1-s2.0-s1074552114002919-main

13
Chemistry & Biology Review Small-Molecule Inhibitors of Protein-Protein Interactions: Progressing toward the Reality Michelle R. Arkin, 1, * Yinyan Tang, 1 and James A. Wells 1 1 Department of Pharmaceutical Chemistry, School of Pharmacy, University of California, San Francisco, San Francisco, CA 94158, USA *Correspondence: [email protected] http://dx.doi.org/10.1016/j.chembiol.2014.09.001 The past 20 years have seen many advances in our understanding of protein-protein interactions (PPIs) and how to target them with small-molecule therapeutics. In 2004, we reviewed some early successes; since then, potent inhibitors have been developed for diverse protein complexes, and compounds are now in clinical trials for six targets. Surprisingly, many of these PPI clinical candidates have efficiency metrics typical of ‘‘lead-like’’ or ‘‘drug-like’’ molecules and are orally available. Successful discovery efforts have integrated multiple disciplines and make use of all the modern tools of target-based discovery—structure, computation, screening, and biomarkers. PPIs become progressively more challenging as the interfaces become more complex, i.e., as binding epitopes are displayed on primary, secondary, or tertiary structures. Here, we review the last 10 years of progress, focusing on the properties of PPI inhibitors that have advanced to clinical trials and prospects for the future of PPI drug discovery. Protein-protein interactions (PPIs) represent a vast class of ther- apeutic targets both inside and outside the cell. PPIs are central to all biological processes and are often dysregulated in disease. Despite the importance of PPIs in biology, this target class has been extremely challenging to convert to therapeutics. Twenty years ago, PPIs were deemed ‘‘intractable.’’ High-resolution structures in the 1980s to 1990s showed that PPI interfaces are generally flat and large (roughly 1,000–2,000 A 2 per side) (Hwang et al., 2010), in stark contrast to the deep cavities that typically bind small molecules (300–500 A 2 )(Fuller et al., 2009). Unlike enzymes or G protein-coupled receptors (GPCRs), nature did not offer simple small molecules that can start a chemical discovery process, and high-throughput screening (HTS) had not provided validated hits. Between 1995 and 2005, hopeful signs were emerging. A clinically approved integrin antagonist (tirofiban) and natural products like taxanes, rapamycin, and cyclosporine inspired confidence that PPIs could be modulated by small molecules. Mutational analysis of protein interfaces showed that not all res- idues at the PPI interface were critical, but rather that small ‘‘hot spots’’ conferred most of the binding energy (Arkin and Wells, 2004; Clackson and Wells, 1995). Hot spots tended to cluster at the center of the interface, to cover an area comparable to the size of a small molecule, to be hydrophobic, and to show conformational adaptivity. These features suggested that at least some PPIs might have small-molecule-sized patches that could dynamically adjust to bind a drug-like molecule. By 2005, about a half-dozen small molecules had been reported to bind with the affinities one would expect for drug leads, at binding sites defined by high-resolution structures (Wells and McClendon, 2007). In parallel, computation and chemical technologies were being developed that might be well suited to PPIs. For instance, fragment-based lead discovery (FBLD) has had a particularly strong impact. FBLD used biophysical methods, including crys- tallography, surface plasmon resonance, and nuclear magnetic resonance (NMR), or disulfide trapping (tethering) to identify low-molecular-weight, low-complexity molecules that bound weakly to subsites on the protein surface (Erlanson et al., 2004; Hajduk and Greer, 2007; Winter et al., 2012). The last decade has seen amazing progress in tackling chal- lenging PPI targets with synthetic molecules. More than 40 PPIs have now been targeted (Basse et al., 2013; Higueruelo et al., 2009; Labbe ´ et al., 2013), and several inhibitors have reached clinical trials. With this advance, it is important to recon- sider the distinction between ligandability (‘‘druggability’’) and our ability to convert PPI inhibitors into drugs. Historically, PPI inhibitors have been larger and more hydrophobic than typical orally available drugs (Wells and McClendon, 2007). Two commonly used metrics to assess the drug-like quality of a compound (or to compare a series of compounds) are ligand efficiency (LE; DG/HA) and lipophilic ligand efficiency (LLE; pIC 50 – logD or logP) (Hopkins et al., 2014). The LE for small- molecule inhibitors of PPIs have hovered around 0.24, whereas a LE of 0.3 or higher is desired. Values of LLE >5 are considered favorable for in vivo activity. Encouragingly, recent PPI inhibitors are approaching these ‘‘drug-like’’ values for several targets (see below). Even inhibitors with properties outside average ranges for oral drugs have been made orally bioavailable. Clinically suc- cessful PPI inhibitors may therefore expand our understanding of the types of molecules that can be made into drugs. Also during the past fifteen years, there has been very prom- ising progress with designing peptides that target PPIs and show promising cell-based (and even in vivo) activities (Azzarito et al., 2013; Bernal et al., 2010; Boersma et al., 2012; Chang et al., 2013; DeLano et al., 2000; Gavenonis et al., 2014). While these approaches are outside the scope of the current review, they represent a parallel strategy that can also inform small- molecule design. Although PPIs come in many shapes and sizes, most of the clinical-stage inhibitors target PPIs in which the hot-spot resi- dues are concentrated in small binding pockets (250–900 A 2 ) (Basse et al., 2013; Smith and Gestwicki, 2012) and partner proteins are characterized by short primary sequences at the interface (London et al., 2013; Perkins et al., 2010). At one 1102 Chemistry & Biology 21, September 18, 2014 ª2014 Elsevier Ltd All rights reserved

