, Rao SN, Lindmark RJ, Panzer-Knodle<br />

S, Nicholson N, Taite B, Salyers A, King L, Feigen L. J Med Chem 1993; 36:1811–1819.<br />

36. Samanen J, Ali FE, Barton L, Bondinell W, Burgess J, Callahan J, Calvo R, Chen W, Chen L,<br />

Erhard K, Heyes R, Hwang S-M, Jakas D, Keenan R, Ku T, Kwon C, Lee C-P, Miller W, Newlander K,<br />

Nichols A, Peishoff C, Rhodes G, Ross S, Shu A, Simpson R, Takata D, Yellin TO, Uzsinskas I,<br />

Venslavasky J, Wong A, Yuan C-K, Huffman W. In: Kaumaya PTP Hodges, RS, Eds. Peptide<br />

Chemistry, <strong>Structure</strong> and Biology Mayflower Scientific Ltd., 1996; 679–681.<br />

37. Flynn DL, Villamil CI, Becker DP, Gullikson GW, Moummi C, Yang D-C. Bioorg Med Chem Lett<br />

1992; 2:1251–1256.<br />

38. Horwell DC. Neuropeptides 1991; 19 (Suppl.):57–64.<br />

39. (a) Ishikawa K, Fukami T, Nagase T, Mase T, Hayama T, Niiyama K, Fujita K, Urakawa Y,<br />

Kumagai U, Fukuroda T, Ihara M, Yano M, Eur Peptide Symp, abs. 1992; 27; (b) Ishikawa K, Ihara M,<br />

Noguchi K, Mase T, Mino N, Saki T,<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_621.html (2 of 2) [4/9/2004 1:44:01 AM]


Document<br />

Page 622<br />

Fukuroda T, Fukami T, Ozaki S, Nagase T, Nishikebe M, Yano M. Proc Natl Acad Sci USA 1994;<br />

91:4892–4896.<br />

40. McDowell<br />

RS, Elias KA,<br />

Stanley MS,<br />

Burdick DJ,<br />

Burnier JP,<br />

Chan KS,<br />

Fairbrother<br />

WJ,<br />

Hammonds<br />

RG, Ingle GS,<br />

Jacobsen NE,<br />

Mortensen<br />

DL, Rawson<br />

TE, Won WB,<br />

Clark RG,<br />

Somers TC.<br />

Proc Natl<br />

Acad Sci USA<br />

1995;<br />

92:1165–1169.<br />

41. (a) Nicolaou KC, Salvino JM, Raynor K, Pietranico S, Reisine T, Freidinger R, Hirschmann R. In:<br />

Rivier JE, Marshall GR, Eds. Peptides-Chemistry, <strong>Structure</strong> and Biology. Leiden, Netherlands: Escom,<br />

1990:881–884. (b) Hirschmann R, Nicolaou KC, Pietranico S, Salvino J, Leahy EM, Sprengeler PA,<br />

Furst G, Smith III AB, Strader CD, Cascieri MA, Candelore MR, Donaldson C, Vale W, Maechler L. J<br />

Amer Chem Soc 1992; 114:9217–9218.<br />

42. Hagiwara D, Miyake H, Igarni N. Reg Peptides 1992; 1 (Suppl):66.<br />

43. Smith PW, McElroy AB, Pritchard JM, Deal MJ, Ewan GB, Hagen RM, Ireland SJ, Ball D,<br />

Beresford I, Sheldrick R, Jordan CC, Ward P. Bioorg Med Chem Lett 1993; 3:931–935.<br />

44. Boden P, Eden JM, Hodgson J, Horwell DC, Hughes J, McKnight AT, Lewthwaite RA, Prichard<br />

MC, Raphy J, Meecham K, Ratcliffe, GS, Suman-Chauhan N, Woodruff GN. J Med Chem 1996;<br />

39:1664–1665.<br />

45. Zablocki JA, Rao SN, Baron DA, Flynn DL, Nicholson NS, Feigen LP. Current Pharm <strong>Design</strong> 1996;<br />

1:533–558.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_622.html (1 of 2) [4/9/2004 1:44:04 AM]


Document<br />

46. Hirschmann R, Yao W, Cascieri MA, Strader CD, Maechler L, Cichy-Knight MA, Hynes J Jr, van<br />

Rijn RD, Sprengeler PA, Smith AB. J Med Chem 1996; 39:2441–2448.<br />

47. Sawyer TK, Maggiora LL, Liu L, Staples DJ, Bradford VS, Mao B, Pals DT, Dunn BM, Poorman R,<br />

Hinzmann J, DeVaux AE, Affholter JA, Smith CW. In: Marshall, GR, Rivier J, Eds. Peptides:<br />

Chemistry and Biology Ae Leiden: Escom Science Publishers, 1990; 855–857.<br />

48. Rosenberg SH, Spina KP, Condon SL, Polakowski J, Yao Z, Kovar P, Stein HH, Cohen J, Barlow<br />

JL, Klinghofer V, Egan DA, Tricaro KA, Perun TJ, Baker WR, and Kleinert HD. J Med Chem 1993;<br />

36:460–467.<br />

49. (a) McQuade TJ, Tomasselli AG, Liu L, Karacostas V, Moss B, Sawyer TK, Heinrikson RL, Tarpley<br />

WG. Science 1990; 247:454–7456; (b) Sawyer TK, Fisher JF, Hester JB, Smith CW, Tomasselli AG,<br />

Tarpley WG, Burton PS, Hui JO, McQuade TJ, Conradi RA, Bradford VS, Liu L, Kinner JH, Tustin J,<br />

Alexander DL, Harrison AW, Emmert DE, Staples DJ, Maggiora LL, Zhang Y-Z, Poorman RA, Dunn<br />

BM, Rao C, Scarbourogh PE, Lowther TT, Craik C, DeCamp D, Moon J, Howe WJ, Heinrikson RL.<br />

Bioorg Med Chem Lett 1993; 3:819–824.<br />

50. Roberts NA, Martin JA, Kinchington D, Broadhurst AV, Craig JC, Duncan IB, Galpin SA, Handa<br />

BK, Kay J, Krohn A, Lambert RW, Merrett JH, Mills JS, Parkes KEB, Redshaw S, Ritchie AJ, Taylor<br />

DL, Thomas GJ, Machin PJ. Science 1990; 248:358–361.<br />

51. Ondetti MA, Rubin B, Cushman DW. Science 1977; 196:441–444.<br />

52. Patchett AA, Harris E, Tristam EW, Wyvratt MJ, Wu MT, Taub D, Peterson ER, Ikeler TJ,<br />

TenBroeke J, Payne LG, Ondeyka DL, Thorsett ED, Greenlee WJ, Lohr NS, Hoffsommer RD, Joshua<br />

H, Ruyle WV, Rothrock, JW, Aster SD, Maycock AL, Robinson FM, Hirschmann R, Sweet CS, Ulm<br />

EH, Grosse DM, Vassil TC, Stone CA. Nature 1980; 288:280–283.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_622.html (2 of 2) [4/9/2004 1:44:04 AM]


Document<br />

Page 623<br />

53. Bird J, Harper GP, Hughes I, Hunter DJ, Karran EH, Markwell RE, Miles-Williams AJ, Rahman SS,<br />

Ward RW. Bioorg Med Chem Lett 1995; 5:2593–2598.<br />

54. Porter JR, Beeley NRA, Boyce BA, Mason B, Millican A, Miller K, Leonard J, Morphy JR,<br />

O'Connell JP. Bioorg Med Chem Lett 1994; 4:2741–2746.<br />

55. Chapman KT, Durette PL, Caldwell CG, Sperow KM, Niedzwiecki LM, Harrison RK, Saphos C,<br />

Christen AJ, Olszewski JM, Moore VL, MacCoss M, Hagmann WK. Bioorg Med Chem Lett 1996;<br />

6:803–806.<br />

56. Edwards PD, Andiskik DW, Strimpler AM, Gomes B, Tuthil PA. J Med Chem 1996; 39:1112–1124.<br />

57. Shuman RT, Rothenberger RB, Campbell CS, Smith GF, Gifford-Moore DS, Gesellechen PD. J<br />

Med Chem 1993; 36:314–319.<br />

58. Okamoto S, Hijikata A. J Med Chem 1980; 23:1293–1299.<br />

59. Dolle RE, Prouty CP, Prasad CVC, Cook E, Saha A, Ross TM, Salvino JM, Helaszek CT, Ator MA.<br />

J Med Chem 1996; 39:2438–2440.<br />

60. Krausslich H-G, Schneider H, Zybarth G, Carter CA, Wimmer E. J Virology 1988; 2:4394–4397.<br />

61. Owens RA, Gesellchen PD, Houchins BJ, DiMarchi RD. Biochem Biophys Res Comm 1991;<br />

181:401–408.<br />

62. Richards AD, Roberts R, Dunn BM, Graves MC, Kay J. FEBS Letters 1989; 247:113–117.<br />

63. Smith AB, Hirschmann R, Pasternak A, Guzman MC, Yokoyama A, Sprengeler PA, Darke PL,<br />

Emini EA, Schleif WA. J Amer Chem Soc 1995; 117:11113–11123.<br />

64. Rich DH, Green J, Toth MV, Marshall GR, Kent SBH. J Med Chem 1990; 33:1288–1295.<br />

65. Kempf J, Sham HL. Current Pharmaceutical <strong>Design</strong> 1996; 2:225–246.<br />

66. Das J, Kimball SD. Bioorg Med Chem 1995; 3:999–1007.<br />

67. Bajusz S, Szell E, Bagdy D, Barbas E, Horvath G, Dioszegi M, Fittler Z, Szabo G, Juhasz A, Tomori<br />

E, Szilagyi G. J Med Chem 1990; 33:1729–1735<br />

68. Sherry S, Alkjaersig N, Fletcher AP. Am J Physiol 1995; 209:577–583.<br />

69. Fusetani N, Matsunaga S, Matsumoto H, Takebayashi Y. J Amer Chem Soc 1990; 112:7053–7054.<br />

70. Okamoto S, Hijikata A. J Med Chem 1980; 23:1293–1299.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_623.html (1 of 2) [4/9/2004 1:44:30 AM]


Document<br />

71. Sturzebecher J, Markwardt F, Voigt B, Wagner G, Walsmann P. Thromb Res 1983; 29:635–642.<br />

72. Rose C, Vargas F, Facchinetti P, Bourgeat P, Babal RB, Bishop PB, Chan SMT, Moore, ANJ,<br />

Ganellin CR, Schwartz J-C. Nature 1996; 380:403–409.<br />

73. Kohl NE, Mosser SD, deSolms SJ, Giuliani EA, Pompliano DL, Graham SL, Smith RL, Scolnick<br />

EM, Oliff A, Gibbs JB. Science 1993; 260:1934–1936.<br />

74. James GL, Goldstein JL, Brown MS, Rawson TE, Somers TC, McDowell RS, Crowley CW, Lucas<br />

BK, Levinson AD Marsters Jr JC. Science 1993; 260:1937–1942.<br />

75. Qian Y, Blaskovich MA, Seong C-M, Vogt A, Hamilton AD, Sebti SM. Bioorg Med Chem Lett<br />

1994; 21:2579–2584.<br />

76. Scholten JD, Zimmerman K, Oxender GM, Sebolt-Leopold J, Gowan R, Leonard D, Hupe DJ.<br />

Bioorg Med Chem 1996; 4:1537–1543.<br />

77. Plummer MS, Lunney EA, Para KS, Vara Prasad JVN, Shahripour A, Singh J, Stankovic CJ,<br />

Humblet C, Fergus JH, Marks JS, Sawyer TK. <strong>Drug</strong> <strong>Design</strong> Discovery 1996; 13:75–81.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_623.html (2 of 2) [4/9/2004 1:44:30 AM]


Document<br />

78. Rodriguez M, Cros<strong>by</strong> R, Alligood K, Gilmer T, Berman J. Lett Peptide Sci 1995; 2:1–6.<br />

79. Burke Jr TR, Barchi Jr JJ, George C, Wolf G, Shoelson SE, Yan X. J Med Chem 1995;<br />

38:1386–1396.<br />

80. Burke Jr TR, Kole HK, Roller PP. Biochem Biophys Res Comm 1994; 204:129–134.<br />

Page 624<br />

81. (a) Hamilton AD, Sebti SM. <strong>Drug</strong> News and Perspectives 1995; 8:138–145; (b) Buss JE, Marsters Jr<br />

JC. Chemistry and Biology 1995; 2:787–791.<br />

82. Manne V, Yan N, Carboni JM, Tuomari AV, Ricca CS, Brown JG, Andahazy ML, Schmidt RJ,<br />

Patel D, Zahler R, Weinmann R, Der CJ, Cox AD, Hunt JT, Gordon EM, Baracid M, Seizinger BR.<br />

Oncogene 1995; 10:1763–1779.<br />

83. Lerner EC, Qian Y, Blaskovich MA, Fossum RD, Vogt A, Sun J, Cox AD, Der JD, Hamilton AD,<br />

Sebti SM. J Biol Chem 1995; 270:26802–26806.<br />

84. Hunt JT, Lee VG, Leftheria K, Seizinger B, Carboni J, Mabus J, Ricaa C, Yan N, Manne V. J Med<br />

Chem 1996; 39:353–358.<br />

85. Botfield MC, Green J. Ann. Rep. Med. Chem 1995; 30:227–237.<br />

86. Kuriyan J, Cowburn D. Current Biology 1993; 3:828–837.<br />

87. Songyang Z, Shoelson SE, Chaudhuri M, Gish G, Pawson T, Haser WG, King F, Roberts T,<br />

Ratnofsky S, Lechleider RJ, Neel BG, Birge RB, Fajardo JE, Cou MM, Hanfusa H, Schaffhausen and<br />

Cantley LC. Cell 1993; 72:767–778.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_624.html (1 of 2) [4/9/2004 1:45:02 AM]<br />

88. Gilmer T, Rodriguez<br />

M, Jordan S, Cros<strong>by</strong> R,<br />

Alligood K, Green M,<br />

Kimery M, Wagner C,<br />

Kinder D, Charifson P,<br />

Hassell AM, Willard D,<br />

Luther M, Rusnak D,<br />

Sternbach DD, Mehrotra<br />

M, Peel M, Shampine L,<br />

Davis R, Robbins J, Patel<br />

IR, Kassel D, Burkhart W,<br />

Moyer M, Bradshaw T,<br />

Berman J. J Biol Chem<br />

1994; 269:31711–31719.


Document<br />

89. Plummer MS, Lunney EA, Shahripour A, Para KS, Stankovic CJ, Humblet C, Fergus J, Marks JS,<br />

Herrera R, Hubbell SE, Saltiel AR, Sawyer TK. Bioorg Med Chem 1997; 5:41–47.<br />

90. (a) Pert, CB, Synder SH. Science 1973; 179:1011–1014; (b) Lord JA, Waterfield AA, Hughes J,<br />

Kosterlitz HW. Nature 1977; 267:495–499.<br />

91. Snider MR, Constantine JW, Lowe JA, Longo KP, Lebel WS, Woody HA, Drozda SE, Desaia MC,<br />

Vinick FJ, Spencer RW, Hess H-J. Science 1991; 251:435–437.<br />

92. Chiu AT, McCall DE, Aldrich PE, Timmermans PBMWM. Biochem Biophys Res Commun 1990;<br />

172:1195–1202.<br />

93. Smith RG, Cheng K, Schoen WR, Pong S-S, Hickey G, Jacks T, Butler B, Chan WW-S, Chaung L-<br />

YP, Judith F, Taylor J, Wyvratt MJ, Fisher MH. Science 1993; 260:1640–1643.<br />

94. Evans BE, Rittle KE, Bock MG, DiPardo RM, Freidinger RM, Whitter WL, Lundell GF, Veber DF,<br />

Anderson PS, Chang RSL, Lotti VJ, Cerno DJ, Chen TB, Kling PJ, Kunkel KA, Springer JP, Hirshfield<br />

JJ. J Med Chem 1988; 31:2235–2246.<br />

95. Bock MG, DiPardo RM, Evans BE, Rittle KE, Whitter WL, Veber DF, Andeson PS, Freidinger RM.<br />

J Med Chem 1989; 32:13–16.<br />

96. Aquino CJ, Armour DR, Berman JM, Birkemo LS, Carr RAE, Croom DK, Dezube M, Dougherty<br />

RW, Ervin GN, Grizzle MK, Head JE, Hirst GC, James MK, Johnson MF, Miller LJ, Queen KL, Rimele<br />

TJ, Smith DN, Sugg EE. J Med Chem 1996; 39:562–569.<br />

97. (a) Doherty AM, Patt WC, Edmunds JJ, Berryman KA, Reisdorph BR, Plummer MS, Shahripour A,<br />

Lee C, Cheng X-M, Walker DM, Haleen SJ, Keiser JA, Flynn<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_624.html (2 of 2) [4/9/2004 1:45:02 AM]


Document<br />

MA, Welch KM, Hallak H, Taylor DG, Reynolds EE. J Med Chem 1995; 38:1259–1263; (b)<br />

Doherty AM. <strong>Drug</strong> Disc Today 1996; 1:60–70.<br />

98. De B, Plattner JJ, Bush EN, Jae H-S, Diaz G, Johnson ES, Perun TJ. J Med Chem 1989;<br />

32:2038–2041.<br />

99. Yamamura Y, Ogawa H, Chihara T, Kondo K, Onogawa T, Nakamura S, Mori T, Tominaga M,<br />

Yabuuchi Y. Science 1991; 252:572–574.<br />

Page 625<br />

100. Valentine JJ, Nakanishi S, Hageman DL, Snider RM, Spencer RW, Vinick FJ. Bioorg Med Chem<br />

Lett 1992; 2:333–338.<br />

101. Collins JL, Dambek PJ, Goldstein SW, Faraci WS. Bioorg Med Chem Lett 1992; 2:915–918.<br />

102. Gully D, Canton MM, Boigegrain R, Jeanjean F, Molimard JC, Poncelete M, Gueudet C, Heaulem<br />

M, Leris R, Brouard A, Pelaprat D, Labbe-Jullie C, Mazella J, Soubrie P, Moffrand JP, Rostene W,<br />

Kitabji P, Le Fur G. Proc Natl Acad Sci USA 1993; 229:23–28.<br />

103. Perlman S, Scham<strong>by</strong>e HT, Rivero RA, Greenlee WJ, Hjorth SA, and Schwartz TW. J Biol Chem<br />

1995; 270:1493–1496.<br />

104. Evans BE, Leighton JJ, Rittle KE, Gilbert KF, Lundell GF, Gould NP, Hobbs DW, DiPardo RM,<br />

Veber DF, Pettitbone DJ, Clineschmidt BV, Anderson PS, Freidinger RM. J Med Chem 1992;<br />

35:3919–3927.<br />

105. Vara Prasad JVN, Para KS, Lunney EA, Ortwine DF, Dunbar JB, Ferguson D, Tummino PJ, Hupe<br />

D, Tait BD, Domagala JM, Humblet C, Bhat TN, Liu B, Guerin DMA, Baldwin ET, Erickson JW,<br />

Sawyer TK. J Amer Chem Soc 1995; 116:6989–6990.<br />

106. Farmer PS, Ariens EJ. Trends Pharm Sci 1982; 3:362–365.<br />

107. Horwell DC. Approaches to Nonpeptide Ligands for Peptide Receptor Sites. Bioorg Med Chem<br />

Lett 1993; 3:No 5.<br />

108. (a) Casey AF. Adv <strong>Drug</strong> Res 1989; 18:177–289; (b) Portoghese PS. J Med Chem 1992;<br />

35:1927–1937.<br />

109. Wexler RR, Greenlee WJ, Irvin JD, Goldberg MR, Prendergast K, Smith RD, Timmermans<br />

PBMWM. J Med Chem 1996; 39:625–656.<br />

110. (a) Doughty MB, Chu SS, Misse GA, Tessel R. Bioorg Med Chem Lett 1992; 2:1497–1502; (b)<br />

Chaurasia C, Misse G, Tessel R, Doughty MB. J Med Chem 1994; 37:2242–2248.<br />

111. Nieps S, Michel MC, Dove S, Buschauer A. Bioorg Med Chem Lett 5:2065–2070.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_625.html (1 of 2) [4/9/2004 1:45:46 AM]


Document<br />

112. (a) Williams PD, Ball RG, Clineschmidt BV, Culberson JC, Erb JM, Freidinger RM, Pawluczyk<br />

JM, Perlow DS, Pettibone DJ, Veber DF. J Med Chem 1994; 2:971–985; (b) Williams PD, Anderson<br />

PS, Ball RG, Bock MG, Carroll LA, Lee Chiu S-H, Clineschmidt BV, Culberson JC, Erb JM, Evans BE,<br />

Fitzpatrick SL, Freidinger RM, Kaufman MJ, Lundell GF, Murphy JS, Pawluczyk JM, Perlow DS,<br />

Pettibone DJ, Pitzenberger SM, Thompson KL, Veber DF. J Med Chem 1994; 37:565–571.<br />

113. Greer J, Erickson JW, Baldwin JJ, Varney MD. J Med Chem 1994; 37:1035–1054.<br />

114. Colman PM. Curr Opinion Struct Biol 1994; 4:868–874.<br />

115. (a) Verlinde LMJ, Hol WGJ. <strong>Structure</strong> 1994; 2:577–587; (b) Hol WGJ. Angew Chemie 1986;<br />

25:767–852.<br />

116. Navia MA, and Murcko MA. Curr Opinion Struct Biol 1992; 2:202–210.<br />

117. Appelt K, Bacquet RJ, Barlett CA, Booth CLJ, Freer ST, Fuhry MA, Gehring MR, Herrmann SM,<br />

Howland EF, Janson CA, Jones TR, Kan C-C, Kathardekar V, Lewis KK, Marzoni GP, Matthews DA,<br />

Mohr C, Moowaw EW, Morse CA, Oatley<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_625.html (2 of 2) [4/9/2004 1:45:46 AM]


Document<br />

Page 626<br />

SJ, Ogden RC, Reddy M, Reich SH, Schoettlin WS, Smith WW, Varney MD, Villafranca JE, Ward<br />

RW, Webber S, Webber SE, Welsch KM, White J. J Med Chem 1991; 34:1925–1934.<br />

118. Marshall GR. Curr Opinion Struct Biol 1992; 2:904–919.<br />

119. Veerapandian P. In: Burger's Medicinal Chemistry and <strong>Drug</strong> Discovery. Fifth Ed. Vol. 1 (Wolff<br />

ME, Ed) Wiley and Sons Inc. 1995; 303–348.<br />

120. (a) Kuntz ID. Science 1992; 237:1078–1082; (b) Roe D, Kuntz ID. Pharm News 1995; 2:13–15.<br />

122. Fong TM, Strader CD. Med Res Rev 1994; 14:387–399.<br />

121. Ajay Murko MA. J Med Chem 1995; 38:4973–4967.<br />

123. (a) Schwartz TW, Gether U, Scham<strong>by</strong> HT, Hjorth SA. Curr Pharm <strong>Design</strong> 1995; 1:355–372; (b)<br />

Schwartz TW. Curr Opinion Biotech 1994; 5:434–444.<br />

124. Humblet C, Mirzadegan T. Ann Rep Med Chem 1992; 27:291–300.<br />

125. Findley J, Eliopoulos E. Trends Pharm Sci 1990; 11:492–499.<br />

126. (a) Hilbert MF, Trumpp-Kallmeyer S, Hoflack J, Bruinvels A. Trends Pharm Sci 1993; 14:7–12;<br />

(b) Trumpp-Kallmeyer S, Hoflack J, Bruinvels A, Hilbert M. J Med Chem 1992; 35:3448–3462; and (c)<br />

Hilbert MF, Trumpp-Kallmeyer S, Bruinvels A, Hoflack J. Molec Pharmacol 1991; 40:8–15.<br />

127. Kontoyianni M, Lybrand TP. Med Chem Res 1993; 3:407–418.<br />

128. Probst WC, Snyder LA, Schuster DI, Brosius J, Sealfon SC. DNA Cell Biol 1992; 11:1–20.<br />

129. Moereels H, Jannsen PAJ. Med Chem Res 1993; 3:335–343.<br />

130. Attwood TK, Findlay JBC. Prot Eng 1994; 7:195–203.<br />

131. Savares TM, Fraser CM. Biochem J 1992; 283:1–19.<br />

132. Brann MR. Molecular Biology of G-Protein-Coupled Receptors, Boston: Birkhauser, 1992.<br />

133. Henderson R, Baldwin JM, Ceska TA, Semlin F, Beckman E, Downing KH. J Mol Biol 1990;<br />

213:899–929.<br />

134. (a) Underwood DJ, Strader CD, Rivero R, Patchett A, Greenlee W, Pendergrast K. Chem Biol<br />