Upload: hashem-el-marimey

Post on 14-Nov-2015

8 views

Category:

Documents


0 download

DESCRIPTION

chem

TRANSCRIPT

  • ve

    dnocvale

    es

    typically bind small molecules (300500 A ) (Fuller et al.,

    dynamically adjust to bind a drug-like molecule. By 2005, about

    inhibitors have been larger and more hydrophobic than

    et al., 2013; Bernal et al., 2010; Boersma et al., 2012; Changbeing developed that might be well suited to PPIs. For instance,

    fragment-based lead discovery (FBLD) has had a particularly

    strong impact. FBLD used biophysical methods, including crys-

    Although PPIs come in many shapes and sizes, most of the

    clinical-stage inhibitors target PPIs in which the hot-spot resi-

    dues are concentrated in small binding pockets (250900 A2)tallography, surface plasmon resonance, and nuclear magnetic

    resonance (NMR), or disulfide trapping (tethering) to identify

    low-molecular-weight, low-complexity molecules that bound

    (Basse et al., 2013; Smith and Gestwicki, 2012) and partner

    proteins are characterized by short primary sequences at the

    interface (London et al., 2013; Perkins et al., 2010). At onea half-dozen small molecules had been reported to bind with

    the affinities one would expect for drug leads, at binding sites

    defined by high-resolution structures (Wells and McClendon,

    2007). In parallel, computation and chemical technologies were

    et al., 2013; DeLano et al., 2000; Gavenonis et al., 2014). While

    these approaches are outside the scope of the current review,

    they represent a parallel strategy that can also inform small-

    molecule design.2009). Unlike enzymes or G protein-coupled receptors (GPCRs),

    nature did not offer simple small molecules that can start a

    chemical discovery process, and high-throughput screening

    (HTS) had not provided validated hits.

    Between 1995 and 2005, hopeful signs were emerging. A

    clinically approved integrin antagonist (tirofiban) and natural

    products like taxanes, rapamycin, and cyclosporine inspired

    confidence that PPIs could be modulated by small molecules.

    Mutational analysis of protein interfaces showed that not all res-

    idues at the PPI interface were critical, but rather that small hot

    spots conferred most of the binding energy (Arkin and Wells,

    2004; Clackson and Wells, 1995). Hot spots tended to cluster

    at the center of the interface, to cover an area comparable to

    the size of a small molecule, to be hydrophobic, and to show

    conformational adaptivity. These features suggested that at least

    some PPIs might have small-molecule-sized patches that could

    typical orally available drugs (Wells and McClendon, 2007).

    Two commonly used metrics to assess the drug-like quality of

    a compound (or to compare a series of compounds) are ligand

    efficiency (LE; DG/HA) and lipophilic ligand efficiency (LLE;

    pIC50 logD or logP) (Hopkins et al., 2014). The LE for small-

    molecule inhibitors of PPIs have hovered around 0.24, whereas

    a LE of0.3 or higher is desired. Values of LLE >5 are consideredfavorable for in vivo activity. Encouragingly, recent PPI inhibitors

    are approaching these drug-like values for several targets (see

    below). Even inhibitors with properties outside average ranges

    for oral drugs have been made orally bioavailable. Clinically suc-

    cessful PPI inhibitorsmay therefore expand our understanding of

    the types of molecules that can be made into drugs.