1994; 1:211–221; (b) Scham<strong>by</strong>e HT, Hjorth SA, Bergsama DJ, Sathe G, Schwartz TW. Proc Natl Acad<br />

Sci USA 1994; 91:7046–7050; (c) Yamano Y, Ohyama K, Chaki S, Guo D-F, Inagami T. Biochem<br />

Biophys Res Comm 1992; 187:1426–1431; (d) Yamano Y, Ohyama K, Kikyo M. J Biol Chem 1995;<br />

270:14024–14030.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_626.html (1 of 2) [4/9/2004 1:46:39 AM]


Document<br />

135. (a) Fong TM, Yu H, Cascieri MA, Underwood D, Swain CJ, Strader CD. J Biol Chem 1994;<br />

269:14957–14961; (b) Fong TM, Yu H, Cascieri MA, Underwood D, Swain C, Strader CD. J Biol Chem<br />

1994; 269:2728–2732; (c) Fong TM, Cascieri MA, Yu H, Bansai A, Swain CJ, Strader CD, Nature<br />

1993; 362:30–353; (d) Gether Y, Johansen TE, Snider RM, Lowe JA, Kakanishi S, Schwartz TW.<br />

Nature 1993; 362:345–348.<br />

136. (a) Beinborn M, Lee Y-M, McBridde EW, Quinn SM, Kopin AS. Nature 1993; 362:348–350; (b)<br />

Kopin AS, McBride EW, Quinn SM, Kolakowski LF, Beinborn M. J Biol Chem 1995; 270:5019–5023;<br />

(c) Silvente-Poirot S, Wank SA. J Biol Chem 1996; 271:14698–14706.<br />

137. (a) Surratt CK, Johnson PS, Moriwaski A, et al. J Biol Chem 1994; 269:20548–20553; (b) Kong H,<br />

Raynor K, Yusuda K, et al. J Biol Chem 1993; 268:23055–23058; (c) Kong H, Raynor K, Yano H,<br />

Takeda J, Bell GI, Reisine T. Proc Natl Acad Sci USA 1994; 91:8042–8046; (d) Onogi T, Miniami M,<br />

Katao Y, Nakagawa T, Aoki Y, Toya T, Katsumata S, Satoh M. FEBS Lett 1995; 357:93–<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_626.html (2 of 2) [4/9/2004 1:46:39 AM]


Document<br />

97; (e) Fujii I, Nakamura H, Sagara T, Kanematsu K. Med Chem Res 1994; 4:424–431.<br />

Page 627<br />

138. (a) Mouillac B, Chin B, Balestre M-N, Elands J, Trumpp-Kallmeyer S, Hoflack J, Hibert M, Jard S,<br />

Barberis C. J Biol Chem 270:25771–25777; (b) Chini B, Mouillac B, Ala Y, Balestre M-N, Trumpp-<br />

Kallmeyer S, Hoflack J, Elands J, Hibert M, Manning M, Jard S, Barberis C, EMBO J 1995;<br />

14:2176–2182; (c) Huggins JP, Trumpp-Kallmeyer S, Hibert M, Hoflack JM, Fanger BO, Jones CR. Eur<br />

J Pharmacol 1993; 245:203–214.<br />

139. Kyle DJ, Chakravarty S, Sinako JA, Stormann TM. J Med Chem 1994; 37:1347–1354.<br />

140. Pang Y-P, Cusack B, Groshan K, Richelson E. J Biol Chem 1996; 271:15060–15068.<br />

141. Haskell-Luevano C, Sawyer TK, Trumpp-Kallmeyer Bikker J, Humblet C, Gantz I, Hru<strong>by</strong> VJ.<br />

<strong>Drug</strong> <strong>Design</strong> Disc. 1996; 14:197–211.<br />

142. (a) Rich DH, In: Sammes PG, Ed. Comprehensive Medicinal Chemistry. New York: Pergamon<br />

Press, 1990: 391–441; (b) Lawton G, Paciorek PM, Waterfall JF. Adv <strong>Drug</strong> Res 1992; 23:12–220; (c)<br />

Wyvratt MJ, Patchett AA. Med Res Rev 1985; 5:483–531.<br />

143. James MN, and Sielecki AR. J Mol Biol 1983; 163:299–361.<br />

144. Suguna K, Bott RR, Padlan EA, Subramanian E, Sheriff S, Cohen GH, Davis DR. J Mol Biol 1987;<br />

196:877–900.<br />

145. Blundell TL, Jenkins JA, Sewell BT, Pearl LH, Cooper JB, Tickle IJ, Veerpandian B, Wood SP. J<br />

Mol Biol 1990; 211:919–941.<br />

146. Sielecki AR, Fujinaga M, Read RJ, James MN, J Mol Biol 1991; 219:671–.<br />

147. (a) Abad-Zapatero C, Rydel TJ, Erickson J, Proteins 1990; 8:62-; (b) Cooper JB, Khan G, Taylor<br />

G, Tickle IJ, Blundell TL. J Mol Biol 1990; 214:199-; (c) Sielecki AR, Fedorov AA, Boodhoo A,<br />

Andreeva NS, James MN. J Mol Biol 1990; 214:143–170.<br />

148. (a) Sibanda BL, Blundell T, Hobart PM, Fogliano M, Bindra JS, Dominy BW, Chirgwin JM. FEBS<br />

Lett 1984; 174:96–101; (b) Carson W, Karplus M, Haber E. Hypertension 1985; 7:13–26; (c) Akahane<br />

K, Umeyama H, Nakagawa S, Moriguchi I, Hirose S, Iizuka K, Murakami Hypertension 1985; 7:3–12;<br />

(d) Sham HL, Bolis G, Stein HH, Fesik SW, Marcotte PA, Plattner JJ, Remple CA, Greer J. J Med<br />

Chem 1988; 31:284–295.<br />

149. (a) Abdel-Meguid SS, Med Res Rev 1993; 13:731–778; (b) Hutchins C, Greer J. Crit Rev Biochem<br />

Mol Biol 1991; 26:77–127; (c) Greenlee W. Med Res Rev 1990; 10:173–236.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_627.html (1 of 2) [4/9/2004 1:47:10 AM]


Document<br />

150. (a) Cooper JB, Foundling SI, Watson FI, Sibanda BL, Blundell TL. Biochem Soc Trans 1987;<br />

15:751–754; (b) Cooper JB, Foundling SI, Hemmings A, Blundell TL, Jones DM, Hallett A, Szelke M,<br />

Eur J Biochem 1987; 169:215–21; (c) Hallett A, Jones DM, Atrash B, Szelke M, Leckie B, Beattle S,<br />

Dunn BM, Valler MJ, Rolph CE, Jay J, Foundling SI, Wood S, Pearl LH, Watson FE, Blundell TL. In:<br />

Kostka V., Ed. Aspartic Proteinases and Their Inhibitors. Berlin: de Gruyter, (1985); 467–478; (d) Sali<br />

A, Veerapandian B, Cooper JB, Foundling SI, Hoover DJ, Blundell TL. EMBO J 1989; 2179–2188; (e)<br />

Veerapandia B, Cooper JB, Sali A, Blundell TL. J Mol Biol 1990; 216:1017–1029; (f) Suguna K, Padlan<br />

EA, Smith CW, Carlson WD, Davies DR, Proc Natl Acad Sci USA 1987; 84:7009–7013; (g) Abad-<br />

Zapatero C, Rydel TJ, Neidhart DJ, Luly J, Erickson JW, Adv Exp<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_627.html (2 of 2) [4/9/2004 1:47:10 AM]


Document<br />

Med Biol 1991; 306:9–21; (h) James MN, Sielecki AR, Hayakawa K, Gelb MH. Biochem 1992;<br />

31:387.<br />

Page 628<br />

151. (a) Sielecki AR, Hayakawa K, Fujinaga M, Murphy ME, Fraser M, Muir AK, Carilli CT, Lewicki<br />

JA, Baxter JD, James, MN. Science 1989; 243:1346–1351; (b) Rahuel J, Priestle JP, Grutter MG. J<br />

Struct Biol 1991; 107:227–236.<br />

152. Plummer MS, Shahripour A, Kaltenbronn JS, Lunney EA, Steinbaugh BA, Ham<strong>by</strong> JM, Hamilton<br />

HW, Sawyer TK, Humblet C, Doherty AE, Taylor MD, Hingorani G, Batley BL, Rapundalo ST. J Med<br />

Chem 1995; 38:2893–2905.<br />

153. Rasetti V, Cohen NC, Rueger H, Boschke R, Maibaum J, Cumin F, Fuhrer W, Wood JM. Bioorg<br />

Med Chem Lett 1996; 6:1589–1594.<br />

154. Weber AE, Halgren TA, Doyle JJ, Lynch RJ, Siegel PKS, Parsons WH, Greenlee WJ, Patchett AA.<br />

J Med Chem 1991; 34:2692–2701.<br />

155. (a) Dutta AS, Gormely JJ, McLachlan PF, Major JS. J Med Chem 1990; 33:2552–2560; (b) Dutta<br />

AS, Gormley JJ, McLachlan PF, Major JS. J Med Chem 1990; 33:2560–2568.<br />

156. (a) Kempf J, Sham HL. Curr Pharm <strong>Design</strong> 1996; 2:225–246; (b) Wlodawer A, Erickson JW. Ann<br />

Rev Biochem 1993; 62:543–585; (c) Appelt K. Perspect <strong>Drug</strong> Disc <strong>Design</strong> 1993; 1:23–48; (d) Meek<br />

TD. J Enz Inhib 1992; 6:65–98 (e); Tomasselli AG, Howe WJ, Sawyer TK, Wlodawer A, Heinrikson<br />

RL. Chim Oggi 1991; 9:6–27; (f) Huff JR. J Med Chem 1991; 34:2305–2314.<br />

157. (a) Erickson J, Neidhart DJ, VanDrie J, Kempf DJ, Wang XC, Norbeck DW, Plattner JJ,<br />

Rittenhouse JW, Turon M, Wideburg N, Kohlbrenner WE, Simmer R, Helfrich R, Pau DA, Knigge M,<br />

Science 1990; 249:529–533; (b) Kempf DJ, Norbeck DW, Codacovi L, Wang XC, Kohlbrenner WE,<br />

Wideburg NE, Paul DA, Knigge MF, Vasavanonda S, Criag-Kennard A, Saldivar A, Rosenbrook JW,<br />

Clement JJ, Plattner JJ, Erickson J. J Med Chem 1990; 33:2687–2689.<br />

158. Lam PYS, Jadhav PK, Eyermann CJ, Hodge CN, Ru Y, Bacheler LT, Meek JL, Otto MJ, Rayner<br />

MM, Wong YN, Chang C-H, Weber PC, Jackson DA, Sharpe TR, Erickson-Viitanen S. Science 1994;<br />

263:380–383.<br />

159. Kempf DJ, March KC, Denissen JF, McDonald E, Vasavanonda S, Flentge CA, Green BE, Fino L,<br />

Park CH, Kong X-P, Wideburg NE, Saldivar A, Ruiz A, Kati WM, Sham HL, Robins T, Stewart KD,<br />

Hsu A, Plattner JJ, Leonard JM, Norbeck DW. Proc Natl Acad Sci USA 1995; 92:2484–2488.<br />

160. Vacca JP, Dorsey BD, Schlief WA, Levin RB, McDaniel SL, Darke PL, Zugay J, Quintero JC,<br />

Blahy OM, Roth E, Sardana VV, Schlabach AJ, Graham PI, Condra JH, Gotlib L, Holloway MK, Lin J,<br />

Chen I-W, Vastag K, Ostovic D, Anderson PS, Emini EA, Huff JR. Proc Natl Acad Sci USA 1994;<br />

91:4096–4100.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_628.html (1 of 2) [4/9/2004 1:47:40 AM]


Document<br />

161. (a) Kalish VJ, Tatlock JH, Davies JF II, Kaldor SW, Dressman BA, Reich S, Pino M, Nyugen D,<br />