    Also during the past fifteen years, there has been very prom-

    ising progress with designing peptides that target PPIs and

    show promising cell-based (and even in vivo) activities (AzzaritoSmall-Molecule Inhibitors oInteractions: Progressing to

    Michelle R. Arkin,1,* Yinyan Tang,1 and James A. Wells11Department of Pharmaceutical Chemistry, School of Pharmacy, Uni*Correspondence: [email protected]://dx.doi.org/10.1016/j.chembiol.2014.09.001

    The past 20 years have seen many advances in our unhow to target themwith small-molecule therapeutics. Ipotent inhibitors have been developed for diverse prtrials for six targets. Surprisingly, many of these PPIlead-like or drug-like molecules and are orally amultiple disciplines andmake use of all themodern tooscreening, and biomarkers. PPIs become progressivcomplex, i.e., as binding epitopes are displayed on primthe last 10 years of progress, focusing on the propertiand prospects for the future of PPI drug discovery.

    Protein-protein interactions (PPIs) represent a vast class of ther-

    apeutic targets both inside and outside the cell. PPIs are central

    to all biological processes and are often dysregulated in disease.

    Despite the importance of PPIs in biology, this target class has

    been extremely challenging to convert to therapeutics. Twenty

    years ago, PPIs were deemed intractable. High-resolution

    structures in the 1980s to 1990s showed that PPI interfaces

    are generally flat and large (roughly 1,0002,000 A2 per side)

    (Hwang et al., 2010), in stark contrast to the deep cavities that21102 Chemistry & Biology 21, September 18, 2014 2014 Elsevier LtChemistry & Biology

    Review

    f Protein-Proteinward the Reality

    rsity of California, San Francisco, San Francisco, CA 94158, USA

    erstanding of protein-protein interactions (PPIs) and2004, we reviewed some early successes; since then,tein complexes, and compounds are now in clinicallinical candidates have efficiency metrics typical ofilable. Successful discovery efforts have integrateds of target-based discoverystructure, computation,ly more challenging as the interfaces become moreary, secondary, or tertiary structures. Here, we reviewof PPI inhibitors that have advanced to clinical trials

    weakly to subsites on the protein surface (Erlanson et al., 2004;

    Hajduk and Greer, 2007; Winter et al., 2012).

    The last decade has seen amazing progress in tackling chal-

    lenging PPI targets with synthetic molecules. More than 40

    PPIs have now been targeted (Basse et al., 2013; Higueruelo

    et al., 2009; Labbe et al., 2013), and several inhibitors have

    reached clinical trials. With this advance, it is important to recon-

    sider the distinction between ligandability (druggability) and

    our ability to convert PPI inhibitors into drugs. Historically, PPId All rights reserved

  • extreme, the peptide epitope is recapitulated by a linear peptide

    epitope 14 amino acids long; there are four clinical-stage exam-

    ples of small-molecule mimetics of these epitopes. PPIs that

    contain a single unit of secondary structure, such as an alpha he-

    lix, binding to a hydrophobic groove have also proven tractable

    Figure 1. The Complexity of the PPI Interface Affects DruggabilityPPIs can be classified bywhether one side of the interface consists of a primary(linear) protein sequence (green), a single region of secondary structure (suchas an alpha helix, yellow), or multiple sequences requiring tertiary structure(red). There are fewer examples of small-molecule inhibitors of PPIs as theinterface becomes more complex (from primary, to secondary, to tertiary epi-topes). Structures shown are BRDt/histone (green; Protein Data Bank [PDB]:2WP1), MDM2/p53 (yellow; PDB: 1YCR), and IL-2/IL-2Ra (red; PDB: 1Z92).

    Chemistry & Biology

    Reviewto small-molecule inhibition. Globular interfaces, requiring ter-

    tiary structure on both sides of the PPI, have fewer published

    successes. In developing guidelines for PPI inhibitor discovery,

    it is instructive to consider each clinical success story in the

    context of the type of interface (Figure 1).