Appelt K, Musick L, Wu B-W. Bioorg Med Chem Lett 1995; 5:727-; (b) Kalish V, Kaldor S, Shetty B,<br />

Tatlock J, Davies J, Hammond M, Dressman B, Fritz J, Appelt K, Reich S, Musick L, Wu BW, Su K.<br />

Eur J Med Chem 1995; 30:S201-.<br />

162. (a) Kim<br />

EE, Baker CT,<br />

Dwyer MD,<br />

Murko MA,<br />

Rao BG, Tung<br />

RD, Navia<br />

MA. J Am<br />

Chem Soc<br />

1995;<br />

117:1181-; (b)<br />

Partaledis JA,<br />

Yamaguchi K,<br />

Tisdale M,<br />

Blair EE,<br />

Falcione C,<br />

Maschera B,<br />

Myers RE,<br />

Pazhanisamy<br />

S, Futer O,<br />

Culliman AB,<br />

Stuver CM,<br />

Byrn RA,<br />

Livingston DJ.<br />

J Virology<br />

1995;<br />

69:5228–5235.<br />

163. (a) Getman DP, DeCrescenzo GA, Heintz RM, Reed KL, Talley JJ, Bryant ML, Clare M,<br />

Houseman KA, Mar JJ, Mueller RA, Vazquez ML, Shieh H-S, Stallings WC, Stegeman RA. J Med<br />

Chem 1993; 36:288–291; (b) Vazquez ML, Bryant<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_628.html (2 of 2) [4/9/2004 1:47:40 AM]


Document<br />

Page 629<br />

ML, Clare M, Decrescenzo GA, Doherty EM, Freskos JN, Getman DP, Houseman KA, Julien JA,<br />

Kocan GP, Mueller RA, Shieh HS, Stallings WC, Stegeman RA, Talley JJ. J Med Chem 1995;<br />

38:581.<br />

164. (a) Desjarlais RL, Seibel GL, Kuntz ID, Furth PS, Alvarez JC, Ortiz DeMontellano PR, DeCamp<br />

DL, Babe LM, Craik CS. Proc Natl Acad Sci USA 1990; 87:6644–6648; (b) Rutenbar E, Fauman EB,<br />

Keenan RJ, Fong S, Furth PS, Ortiz de Montellano PR, Meng E, Kuntz ID, DeCamp DL, Salto R, Rose<br />

JR, Craik CS. Strand RM. J Biol Chem 1993; 268:15343–15346.<br />

165. Vara Prasad JVN, Para KS, Lunney EA, Ortwine DF, Dunbar JB, Jr Ferguson D, Tummino PJ,<br />

Hupe D, Tait BD, Domagala JM, Humblet C, Bhat TN, Liu TN, Buerin DMA, Baldwin ET, Erickson<br />

JW, Sawyer TK. J Am Chem Soc 1994; 116:6989–6990.<br />

166. Thaisrivongs S, Watenpaugh KD, Howe WJ, Tomich PK, Dolak LA, Chong KT, Tomich CSC,<br />

Tomasselli AG, Turner SR, Strohbach JW, Mulichack AM, Janakiraman MN, Moon JB, Lynn JC,<br />

Horng MM, Hinshaw RR, Curry KA, Rothcrock DJ. J Med Chem 1995; 38:3624–3637.<br />

167. Skulnick HI, Johnson PD, Howe WJ, Tomich PK, Chong KT, Watenpaugh KD, Janakiraman MN,<br />

Dolak LA, McGrath JP, Lynn JC, Horng MM, Hinshaw RR, Zipp GL, Ruwart MJ, Schwende FJ, Zhong<br />

WZ, Padbury GE, Dalga RJ, Shiou LH, Possert PL, Rush BD, Wilkinson KF, Howard GM, Toth LN,<br />

Williams MG, Kakuk TJ, Cole SL, Zaya TM, Lovasz KD, Morris JK, Romines KR, Thaisrivongs S,<br />

Aristoff PA. J Med Chem 1995; 38:4968–4971.<br />

168. Tummino PJ, Vara Prasad JVN, Ferguson D, Nouhan C, Graham N, Domagala JM, Ellsworth E,<br />

Gadja C, Hagen SE, Lunney EA, Para KS, Tait BD, Pavlosky A, Erickson JW, Gracheck S, McQuade<br />

TJ, Hupe DJ. Bioorg Med Chem. 1996; 4: 1401–1410.<br />

169. (a) Navia MA, Fitzgerald PM, McKeever BM, Leu CT, Heimbach JC, Herber WK, Sigal IS, Darke<br />

PL, Springer JP. Nature 1989; 337:615–620; (b) Wlodawer A, Miller M, Jaskolski M, Sathyranarayana<br />

BK, Baldwin E, Weber IT, Selk LM, Clawson L, Schneider J, Kent SB. Science 1989; 245:616–621.<br />

170. Miller M, Schneider J, Sathyanarayana BK, Toth MV, Marshall GR, Clawson L, Selk L, Kent<br />

SBH, Wlodawer A, Science 1989; 246:1149–1152.<br />

171. Jaskolski M, Tomasselli AG, Sawyer TK, Staples DJ, Heinrikson RL, Schneider J, Kent SBH,<br />

Wlodawer A. Biochem 1991; 30:1600–1609.<br />

172. Swain AL, Miller MM, Green J, Rich DH, Schneider J, Kent SBH, Wlodawer A. Proc Natl Acad<br />

Sci USA 1990; 87:8805–8809.<br />

173. Fitzgerald PMD, McKeever BM, Van Middlesworth JF, Springer JP, Heimbach JC, Leu C-T,<br />

Herber WK, Dixon AF, Darke PL. J Biol Chem 1990; 265:14209–14219.<br />

174. Weber PC, Ohlendorf DH, Wendoloski JJ, Salemme FR. Science 1989; 243:85–88.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_629.html (1 of 2) [4/9/2004 1:47:58 AM]


Document<br />

175. (a) Stubbs MT, Bode W. Trends Cardiovasc Med 1995; 5:157–166. (b) Stubbs MT, Bode W.<br />

Perspectives in <strong>Drug</strong> Disc <strong>Design</strong> 1994; 1:431–452; (c) Bode W, Huber R, Rydel TJ, Tulinsky A. In:<br />

Berliner LJ, Ed. Thrombin: <strong>Structure</strong> and Function. New York: Plenum Press, 1992:3–62; (d) Powers<br />

MC, Kam C-M. In: Berliner LJ, Ed. Thrombin: <strong>Structure</strong> and Function. New York: Plenum Press,<br />

1992:117–158.<br />

176. Weber PC, Lee S-L, Lewandowski FA, Schadt MC, Chang C-H, Kettner CA. Biochem 1995;<br />

34:3750–3757.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_629.html (2 of 2) [4/9/2004 1:47:58 AM]


Document<br />

Page 630<br />

177. (a) Bode W, Turk D, Karshikov A, Protein Sci 1992; 1:426–471; (b) Bode W, Mayr I, Baumann U,<br />

Huber R, Stone SR, Hofsteenge J. EMBO J 1989; 8:3467–3477; (c) Rydel TJ, Tulinksy A, Bode W,<br />

Huber R. J Mol Biol 1991; 221:583–601.<br />

178. Arni RK, Padmanabhan K, Padmanabhan KP, Wu TP, Tulinsky A. Biochem 1993; 32:4727–4737.<br />

179. Maryanoff BE, Qiu X, Pamanhabhan KP, Tulinsky A, Almond HR Jr, Andrade-Gordon P, Greco<br />

MN, Kauffman JA, Nicolaou KC, Liu A, Burngs PH, Fusetani N. Proc Natl Acad Sci USA 1993;<br />

90:8084–8052.<br />

180. Wiley MR, Chirgadze NY, Clawson DK, Craft TJ, Gifford-Moore DS, Jones ND, Olkowski JL,<br />

Schacht AL, Weir LC, Smith GF. Bioorg Med Chem Lett 1995; 5:2835–2840.<br />

181. (a) Brandstetter H, Turk D, Hoeffken HW, Grosse D, Sturzebecher J, Martin PD, Edwards BFP,<br />

Bode W. J Mol Biol 1992; 266:1085–1099; (b) Banner DW, Hadvary P. J Biol Chem 1991;<br />

266:20085–20093.<br />

182. Hilpert K, Ackermann J, Banner DW, Gast A, Gubernator K, Hadvary P, Labler L, Muller K,<br />

Schmid G, Tachopp TB, van de Waterbeemd H. J Med Chem 1994; 37:3889–3901.<br />

183. Padmanabhan K, Padmanabhan KP, Tulinsky A. J Mol Biol 1993; 232:947–966.<br />

184. (a) Walter J, Steigemann W, Singh TP, Bartunik H, Bode W, Huber R. Acta Cryst 1982;<br />

B38:1462–1472; (b) Bode W, Turk D, Sturzebecher J. Eur J Biochem 1990; 193:175–182; (c) Stubbs<br />

MT, Huber R, Bode W. FEBS Lett 1995; 375:103–107.<br />

185. Chen ZG, Bode W. J Mol Biol 1983; 164:283–311.<br />

186. Meyer EF. Acta Crystallogr 1988; B44:26–38.<br />

187. Bernstein PR, Andiski D, Bradley PK, Bryant CB, Ceccarelli C, Damewood JR Jr, Earley R,<br />

Edwards PD, Feeney S, Gomes BC, Kosmider BJ, Steelman GB, Thomas RM, Vacek EP, Veale CA,<br />

Williams JC, Wolanin DJ, Woolson SA. J Med Chem 1994; 37:3313–3326.<br />

188. Edwards PD, Meyer EF Jr, Vijayalakshmi J, Tuthill PA, Andisik DA, Gomes B, Strimpler A. J Am<br />

Chem Soc 1992; 114:1854–1863.<br />

189. (a) Navia MA, McKeever BM, Springer JP, Lin T-Y, Williams HR, Fluder EM, Dorn CP,<br />

Hoogsteen K. Proc Natl Acad Sci USA 1989; 86:7–11; (b) Takahashi LH, Radhakrishnan R, Rosenfield<br />

RE Jr, Meryer EF Jr, Trainor DA, Stein M. J Mol Biol 1988; 201:423–428.<br />

190. Takahashi LH, Radhakrishnan R, Rosenfield RE Jr, Meyer EF, Trainor DA. J Am Chem Soc 1989;<br />

111:3368–3374.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_630.html (1 of 2) [4/9/2004 1:48:30 AM]


Document<br />

191. (a) Peisach E, Casebier D, Gallion SL, Furth P, Petsko GA, Hogan JC Jr, Ringe D. Science 1995;<br />

269:66–69; (b) Mattos C, Giammona DA, Petsko GA, Ringe D. Biochem 1995; 34:3193–3205; (c)<br />

Mattos C, Rasmussen B, Ding X, Petsko G, Ringe D. Nature Struct Biol 1994; 1:55–58.<br />

192. (a) Edwards PD, Bernstein PR. Med Res Rev 1994; 14:127–194; (b) Hiasta DJ, Pagani ED. Ann<br />

Rep Med Chem 1994; 29:195–204.<br />

193. Drenth J, Jansonius JN, Koekoek R, Swen HM, Wolthers BG. Nature 1968; 218:929–932.<br />

194. Drenth J, Kalk KH, Swen HM. Biochem 1976; 15:3731–3738.<br />

195. Schroeder E, Phillips C, Gaman E, Harlos K, Crawford C. FEBS Lett 1993; 315:38–42.<br />

196. Rauber P, Angliker H, Walker B, Shaw E. Biochem J 1986; 239:633–640.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_630.html (2 of 2) [4/9/2004 1:48:30 AM]


Document<br />

Page 631<br />

197. Musil D, Zucic D, Turk D, Engh RA, Mayr I, R Huber Popovic T, Turk V, Towatari T, Katunuma<br />

N, Bode W. EMBO J 1991; 10:2321–2330.<br />

198. (a) Allair M, Chernaia M, Malcolm BA, James MNG. Nature 1994; 369:72–77; (b) Matthews DA,<br />

Smith WW, Ferre RA, Condon B, Budahazi G, Sisson W, Villafranca JE, Janson CA, McElroy HE,<br />

Gribskow CL, Worland S. Cell 1994; 77:761–771.<br />

199. Wilson KP, Black JF, Thomson JA, Kim EE, Griffith JP, Navia MA, Murcko MA, Chambers SP,<br />

Aldape RA, Raybuck SS, Livingson DJ, Nature 1994; 370:270–275.<br />

200. Walker NPC, Talanian RV, Brady KD, Dang LC, Bump NJ, Ferenz CR, Franklin S, Ghayur T,<br />

Hackett MC, Hammill LD, Herzog L, Hugunin M, Houy W, Mankovich JA, McGuiness L, Orlewicz E,<br />

Paskind M, Pratt CA, Reis P, Summani A, Terranova M, Welch JP, Xiong L, Moller A, Tracey DE,<br />

Kamen R, Wong WW. Cell 1994; 78:343–352.<br />

201. Ator MA, Dolle RE. Curr Pharm <strong>Design</strong> 1995; 1:191–210.<br />

202. Mullikan MD, Lauffer DJ, Gillespie RJ, Matharu SS, Kay D, Porritt GM, Evans PL, Golec JMC,<br />

Murcko MA, Luong Y-P, Raybuck SA, Livingston DJ. Bioorg Med Chem Lett 1994; 4:2359–2364.<br />

203. Dole RE, Singh J, Rinker J, Hoyer D, Prasad CVC, Graybill TL, Salvino JM, Helaszek CT, Miller<br />

RE, Ator MA. J Med Chem 1994; 37:3863–3866.<br />

204. Dole RE, Prouty CP, Prasad CVC, Cook E, Saha A, Morgan Ross T, Salvino JM, Helaszek CT,<br />

Ator MA. J Med Chem 1996; 39:2438–2440.<br />

205. Rotonda J, Nicholson DW, Fazil KM, Gallant M, Gareau Y, Labelle M, Peterson EP, Rasper DM,<br />

Ruel R, Vaillancourt JP, Thornberry NA, Becker JW, Nature (Struct Biol) 1996; 3:619–625.<br />

206. Matthews BW. Acc Chem Res 1988; 21:333–340.<br />

207. (a) Matthews BS, Jansonius JN, Colman PM, Schoenborn B, Dupourque D. Nature (London) New<br />

Biol 1972; 238:37–41; (b) Holmes MA, Matthews BW. J Mol Biol 1982; 160:623–639.<br />

208. Weaver LH, Kester WR, Matthews BW. J Mol Biol 1977; 114:119–132.<br />

209. (a) Holden HM, Tronrud DE, Monzingo AF, Weaver LH, Matthews BW. Biochem 1987; 26:8542;<br />

(b) Bartlett PA, Marlowe CK. Biochem 1987; 26:8553–8561.<br />

210. (a) Montzingo AR, Matthews BW. Biochem 1984; 23:5724-; (b) Maycock AL, DeSousa DM,<br />

Payne LG, ten Broeke J Wu MT, Patchett AA. Biochem Biophys Res Commun 1981; 102:963–969.<br />

211. (a) Holmes MA, Matthews BW, Biochem 1981; 20:6912; (b) Nishino N, Powers JC. Biochem<br />

1978; 17:2846–2850.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_631.html (1 of 2) [4/9/2004 1:48:58 AM]


Document<br />

212. (a) Monzingo AF, Matthews BW, Biochem 1982; 21:3390. (b) Nishino N, Powers JC. Biochem<br />

1979; 18:4340–4347.<br />

213. (a) Lipscomb WN, Hartsuck JA, Quiocho FA, Reeke GN Jr, Proc Natl Acad Sci NSA 1969;<br />

64:28–35. (b) Christianson DW, Lipscomb WN. J Am Chem Soc 1988; 110:5560–5565.<br />

214. Krapcho J, Turk C, Cushman DW, Rubin B, Powell JR, DeForrest JM, Spitzmiller ER,<br />

Karanewsky DS, Duggan M, Rovnyak G, Schwartz J, Natarajan S, Godfrey JD, Ryono DE, Neubeck R,<br />

Atwal KS, Petrillo EW Jr. J Med Chem 1988; 31:1148–1160.<br />

215. Flynn GA, Beight DW, Mehdi S, Koehl JR, Giroux EL, French JF, Hake PW, Dage RC. J Med<br />

Chem 36:2420–2423.<br />

216. (a) Browner MF, Persp <strong>Drug</strong> Disc <strong>Design</strong> 1994; 2:343–351; (b) Morphy JR, Millican TA, Porter<br />

JR. Curr Med Chem 1995; 2:743–762; (c) Beckett RP, David-<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_631.html (2 of 2) [4/9/2004 1:48:58 AM]


Document<br />

son AH, Drummond AH, Huxley P, Whittaker M. <strong>Drug</strong> Disc Today 1996; 1:16–26.<br />

Page 632<br />

217. (a) Spurlino JC, Smallwood AM, Carlton DD, Banka TM, Vavra KJ, Johnson JS, Cook ER, Falvo<br />

J, Wahl RC, Pulvino TA, Wendoloski JJ, Smith DL. Proteins: Struct Funt Genet 1994; 19:98–109; (b)<br />

Lovejoy B, Hassell AM, Luther MA, Weigl D, Jordon SR. Biochem 1994; 33:8207–8217.<br />

218. Stams T, Spurlino JC, Smith DL, Wahl RC, Ho TF, Qoronfleh MW, Banks TM, Rubin B. Nature<br />

(Struct Biol) 1994; 1:119–123.<br />

219. Borkakoti N, Winkler FK, Williams DH, D'Arcy A, Broadhurst MJ, Brown PA, Johnson WH,<br />

Murray EJ. Nature (Struct Biol) 1994; 1:106–110.<br />

220. Lovejoy B, Cleas<strong>by</strong> A, Hassell AM, Longley K, Luther MA, Weigl D, McGeehan G, McElroy AB,<br />

Drewry D, Lambert MH, Jordon ST. Science 1994; 263:375–377.<br />

221. Van Doren SR, Kurochkin AV, Ye, Qi-Zhuang, Johnson LL, Hupe DJ, Zuiderweg ERP. Biochem<br />

1993; 32:13109–13122.<br />

222. Gooley PR, O'Connell JF, Marcy AI, Cuca GC, Salowe SP, Bush BL, Hermes JD, Esser CK,<br />

Hagmann WK, Springer JP, Johnson BA. Nature (Struct Biol) 1994; 1:111–119.<br />

223. Becker JW, Marcy AI, Rokosz LL, Axel MG, Burbaum JJ, Fitzgerald PMD, Cameron PJ, Esser<br />

CK, Hagmann WK, Hermes JD, Springer JP. Protein Sci 1995; 4:1966–1976.<br />

224. Dhanarij V, Ye Q-Z, Johnson LL, Hupe DJ, Ortwine DF, Dunbar JB Jr, Rubin JR, Pavlosky A,<br />

Humblet, Blundell TL. <strong>Structure</strong> 1996 4:375–386.<br />

225. Gromis-Ruth FX, Stocker W, Huber R, Zwilling R, Bode W. J Mol Biol 1993; 229:945–968.<br />

226. Grams R, Diver V, Yiotakis A, Yiallouros I, Vassiliou S, Zwilling R, Bode W, Stocker W. Nature<br />

(Struct Biol) 1996; 3:671–675.<br />

227. Beszant B, Bird J, Gaster LM, Harper GP, Hughes I, Karran EH, Markwell RE, Miles-Williams AJ,<br />

Smith SA. J Med Chem 1993; 36:4030–4039.<br />

228. Bird J, De Mello RC, Harper GP, Hunter DJ, Karran EH, Markwell RE, Miles-Williams AJ,<br />

Rahman SS, Ward RW. J Med Chem 1994; 37:158–169.<br />

229. Brown FK, Brown PJ, Bickett DM, Chambers CL, Davies HG, Deaton DN, Drewry D, Foley F,<br />

McElroy AB, Gregson M, McGeehan GM, Myers PL, Norton D, Salovich JM, Schoenen FJ, Ward P. J<br />

Med Chem 1994; 37:674–688.<br />

230. Castelhano AL. Bioorg Med Chem Lett 1995; 5:1415–1420.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_632.html (1 of 2) [4/9/2004 1:49:16 AM]


Document<br />

231. (a) Sun H, Tongs NK. TIBS 1994; 19:480–485; (b) Fry MJ, Panayotou G, Booker GW, Waterfield<br />

MD, Protein Sci 1993; 2:1785–1797; (c) Pawson T. Curr Opinion Genet Develop 1992; 2:4–12; (d)<br />

Ulrich A, Schlessinger J. Cell 1990; 61:203–212.<br />

232. (a) Taylor SS, Radzio-Andzelm <strong>Structure</strong> 1994; 2:345–355; (b) Taylor SS, Knighton DR, Zheng J,<br />

Sowadski JM, Gibbs CS, Zoller MJ. TIBS 1993; 18:84–89.<br />

233. (a) Mauro LJ, Dixon JE. TIBS 1994; 19:152–155; (b) Feng G-S, Pawson T. Trends Genetics 1994;<br />

10:54–58; (c) Fischer EH, Charbonneau H, Tonks NK. Science 1991; 253:401–406.<br />

234. Shenolikar S. Ann Rev Cell Biol 1994; 10:55–86.<br />

235. (a) O'Keefe SJ, O'Neill EA. Perspect <strong>Drug</strong> Discovery Des 1994; 2:57–84; (b) Fischer G. Angew<br />

Chem Int Ed Engl 1994; 33:1415–1436; (c) Armistead DM, Harding MW. Ann Rep Med Chem 1993;<br />

28:207–215.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_632.html (2 of 2) [4/9/2004 1:49:16 AM]


Document<br />

236. (a) Marshall MS. TIBS 1993; 18:250–255; (b) Milburn MV, Tong L, DeVos AM, Brunger A,<br />

Yamaizumi Z, Nishimura S, Kim S-H. Science 1990; 247:939–945.<br />

237. (a) Neer EJ. Protein Sci 1994; 3:3–14; (b) Simon MI, Strathmann MP, Gautam N. Science 1991;<br />

252:802–808.<br />

Page 633<br />

238. Overduin M, Rios CB, Mayer BJ, Baltimore D, Cowburn D. Cell 1992; 70:69–704.<br />

239. Waksman G, Shoelson SE, Pant N, Cowburn D, Kuriyan J. Cell 1993 72:779–790.<br />

240. Xu RX, Word JM, Davis DG, Rink MJ, Willard MJ Jr, Gampe RT Jr. Biochem 1995;<br />

34:2107–2121.<br />

241. Pascal SM, Singer AU, Gish G, Yamazaki T, Shoelson SE, Pawson T, Kay LE, Forman-Kay JD.<br />

Cell 1994 77:461–472.<br />

242. Zhou M-M, Meadows RP, Logan TM, Yoon HS, Wade WS, Ravichandran KS, Burakoff SJ, Fesik<br />

SW. Proc Natl Acad Sci USA 1995; 92:7784–788.<br />

243. (a) Eck MJ, Shoelson SE, Harrison SC. Nature 1993; 362:87–91; (b) Eck MJ, Atwell SK, Shoelson<br />

SE, Harrison SC. Nature 1994; 368:764–769; (c) Mikol V, Baumann G, Keller TH, Manning U, Zurini<br />

MGM. J Mol Biol 1995; 246:344–355; (d) Tong L, Warren TC, King J, Betageri R, Rose J, Jakes S. J<br />

Mol Biol 1996; 256:601–610.<br />

244. Nolte RT, Eck MJ, Schlessinger J, Shoelson SE, Harrison SC. Nature (Struct Biol) 1996;<br />

3:364–374.<br />

245. Lee C-H, Kominos D, Jacques S, Margolis B, Schlessinger J, Shoelson SE, Kuriyan J. <strong>Structure</strong><br />

1994; 2:423–438.<br />

246. Rahuel J, Gay B, Erdmann D, Strauss A, Garcia-Echeverria C, Furet P, Caravatti G, Fretz H,<br />

Schoepfer J, Grutter MG. Nature (Struct Biol) 1996; 3:586–589.<br />

247. Narula SS, Yuan RW, Adams SE, Green OM, Green J, Philips TB, Zydowsky LD, Botfield MC,<br />

Hatada M, Laird ER, Zoller MJ, Karas JL, Dalgarno DC. <strong>Structure</strong> 1995; 3:1061–1073.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_633.html (1 of 3) [4/9/2004 1:50:03 AM]