    Primary Peptide Epitopes: Short, Continuous, LinearPeptidesOne of the first examples of a clinically successful PPI inhibitor is

    tirofiban,whichbinds to the integrin IIbIIIa. Tirofibanwasdesigned

    to mimic the linear tripeptide Arg-Gly-Asp (RGD), the epitope of

    fibrinogen that binds to the I-like domain of IIbIIIa (Hartman

    et al., 1992). While this small, linear peptide epitope seems like a

    special case of PPI, this motif is observed in turns, at protein

    termini, and in unstructured regions (Perkins et al., 2010). Unsur-

    prisingly, these types of interfaces have been particularly

    amenable to inhibition by small molecules, and four such targets

    have recently entered clinical trials, which we discuss below.

    LFA1

    Leukocyte function-associated antigen-1 (LFA-1) has been the

    target of drug discovery efforts aimed at reducing inflammatory

    immune responses. LFA-1 is a beta2 integrin involved in T cell

    activation and adhesion (Ford and Larsen, 2009) through binding

    to its ligand ICAM1. LFA-1 is a heterodimer consisting of an alpha

    chain (CD11a/aL) and a beta chain (CD18/b2). ICAM1 binds to an

    inserted domain (I domain) located on CD11a. CD18 also has an

    I-like domain that is located near the I domain. Three types of

    drugs have been in the clinic for LFA-1. The anti-LFA-1 antibody

    efalizumab was approved for psoriasis in 2003, but it was with-

    drawn in 2009 due to rare but fatal viral infections associated

    Chemistry & Biology 2with immunosuppression (Schwab et al., 2012). Small molecules

    that bound to an allosteric site on the I domain were in clinical

    trials for psoriasis, but no development has been reported since

    2010 (Guckian and Scott, 2013). Lifitegrast (SAR1118), however,

    recently completed its registrational phase 3 trial, and a new

    drug application will be filed in early 2015 (Shire, 2014).

    Lifitegrast was discovered at Sunesis, building from a series

    originally published by Genentech (Gadek et al., 2002), and

    was developed clinically by SARcode/Shire (Zhong et al.,

    2012). There are no crystal structures of lifitegrast or its analogs

    bound to LFA-1. The mechanism of action of the compound is

    still under debate; it either binds directly to the ICAM site on

    the I domain (Keating et al., 2006) or to the related site on the

    I-like domain (Shimaoka et al., 2003) (Figure 2), analogous to

    the binding of tirofiban. SARcode developed lifitegrast for the in-

    flammatory disorder dry eye syndrome, the most common eye

    disease in humans (Schaumberg et al., 2003). SAR1118 has an

    optimal pharmacokinetic profile for ocular formulation, with low

    systemic exposure and high aqueous solubility. Since this indi-

    cation does not require systemic exposure, lifitegrast could

    achieve the efficacy that eluded I domain antagonists (Guckian

    and Scott, 2013) and avoid the toxicity that led to efalizumabs

    withdrawal (Schwab et al., 2012).

    IAPs

    Inhibitor of apoptosis proteins (IAPs), including cIAP1, cIAP2,

    and XIAP, are important regulators of cell fate, including

    apoptotic cell death and immunity (Dubrez et al., 2013; Fulda

    and Vucic, 2012). The IAPs have PPI interaction domains called

    BIR domains and a RING domain E3 ubiquitin-ligase domain.

    The BIRs bind directly to an N-terminal tetrapeptide sequence

    exposed during proteolytic activation of caspase-3, caspase-7,

    and caspase-9, thus inhibiting the activity of these proapoptotic

    proteases. IAPs ubiquitin-ligase activity also leads to the ubiq-

    uitylation and degradation of caspases (Dueber et al., 2011)

    and other proteins involved in apoptosis and NF-kB signaling

    (reviewed in Fulda and Vucic, 2012). Current thinking holds

    that XIAP inhibition directly affects caspase activity, whereas

    cIAP inhibition leads to stronger ubiquitin-mediated effects.

    Development of homolog-selective inhibitors of the IAPs will

    allow evaluation of these pathways in cancer therapy and toxicity

    (Condon et al., 2014; Ndubaku et al., 2009; Sun et al., 2014).