Document<br />

248. Hatada<br />

MH, Lu X,<br />

Laird ER,<br />

Green J,<br />

Morgenstern<br />

JP, Lou M,<br />

Marr CS,<br />

Phillips TB,<br />

Ram MK,<br />

Theriault K,<br />

Zoller MJ,<br />

Karas JL.<br />

Nature 1995;<br />

377:32–38.<br />

249. (a) Yu H, Rosen MK, Shin TB, Seidel-Dugan C, Brugge JS, Schreiber SL. Science 1992;<br />

258:1655–1668; (b) Feng S, Chen JK, Yu H, Simon JA, Schreiber SL. Science 1994; 266:1241–1247.<br />

250. Musacchio A, Saraste M, Wilmann M. Nature (Struct Biol) 1994; 1:546–551.<br />

251. Wu X, Knudsen B, Feller SM, Zheng J, Sali A, Cowburn D, Hanafusa H, Kuriyan J. <strong>Structure</strong><br />

1995; 3:215–226.<br />

252. Goudreau N, Cornille F, Duchesne M, Parker F, Tocque B, Garbay C, Roques BP. Nature (Struct<br />

Biol) 1994; 1:898–907.<br />

253. Maignan S, Guilloteau J-P, Fromage N, Arnoux B, Becquart J, Ducruix A. Science 1995;<br />

268:291–293.<br />

254. Hubbard SR, Wei L, Ellis L, Hendrickson WA. Nature 1994; 372:746–754.<br />

255. (a) Zheng J, Knighton DR, Xuong H-H. Taylor SS, Sowadski JM, Ten Eyck LF. Protein Sci 1993;<br />

2:1559–1573; (b) Karlsson R, Zheng J, Xuong N-H, Taylor SS, Sowadski JM. Acata Crystallogr 1993;<br />

D49:381–388.<br />

256. (a) De Bondt HL, Rosenblatt J, Jancarik J, Jones HD, Morgan DO, Kim SH. Nature 1993;<br />

363:595–02; (b) Russo AA, Jeffrey PD, Patten AK, Massague J, Pavletich NP. Nature 1996;<br />

38:325–331.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_633.html (2 of 3) [4/9/2004 1:50:03 AM]


Document<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_633.html (3 of 3) [4/9/2004 1:50:03 AM]<br />

257. Zhang<br />

F, Strand A,<br />

Robbins D,<br />

Cobb MH,<br />

Goldsmith<br />

EJ. Nature<br />

1994;<br />

367:704–710.


Document<br />

Page 634<br />

258. (a) Zhang M, Van Etten RL, Stauffacher CV. Biochem 1994; 33:11097–11105; (b) Barford D, Flint<br />

AJ, Tonks NK. Science 1995; 263:1397–1404; (c) Jia Z, Barford D, Flint AJ, Tonks NK. Science 1995;<br />

268:1754–1758.<br />

259. Yuvaniyama J, Denu JM, Dixon JE, Saper MA. Science 1996; 272:1328–1331.<br />

260. Goldberg J, Huang H-B. Kwon Y-G, Greengard P, Nairn AC, Kuriyan J. Nature 1995;<br />

376:745–753.<br />

261. Eck MJ, Dhe-Paganon S, Trub T, Nolte RT, Shoelson SE. Cell 1996; 65:695–705.<br />

262. Zhou M-M, Huang B, Olejniczak ET, Meadows RP, Shuker SB, Miyazaki M, Trub T, Shoelson<br />

SE, Fesik SW. Nature (Struct Biol) 1996; 3:388–393.<br />

263. Pflugl G, Kallen J, Scuirmer T, Jansonius JN, Zurini MGM, Walkinshaw MD. Nature 1993;<br />

361:91–94.<br />

264. Van Duyne GD, Standaert RG, Karplus PM, Schreiber SL, Clardy J. Science 1991; 252:839–842.<br />

265. Saltiel AR. Scientific American (Sci Med) 1995; 2:58–67.<br />

266. Bridges AJ. Chemtracts (Org Chem) 1995; 8:73–107.<br />

267. Huber HE, Koblan KS, Heimbrook DC. Curr Med Chem 1994; 1:13–34.<br />

268. (a) Cohen GB, Ren R, Baltimore D. Cell 1995; 80:237–248; (b) Liu X, Pawson T. Recent Prog<br />

Hormone Res 49:149–160; (c) Fry MJ, Panayotou G, Booker GW, Waterfield MD. Protein Sci 1993;<br />

2:1785–1797.<br />

269. Cantley LC, Songyang Z. J Cell Sci 1994; 18 (Suppl):121–126.<br />

270. Shahripour A, Para KS, Plummer MS, Lunney<br />

EA, Stankovic CJ, Holland DR, Rubin JR, Humblet<br />

C, Fergus JH, Marks JS, Saltiel AR, Sawyer TK.<br />

Bioorg Med Chem 1997; in press..<br />

271. Shahripour A, Plummer MS, Lunney EA, Para KS, Stankovic CJ, Rubin JR, Humblet C, Fergus<br />

JH, Marks JS, Herrera R, Hubbell SE, Saltiel AR, Sawyer TK. Bioorg Med Chem Lett 1996;<br />

6:1209–1214.<br />

272. Zhang Y-Z, Maclean D, McNamara DJ, Sawyer TK, Dixon JE, Biochemistry 1994; 33:2285–2290.<br />

273. Burke TR Jr, Kole HK, Roller PP. Biochem Biophys Res Comm 1994; 204:129–134.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_634.html (1 of 2) [4/9/2004 1:50:06 AM]


Document<br />

274. (a) Kavanaugh WM, Williams LT. Science 1994; 266:1862–1865; (b) Kavanaugh WM, Turck CW,<br />

Williams LT. Science 1995; 268:1177–1179.<br />

275. Eck MJ. <strong>Structure</strong> 1995; 3:421–424.<br />

276. (a) Thompson LA, Ellman JA. Chem Rev 1996; 96:555–600; (b) Patel DV, Gordon EM. <strong>Drug</strong> Disc<br />

Today 1996; 1:134–144; (c) Chen JK, Schreiber SL. Angew Chem Int Ed Engl 1995; 34:953–969; (d)<br />

Andrews P, Cody WL, Leonard DM, Sawyer TK. In: Pennington MW, Dunn BM, Eds. Peptide<br />

Synthesis and Purification Protocols. Vol. II. Humana Press, 1994:305–328; (e) Zuckermann RN. Curr<br />

Opinion Struct Biol 1993; 3:580–584; (f) Pavia MR, Sawyer TK, Moos WH. Bioorg Med Chem Lett<br />

1993; 3:387–396; (g) Jung G, Beck-Sickinger AG. Angew Chem Int Ed Engl 1992; 31:367–486.<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_634.html (2 of 2) [4/9/2004 1:50:06 AM]


Document<br />

Index<br />

A<br />

Acquired immunodeficiency virus (AIDS), 41, 56, 62, 65, 70<br />

Active site, 45, 48, 50, 54, 55, 56, 59, 61, 62, 63, 64, 65<br />

Acyclovir diphosphate, 154, 160, 164<br />

Alcohol dehydrogenase (ADH) 202-203<br />

catalytic mechanism, 233-234<br />

inhibitors, 231-232<br />

mutations, 234<br />

NADPH cofactor binding, 232-233<br />

relation to diabetic complications, 229-231<br />

structure<br />

active site, 233-235<br />

NADPH-bound form, 231-233<br />

ternary complex with zopolrestat, 235-239<br />

α-ketoamide transition state, 282<br />

Page 635<br />

Activate platelets, 247<br />

Activity, 41, 48, 50, 55, 56, 64, 65, 67, 68, 69, 70<br />

Addison's disease, 195<br />

Aldo-keto reductase superfamily, 240<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_635.html (1 of 2) [4/9/2004 1:50:42 AM]<br />

Aldose reductase (ALR2)


Document<br />

replacements, 563-565<br />

1-amidinopiperidine, 256, 257<br />

Amino acids<br />

C alpha, 128<br />

N-methyl, 128<br />

side chain constraints, 124-128<br />

converting enzyme, 120<br />

Animal models<br />

thrombosis of, 267<br />

Anthopleurins, 297, 300, 301, 302, 312, 313, 314<br />

Antibody-neuraminidase complexes, 470<br />

Anticoagulation, 247<br />

Anticoagulant protein, 257<br />

ALR2 (see Aldose reductase)<br />

Amantidine,<br />

462<br />

Amide bond<br />

Aminoalcohols 326<br />

Amphipathic alpha-helices (see Leucine zipper)<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_635.html (2 of 2) [4/9/2004 1:50:42 AM]<br />

Angiotensin, 321


Document<br />

Antirhinoviral agents<br />

capsid-binding compounds, 497-500<br />

clinical trials, 517-518<br />

drug sensitivity groups, 503<br />

resistance, 514-517<br />

structure-activity relationships, 502-514<br />

WIN compounds, 498-500<br />

Antithrombin III, 248<br />

eflornithine, 365-367<br />

melarsoprol, 365-366<br />

pentamidine, 365<br />

suramin, 365-366<br />

Argatroban, 251, 255<br />

Arginine<br />

boronate esters, 250<br />

guanidinium, 251<br />

Arrhythmias, 296, 298<br />

Aspartic proteinases<br />

Page 636<br />

Anti-influenza drugs, 462, 477-480<br />

Antitrypanosomal agents<br />

Antiviral<br />

agent, 41, 67<br />

Apparent mineralocorticoid excess (AME), 191<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_636.html (1 of 4) [4/9/2004 1:52:05 AM]


Document<br />

inhibitor binding, 323, 332<br />

inhibitors, 323<br />

mechanism, 328<br />

specificity, 333<br />

structure, 322<br />

transition state analogs, 323<br />

Atherosclerosis 395, 398<br />

Autoimmune<br />

disease, 395,398<br />

atherosclerosis, 395, 398<br />

insulin-dependant diabetes, 395, 398<br />

myasthenia gravis, 398<br />

rheumatoid arthritis, 395, 398<br />

systemic lupus erythematosus, 398<br />

disorders, 152<br />

Available chemicals database (ACD), 381<br />

B<br />

Bacteriorhodopsin, 131<br />

BCX-34, 166, 167<br />

Benzodiazepinone peptidomimetic, 571<br />

β-barrels, 248<br />

β-turn, 122, 124, 126, 127<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_636.html (2 of 4) [4/9/2004 1:52:05 AM]


Document<br />

Bicyclic peptidomimetic, 251<br />

Bidentate hydrogen bond, 252<br />

Binding<br />

cleft, 45, 48, 61, 65<br />

pocket, 56, 58, 60<br />

Bivalent inhibitors, 257<br />

Blood clot formation, 247<br />

Boroarginine, 261<br />

Boronate esters, 251<br />

Boronic acid analog, 250<br />

Bovine trypsin inhibitor mutants<br />

C<br />

phage display, 287<br />

positional requirements for Fxa inhibition, 286-288<br />

random mutagenesis, 287<br />

site specific mutagenesis, 285<br />

structure of complexes, 285<br />

Calcium regulators, 296<br />

cAMP generators, 296<br />

Cancer, 395<br />

Carbonyl reductase, 199<br />

Cardiotonic activity, 301, 304, 305, 306, 314<br />

Catalytic<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_636.html (3 of 4) [4/9/2004 1:52:05 AM]


Document<br />

site, 52, 55, 61<br />

triad, 247, 275<br />

Catechol O-methyltransferase<br />

active site, 349-350, 355-356<br />

catalysis, 345, 350-351<br />

inhibition mechanism, 356-358<br />

inhibitors, 351-353<br />

kinetics, 346<br />

physiological role, 344-345<br />

structure<br />

AdoMet (see S-Adenosyl-L-methionine)<br />

crystal structure, 347-350<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_636.html (4 of 4) [4/9/2004 1:52:05 AM]


Document<br />

[Catechol O-methyltransferase] drug complexes, 354-355<br />

Mg++ ion, 349-350<br />

S-adenosyl-L-methionine, 349<br />

sequence, 345, 347<br />

substrates, 345,346<br />

Cation-p site, 277, 282<br />

Ceriamide, 473<br />

Charge calculation, 381<br />

CHARMm, 133<br />

databases, 536<br />

selection algorithm, 536<br />

systems integration, 536<br />

virtual library production, 536<br />

Chemotherapy, 66<br />

Chimeric receptor, 126<br />

Chymotrypsin serine protease family, 248<br />

Coagulation<br />

cascade, 247, 265<br />

Page 637<br />

Chemical<br />

library, 142<br />

Chemi-informatics, 526, 535, 537<br />

Circular dichroism (CD), 438, 444<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_637.html (1 of 4) [4/9/2004 1:52:52 AM]


Document<br />

factors, 247<br />

Combination therapy, 62, 68, 69<br />

approach, 550-552<br />

chemistry, 141, 525, 537<br />

combinatorial explosion, 552<br />

drug lead source, 528, 529<br />

focusing, 530<br />

parallel synthesis, 528, 532<br />

refinement, 530<br />

robotic instrumentation, 528, 532<br />

scaffolds, 530, 532<br />

structure-activity relationship (SAR), 529<br />

pruning, 552<br />

Compound selection, 531<br />

chemi-informatics, 536<br />

drug properties, 535<br />

receptor fit, 536<br />

similarity, 536<br />

structure-activity relationship (SAR) models, 533, 536<br />

virtual libraries, 533, 536<br />

Computer programs<br />

BIOGRAF, 381<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_637.html (2 of 4) [4/9/2004 1:52:52 AM]<br />

Combinatorial


Document<br />

CHARMM, 542-544<br />

computational combinatorial ligand design (CCLD), 550-552<br />

DELPHI, 383<br />

DOCK, 379-383<br />

MCSS, 542-544<br />

Conformational<br />

change, 59, 60, 61, 70<br />

degrees of freedom, 252<br />

protein, 235-237<br />

Congestive heart failure, 295, 296, 298, 301, 314<br />

Cortiso, 193<br />

Cortisone, 193<br />

Cross linking, 137-138<br />

Crystallography<br />

cocktail soak approach, 377-379<br />

Crystal structure, 45, 54, 56, 59, 61, 64<br />

aldose reductase, 231-239<br />

other aldo-keto reductases, 240-241<br />

glyceraldehyde-3-phosphate dehydrogenase, 374-375<br />

phosphoglycerate kinase, 376-377, 388<br />

triosephosphate isomerase, 371-373<br />

Cutaneous T-cell lymphoma, 167<br />

Cyclic template, 252<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_637.html (3 of 4) [4/9/2004 1:52:52 AM]


Document<br />

Cyclotheonamide A (CtA), 254<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_637.html (4 of 4) [4/9/2004 1:52:52 AM]


Document<br />

Cytokines, 395-398, 412, 420<br />

alpha-beta, 396<br />

atherosclerosis, 398<br />

autoimmune disease and, 395, 98<br />

beta sandwich 396<br />

beta strands, 402<br />

beta-trefoils, 396<br />

4-helix bundle, 396<br />

growth hormone, 398<br />

immune system, 398<br />

immunomodulation, 398<br />

insulin-dependant diabetes, 398<br />

myasthenia gravis, 398<br />

network, 396<br />

nuclear magnetic resonance (NMR) and, 396<br />

production of, 395<br />

prolactin, 398<br />

receptors, 396<br />

rheumatoid arthritis, 398<br />

signalling <strong>by</strong>, 396<br />

structures, 396<br />

systemic lupus erythematosus, 398<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_638.html (1 of 4) [4/9/2004 1:53:57 AM]<br />

Page 638


Document<br />

D<br />

ddI, 152<br />

tumour necrosis factor, 398<br />

x-ray crystallography and, 396<br />

9-deazaguanine, 161, 163, 164, 165, 167<br />

Diabetic complications, 229-231<br />

Digoxin,295, 296, 297, 300<br />

Dihydropteridine reductase, 197, 202<br />

Diversity<br />

directed, 536<br />

of candidate ligands, 542,555<br />

dNTP binding, 56, 61<br />

<strong>Drug</strong>, 43, 45, 53, 55, 56, 60, 62, 63, 65, 66, 67, 68, 69, 70<br />

binding affinity prediction, 555<br />

combinatorial, 550-552<br />

design, 55, 66, 70, 402<br />

docking, 241-242, 379-383, 542-544<br />

fragment approach, 541<br />

inhibition, 70<br />

lead<br />

discovery, 377-384<br />

optimization, 384-387<br />

linked-fragment approach, 378<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_638.html (2 of 4) [4/9/2004 1:53:57 AM]


Document<br />

resistance, 56, 66, 68, 70<br />

scoring, 382-383<br />

solvation effects, 544-547<br />

<strong>Drug</strong> properties, 527, 535<br />

absorption, 535<br />

excretion, 535<br />

metabolism, 535<br />

refinement, 535, 536, 537<br />

structure-<strong>based</strong> design, 527<br />

DTic (tetrahydroisoquinnoline carboxylic acid), 124<br />

E<br />

Energy<br />

F<br />

minimization, 132-133<br />

with CHARMM, 542-544<br />

Fab, 45, 49, 51, 56, 58, 59, 69<br />

Factor Xa<br />

active site blocked (DEGR-Xa), 267, 269<br />

active site substrate sequences, 271<br />

modeled inhibitor complexes<br />

amidinoaryls, 279<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_638.html (3 of 4) [4/9/2004 1:53:57 AM]<br />

DUP714, 250


Document<br />

antistasin peptides, 281<br />

cyclotheonamide, 284<br />

dansyl-Glu-Glu-Arg-CMK, 280<br />

DX-9065a, 276<br />

SEL-2711, 283<br />

natural inhibitors of<br />

AcAP's, 273-274<br />

antistasin, 266, 268, 270, 271-272<br />

ecotin, 268, 270, 274, 288<br />

tick anticoagulant peptide (TAP), 266, 268, 270, 272-273<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_638.html (4 of 4) [4/9/2004 1:53:57 AM]


Document<br />

structure and function, 267-269, 274<br />

synthetic inhibitors of<br />

amidinoaryls, 277-279<br />

antistasin peptides, 280-282<br />

bisamidines, 277<br />

BPTI mutants, 285-288<br />

cyclotheonamide, 282<br />

dansyl-Glu-Glu-Arg-CMK, 280<br />

DX-9065a, 275-277<br />

peptidyl argininals, 288<br />

PPACK, 275<br />

SEL-2711, 282<br />

Factor XIII, 247<br />

Factor XIIa, 119<br />

Feline immunodeficiency virus (FIV), 441<br />

Fibrin, 247<br />

Fibrinogen, 247<br />

Fibrinopeptide A, 250, 261<br />

Flavonoid, 473<br />

Fluoroketone analogs, 327<br />

Fragment approach (see Computer programs, MCSS, CCLD)<br />

Page 639<br />

[Factor Xa] TFPI, 270-271, 285, 287<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_639.html (1 of 4) [4/9/2004 1:55:20 AM]


Document<br />

fragment-<strong>based</strong> programs, 541-542<br />

Fourier maps, 154, 161, 166<br />

G<br />

Geminal diol analogs, 327<br />

Genomic data, 525, 527<br />

GG167 (see Neu5Ac2en,4-guanidino)<br />

Glucopyranoside<br />

peptidomimetic, 571<br />

Glyceraldehyde-3-phosphate dehydrogenase (GAPDH)<br />

catalysis, 372, 374<br />

crystal structure, 374-375<br />

human, 374-375<br />

inhibitors, 381-382, 384-387<br />

Leishmania mexicana, 375-376, 385<br />

sequence, 374<br />

Glucocorticoid 193<br />

G-protein coupled receptors (GPCR), 592-594<br />

H<br />

Heart attack, 247<br />

Hemagglutinin, 459, 460, 462, 463, 464, 477<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_639.html (2 of 4) [4/9/2004 1:55:20 AM]<br />

Glycol analogues 326<br />

GRID maps, 475<br />

Hematophageous organisms


Document<br />

Ancylostoma caninum, 273<br />

Haementeria officinalis, 266<br />

Ornithidorous moubata, 266<br />

Hemopexin domain, 172-174<br />

Hemostasis, 247<br />

Heparin, 248 cofactor II, 248<br />

High throughput screening, 530<br />

Hirudin, 251<br />

HOE 140, 127-128<br />

Homology<br />

model building, 275-276,278<br />

models, 527<br />

scaffold development, 532<br />

structure-<strong>based</strong> design, 527<br />

sequence, 176<br />

structure, 176-177<br />

Hormone therapy, 206<br />

Human immunodeficiency virus (HIV), 441<br />

protease<br />

cleavage sites, 7<br />

flexibility, 7-8<br />

inhibitors, 586-587 AG1284, 22-27<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_639.html (3 of 4) [4/9/2004 1:55:20 AM]


Document<br />

[Human immunodeficiency virus]<br />

cyclic ureas, 21-22<br />

hydroxycoumarins, 27-28<br />

indinavir, 9, 15-17, 28-29<br />

inversion of binding mode, 24<br />

nelfinavir, 17-21, 29-30<br />

nonpeptidic, 17-28<br />

peptidic, 9-10<br />

peptidomimetic, 10-19<br />

ritonavir, 13-16, 28-29<br />

saquinavir, 10-13, 28-29<br />

symmetric binding, 13-15, 21-22<br />

mutations, 28-32<br />

resistance<br />

primary mutations, 29-30<br />

secondary mutations, 30-32<br />

sequential passage, 28<br />

vitality factor, 32<br />

structure<br />

active site, 5-7<br />

three dimensional, 3-5<br />

Wat 301, 6, 7, 10, 13, 16, 17, 21, 22, 24, 27<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_640.html (1 of 4) [4/9/2004 1:56:12 AM]<br />

Page 640


Document<br />

integrase<br />

amino acid sequence, 90-91<br />

amino terminal domain, 92, 102<br />

biochemical properties, 85-88<br />

biophysical properties, 88, 92<br />

carboxyl terminal domain, 89, 92<br />

structure of, 102-103<br />

catalytic core domain, 88-89<br />

biophysical properties of, 92-93<br />

mutation of hydrophobic residues of, 102-103<br />

structure of, 93-102<br />

comparison to toher<br />

polynucleotidyl<br />

transferases, 96-100<br />

conserved acidic residues in<br />

active site, 95-96<br />

dimer, 100-102<br />

domain structure, 88-92<br />

inhibitors, 103-109<br />

3'-azido-3'-deoxy-thymidine (AZT), 107-108<br />

common pharmacophore of, 104-107<br />

curcumin, 107, 111<br />

design of, 110-112<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_640.html (2 of 4) [4/9/2004 1:56:12 AM]


Document<br />

overview, 103-104, 108-109<br />

rationale for, 85<br />

Human prothrombin fragment F1, 261<br />

Hydroxamate, 172, 182-184<br />

Hydroxysteroid dehydrogenases, 191<br />

7α, 197, 199<br />

11β, 191-194, 203<br />

17β, 193-199, 205-207<br />

20β, 197, 199<br />

Hypertension, 193<br />

I<br />

Inflammation, 395<br />

Influenza virus<br />

antigenic variation, 462, 468-470<br />

classification, 459<br />

drug resistance, 478, 479<br />

inhibition, 478<br />

replication cycle, 461<br />

vaccines, 463<br />

Inhibition, 50, 54, 60, 61, 65, 69, 70<br />

mechanism, 65<br />

Inhibitors, 41, 43, 45, 48, 50, 55, 56, 58, 59, 60, 61, 62, 63, 65, 66, 67, 68, 69, 70<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_640.html (3 of 4) [4/9/2004 1:56:12 AM]<br />

Hydrophobic collapse, 252


Document<br />

design, 358-359<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_640.html (4 of 4) [4/9/2004 1:56:12 AM]


Document<br />

[Inhibitors]<br />

2-((3,4-dihydroxy-2-nitrophenyl)vinyl)phenyl-ketone, 352-353, 356<br />

entacapone, 352-353, 360<br />

first generation inhibitors, 351-352<br />

tolcapone, 352-353, 360<br />

Insulin-dependent diabetes 395,398<br />

Interactions<br />

electrostatic, 544<br />

hydrophobic, 545<br />

van der Waals, 545<br />

Interferon<br />

α, 435, 439, 440<br />

β, 435<br />

activity, 442<br />

cytotoxicity, 439, 443<br />

receptor, 441<br />

structural studies, 443, 444<br />

clinical uses, 436<br />

definition, 435<br />

γ, 435<br />

activity, 448, 449<br />

antibody studies, 445<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_641.html (1 of 4) [4/9/2004 1:56:34 AM]<br />