    An endogenous inhibitor of IAPs, called Smac (second mito-

    chondrial activator of caspases), competeswith caspase binding

    to BIR domains and thereby stimulates apoptosis (Liu et al.,

    2000). Small-molecule IAP inhibitors mimic the tetrapeptide-

    binding motif on Smac and are therefore called Smac mimetics

    (SMs). The Smac peptide motif, Ala-Val-Pro-Ile (AVPI), binds to

    two adjacent subpockets on BIR domains (Liu et al., 2000).

    Hot-spot residues on the XIAP BIR3 domain focus on the pocket

    that binds the N-terminal amine (E314) and alanine residue (L307

    and W310) and W323, which interacts with proline. Smac pep-

    tides are impressively tight binders; AVPI-amide binds cIAP1

    with a Kd value of 2 nM (Condon et al., 2014). Since the first pep-tidomimetic inhibitor of the IAP/Smac (and IAP/caspase) interac-

    tion was disclosed (Oost et al., 2004), several groups have devel-

    oped inhibitors with improved potency, PK, and in vivo activity.

    These compounds mimic the key interactions between AVPI

    and BIR domains (Figure 2). Seven SMs have reached clinical tri-

    als, and five compounds remain active in clinical development.

    1, September 18, 2014 2014 Elsevier Ltd All rights reserved 1103

  • Clinical IAP inhibitors fall into two structural camps. The

    monovalent antagonists, LCL161 (Novartis) (Dubrez et al.,

    2013), GDC-0917/CUDC-427 (Genentech/Curis) (Flygare et al.,

    2012; Wong et al., 2013) and SM-406/AT-406 (Wang lab/

    Ascenta) (Cai et al., 2011) were designed from the AVPI peptide

    and have IC50 values in the 260 nM range and a LE of0.3 (LLEsare in the drug-like range of 67.5). The bivalent antagonists,

    birinapant (TetraLogic) (Benetatos et al., 2014; Condon et al.,

    2014) and SM-1387 (Bai et al., 2014), are dimerized SMs, de-

    signed based on the idea that the Smac homodimer simulta-

    neously binds to XIAPs BIR2 and BIR3 domains (Huang et al.,

    2003). These dimers demonstrate better cellular potencies

    than their monovalent analogs. Unlike the orally active, monova-

    lent compounds, the bivalent drug candidates are dosed intrave-

    nously. SM-406 and birinapant bind 50-fold more tightly tocIAP1 than XIAP and may allow clinical evaluation of whether

    XIAP inhibition is required for efficacy.

    Identifying patients who will respond to IAP inhibitors as single

    agents has been challenging. SMs induce apoptosis in 15% ofcancer cell lines (Oost et al., 2004), suggesting that single-agent

    activity of clinical compounds could be unpredictable. Transcrip-

    tion profiling and genetic data suggest that certain tumor types

    might be more sensitive, particularly MALT lymphomas with

    constitutively active IAP protein fusions (Fulda and Vucic, 2012).

    Importantly, preclinical and clinical studies have identified serum

    cytokines and cIAP levels as potential biomarkers of response

    (Dubrez et al., 2013), which could be used to quickly assess

    whether a patient is resistant to SM treatment. Furthermore, pre-

    clinical data provides a strong rationale for several combinations

    with radiation and chemotherapeutics, including the apoptotic

    agonist TRAIL (Fulda and Vucic, 2012). The preclinical data also

    scription, and epigene

    histones. Transcriptio

    contain multiple enzym

    ing proteins (Lessard

    represent multiple opp

    tion using small-molec

    Bromodomains are

    ylated lysines (Kacs)

    complexes to turn o

    2014). Bromodomains

    formed by a left-hand

    of various lengths and

    ordered water molecu

    100 mM affinity (Vollbeen exploited by syn

    higher affinity than the

    examples, (+)-JQ1, sti

    containing protein 4 (B

    ability of bromodomain

    Vidler et al., 2012). (+

    49 nM. Crystal struct

    cupies the acetyl-lysin

    interactions not seen

    (Figure 3). In the pas

    been described and h

    bromodomains in canc

    kopoulos and Knapp,

    Five compounds h

    I-BET762 (Glaxo Smit

    (Constellation) (Gehlin

    2012), and OTX15 (On

    1104 Chemistry & Biology 21, September 18, 2014 2014 Elsevier Ltd All rights reservedrequires the complex integration of signal

    transduction, the action of DNA-binding

    proteins that repress or stimulate tran-

    tic signals on the DNA and DNA-bound

    nal complexes are highly dynamic and

    es, scaffolding proteins, and DNA-bind-

    and Crabtree, 2010). These complexes

    ortunities to activate or repress transcrip-

    ule modulators of PPIs.