Page 641


Document<br />

receptor, 446, 447, 450<br />

intron A, 436<br />

roferon, 436<br />

side effects, 436<br />

sub types, 435, 436<br />

synthetic peptide studies, 448<br />

signal transduction, 451<br />

structural studies, 449-451<br />

synthetic peptide studies, 446, 448<br />

τ, 435, 439, 440,<br />

activity, 441-443<br />

antibody studies, 442<br />

expression, 442, 443<br />

receptor, 441<br />

signal transduction, 443<br />

structural studies, 443, 44<br />

synthetic peptide studies, 441, 442, 444<br />

ω, 435, 440<br />

Interleukin-1<br />

accessory protein, 398, 401<br />

affinity for receptors, 401<br />

α, 398, 399, 401, 402, 406, 420, 421, 423, 427, 428<br />

alternatively spliced, 399<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_641.html (2 of 4) [4/9/2004 1:56:34 AM]


Document<br />

antagonistic activity, 416<br />

auto-antibodies and, 401<br />

autoimmune disease and, 401<br />

barrel, 402, 404<br />

β, 398, 399, 401, 402, 420, 421, 423, 427, 428<br />

barrel, 406, 409<br />

bulge, 415-416<br />

converting enzyme (ICE), 399, 402, 412<br />

hairpins, 405<br />

strands and, 402<br />

beta-trefoil fold, 402, 403, 404, 407<br />

binding, 412-416<br />

catalytic<br />

affinity, 421<br />

epitope, 410<br />

pocket, 412<br />

activity of, 412<br />

diad, 412<br />

cysteine protease, 399, 412<br />

enzyme mechanism, 412<br />

epitopes of, 405, 413-416, 418<br />

expression of, 399<br />

fibroblast growth factors, 402<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_641.html (3 of 4) [4/9/2004 1:56:34 AM]


Document<br />

gene duplication, 398<br />

glycosylation of, 399, 409<br />

GTPase proteins and, 401<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_641.html (4 of 4) [4/9/2004 1:56:34 AM]


Document<br />

[Interleukin-1]<br />

homology with CED-3 protein, 412<br />

hydrophobic patch, 406<br />

hydrophobic residues, 404<br />

immunoglobulin superfamily and, 399<br />

Kunitz family, 402<br />

leukemia, 395<br />

location of, 399<br />

low molecular weight antagonists, 421-427<br />

monoclonal antibodies and, 421<br />

monocyte phagocytes <strong>by</strong>, 399<br />

mutational studies of, 407, 409<br />

N-terminal extension<br />

nuclear magnetic resonance (NMR), 402, 404<br />

overexpression of, 419<br />

peptide fragments, 416<br />

peptomimetics, 412<br />

precursor, 399<br />

form, 399<br />

receptor, 398, 402, 420<br />

antagonist, 398, 399, 401, 407, 420, 421<br />

binding, 404-405, 406, 407-410, 412-416<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_642.html (1 of 4) [4/9/2004 1:57:29 AM]<br />

Page 642


Document<br />

recombinant, 421<br />

sequence identity of, 398<br />

signal transduction and, 401<br />

site-directed mutagenesis, 412-416<br />

structure of, 401-412, 419<br />

synthesis of, 395, 419<br />

substrate, 399<br />

specificity, 412<br />

subunits of, 412<br />

systemic lupus erythematosus, 398<br />

therapeutic strategies and, 412<br />

tumor necrosis factor and, (TNF) 421<br />

type I, 399, 401, 420<br />

type 2, 399, 401, 420<br />

x-ray crystallography, 401-412<br />

Ion-channel modulators, 296, 297, 301<br />

Ischemia, 296, 297<br />

K<br />

Kinins, 119<br />

turn propensity of, 123<br />

bradykinin, 119<br />

Kampo drug, 473<br />

Ketomethylene pseudo peptide bond, 260<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_642.html (2 of 4) [4/9/2004 1:57:29 AM]


Document<br />

Kininogen, 119<br />

high molecular weight, 119<br />

low molecular weight, 119<br />

Kinin receptors, 120<br />

agonist binding site, 131-133<br />

antagonists of, 124<br />

antagonist site, 137-138<br />

B1 subtype, 120<br />

B2 subtype, 119, 120<br />

chimeras and, 138, 139<br />

mutagenesis of, 133-134<br />

Kallidin, 119<br />

molecular dynamics of, 123<br />

pharmacology of, 121<br />

solution conformation of, 121<br />

Kunitz domains, 271, 273, 282, 285<br />

Kyte-Doolittle, 131<br />

L<br />

Lactam amide, 222-223<br />

Leucine zipper<br />

propane functional map, 545-546<br />

stereo view, 546<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_642.html (3 of 4) [4/9/2004 1:57:29 AM]


Document<br />

yeast transcriptional activator protein GCN4, 545-546<br />

Leydig cells, 193<br />

Licorice, 193, 195-196<br />

Ligand docking, 133<br />

Lock-and-key, 159<br />

M<br />

Malignant<br />

tissue, 214<br />

transformation, 214<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_642.html (4 of 4) [4/9/2004 1:57:29 AM]


Document<br />

Matrix-metalloproteinase (MMP), 171-186<br />

fibroblast collagenase, 171-173, 176, 179, 183-184<br />

gelatinase, 171-173, 184<br />

matrilysin, 171-175, 182-184<br />

neutrophil collagenase, 171-175, 183-184<br />

stromelysin, 171-173, 184<br />

Medicinal leech, 257<br />

Metal requirement, 88, 89<br />

Mineralocorticoid receptor, 193<br />

Molecular<br />

fragments<br />

aliphatic, 542<br />

aromatic, 542<br />

charged, 542<br />

polar, 542<br />

modelling, 182-183<br />

MuA transposase<br />

structural comparison to integrase, 97-100<br />

Mutation, 62, 67, 68, 69<br />

Myristic acid, 214<br />

N<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_643.html (1 of 4) [4/9/2004 1:58:26 AM]<br />

Page 643<br />

Myocardial infarction, 247


Document<br />

NAD, 199, 201<br />

NADP, 199, 201<br />

NADPH, 199-201<br />

active site, 470-472<br />

carbohydrate structure, 467<br />

dimensions, 465<br />

enzyme function, 464, 473<br />

inhibitors, 472-473, 476<br />

molecular weight, 465<br />

morphology, 464<br />

topology, 466<br />

Neu5Ac2en, 472, 473, 474, 477, 478, 480<br />

4-guanidino, 475, 476, 477, 479<br />

4-amino, 475, 476<br />

Nonnecleoside ingibitor binding pocket (NNIBP), 56, 58, 59, 60, 61, 62, 63, 66, 67<br />

NAPAP, 255, 261<br />

Napthalenesulfonyl, 257<br />

Neuraminidase<br />

Nonnucleoside reverse transcriptase inhibitors (NNRTI), 41, 45, 48, 49, 50, 56, 58, 59, 60, 61, 62, 63,<br />

64, 65, 66, 67, 68, 69, 70<br />

Nonnucleoside, 41, 45, 56<br />

inhibitor, 45, 56<br />

inhibitor binding pocket, 56<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_643.html (2 of 4) [4/9/2004 1:58:26 AM]


Document<br />

RT inhibitor, 41, 56<br />

Nonpeptide<br />

angiotensin agonist, 586-587<br />

antagonist<br />

angiotensin, 586-587<br />

cholecystokinin, 586-587<br />

endothelin, 586-587<br />

gastrin-releasing hormone, 586-587<br />

glucagon, 586-587<br />

neurokinin, 586-587<br />

neuropeptide Y, 588-589<br />

neurotensin, 586-587<br />

oxytocin, 586-587<br />

drug discovery, 559, 570, 586-587<br />

vasopressin, 586-587<br />

NPC 17731, 124<br />

NPC 18325, 137<br />

NPC 567, 124<br />

Nuclear magnetic resonance (NMR), 402, 404, 407<br />

Nucleic acid, 48, 51, 54, 64, 65, 69<br />

Nucleophile, 221<br />

Nucleoside, 41, 50, 52, 53, 56<br />

reverse transcriptase (RT) inhibitor, 41, 50, 52, 53, 54, 55, 56, 65<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_643.html (3 of 4) [4/9/2004 1:58:26 AM]


Document<br />

Nucleotide, 51, 53, 54, 55, 69<br />

O<br />

binding domain, 200<br />

Oxyanion hole, 248, 250<br />

P<br />

Parkinson's disease<br />

Peptides<br />

disease, 359<br />

therapy, 359-360<br />

synthetic<br />

structure-function studies in, 437, 438, 440-442, 444, 446, 448<br />

Peptidomimetic<br />

drug discovery, 591-598, 613<br />

Page 644<br />

Octahydroindole carboxylic acid (Oic), 124<br />

(see also Protease targets, Receptor targets, Signal transduction protein targets, <strong>Structure</strong>-<strong>based</strong><br />

drug design)<br />

Peptidomimetics, 251<br />

peptide and nonpeptide models, 593-594<br />

copper complexes as integrase inhibitors, 107<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_644.html (1 of 4) [4/9/2004 1:59:33 AM]<br />

PGK (see<br />

Phosphoglycerate<br />

kinase)<br />

Pharmacophore, 225, 560<br />

Phenanthroline


Document<br />

Phosphoglycerate kinase (PGK)<br />

catalysis, 376-377<br />

crystal structure, 376-377, 388<br />

human, 377<br />

sequence, 377<br />

Trypanosoma brucei, 376-377<br />

Phosphorylysis, 151<br />

Phosphotransfer, 222<br />

Platelet aggregation, 247<br />

active site, 45, 48, 50, 54, 55, 56, 59, 61, 62, 63, 64<br />

catalytic site, 52, 55, 61<br />

Polymerization, 41, 48, 50, 55, 61, 67, 70<br />

mechanism, 56<br />

Polynucleotidyl transferases, 96-100<br />

Positive<br />

inotropes, 295, 296, 297, 298, 300, 309, 310, 314<br />

inotropy, 295, 297, 298, 305, 314<br />

PPACK, 250, 261<br />

Primer grip, 48, 61<br />

Proliferative diseases<br />

asthma, 214<br />

Phosphostatine analogs, 327<br />

Polymerase, 41, 45, 48, 50, 52, 54, 55, 56, 59, 61, 62, 63, 64, 65, 67, 69<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_644.html (2 of 4) [4/9/2004 1:59:33 AM]


Document<br />

atherosclerosis, 214<br />

fibrosis, 214<br />

osteo arthritis, 214<br />

psoriasis, 214<br />

restenosis, 214<br />

rheumatoid arthritis, 214<br />

septic shock, 214<br />

Protein C, 247<br />

Protease targets<br />

angiotensin-converting enzyme (ACE) inhibitors, 576, 607, 610<br />

aspartyl proteases, 595-596<br />

classes, 567-568<br />

cysteinyl proteases, 605<br />

inhibitors<br />

elastase, 576, 597, 606<br />

gelatinase, 576<br />

human immunodeficiency virus (HIV) protease, 576-578, 595-596, 598, 600-602<br />

interleukin-converting enzyme (ICE), 576, 597, 607-608<br />

stromelysin, 576, 596, 609-612<br />

thermolysin, 597, 610<br />

thrombin, 576, 578-579, 596, 604<br />

metalloproteases, 597-598, 607, 609-612<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_644.html (3 of 4) [4/9/2004 1:59:33 AM]


Document<br />

[Protease targets]<br />

serinyl proteases, 603<br />

x-ray structures<br />

apoprotein and complexes, 595-598<br />

Protein data bank, 173<br />

activation loop, 218, 219<br />

calmodulin dependent, 219<br />

caseine (CK-1), 218, 220, 222, 224<br />

catalytic<br />

core, 214, 219-221<br />

loop, 217, 221<br />

cyclin dependent (CDK-2), 214, 215, 218-219<br />

EGFR, 223<br />

insulin receptor (IRK), 214-216, 218-220, 222<br />

MAP, 214, 218-219<br />

phosphorylase, 218-220, 222, 224<br />

protein A (cAPK) 214-216, 218-220, 222-225<br />

protein C, 224<br />

phosphorylation, 217, 218<br />

http://legacy.netlibrary.com/nlreader/nlReader.dll?bookid=12640&filename=Page_645.html [4/9/2004 2:00:20 AM]<br />

Page 645<br />

Protein<br />

kinase

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!