    epigenetic readers that recognize acet-

    on histone tails and direct transcription

    n genes (Filippakopoulos and Knapp,

    share a conserved hydrophobic pocket

    ed four-helical bundle and loop regions

    charges. An asparagine residue and fiveFigure 2. Peptide-like Inhibitors of PPIswith Primary Epitopes(A) SAR1118 (lifitegrast) inhibits binding of LFA1(CD11a/CD18) to its ligand ICAM. The mechanismeither involves binding directly to the ICAM site inthe I domain of CD11a (Keating et al., 2006) orallosteric inhibition through binding to the I-likedomain in CD18 (Shimaoka et al., 2003).(B) X-ray structure of a cIAP1/XIAP chimera (whitesurface) bound to GDC-0152 (green sticks; PDB:3UW4) and overlaid with SMAC peptide (magentacartoon; PDB: 1G73). Chemical structures ofSMAC mimetics in clinical trials are shown.

    point to the potential for on-target toxic-

    ities (Erickson et al., 2013; Yang and

    Novack, 2013). To date, clinical trials are

    too early to assess anticancer activity,

    but the clinical compounds appear to

    behave well and are tolerated at doses

    that show strong pharmacodynamic ef-

    fects (Infante et al., 2014; A.W. Tolcher

    et al., 2013, J. Clin. Oncol., abstract).

    Bromodomains

    The cellular decision to transcribe a gene

    Chemistry & Biology

    Reviewles recognize Kac-histone peptides with

    muth et al., 2009). This deep pocket has

    thetic Kac mimetics that bind with much

    native peptide. One of the first published

    mulated strong interest in bromodomain-

    RD4) as a cancer target and in the drugg-

    s as a class (Filippakopoulos et al., 2010;

    )-JQ1 binds to BRD4 with an affinity of

    ures show that (+)-JQ1 completely oc-

    e binding groove, including hydrophobic

    in the histone H3 peptide/BRD4 complex

    t 5 years, several potent inhibitors have

    ave been used to explore the biology of

    er and inflammation (reviewed in Filippa-

    2014).

    ave already advanced to clinical trials.

    h-Klein) (Mirguet et al., 2013), CPI-0610

    g et al., 2013), Ten-010 (Tensha) (Rohn,

    cEthix) (P. Herait et al., 2014, Am. Assoc.

  • Chemistry & Biology

    ReviewCancer Res., conference) are based on the JQ1 scaffold

    (Figure 3) and are in clinical trials for cancer, including NUT

    midline carcinoma, a rare cancer caused by BRD4-NUT fusions

    (Filippakopoulos et al., 2010). RVX-208 is a quinazolinone that

    binds deeply into the Kac pocket, making a hydrogen bond

    to the Kac-recognizing asparagine residue; unlike the JQ1-like

    compounds, it binds preferentially to the second bromodomain

    of BRD3 (McLure et al., 2013). RVX-208 has been in phase 2

    trials for atherosclerosis, with promising results. The clinical

    compounds were identified through a range of methods. For

    instance, I-BET762 and RVX-208 were found in cell-based HTS

    focused on ApoA1 upregulation (for atherosclerosis) and were

    later found to bind BET bromodomains (McLure et al., 2013; Mir-

    guet et al., 2013); compound 3, a published analog of CPI-0610,

    was evolved from an isoxazole fragment merged with the core

    structure of JQ1 (Gehling et al., 2013). BET proteins may repre-

    sent a best-case scenario for PPI inhibition, since the primary

    recognition element is a single amino acid. However, the ability

    to develop potent, ligand-efficient (average LE = 0.34; Table 1)

    tecting IN from prot

    catalytic activity (Enge

    of LEDGF binding dom

    shows a short turn o

    dimer interface, with

    (Cherepanov et al., 20

    peptides provided a p

    IN PPI, and stimulat

    based on virtual scree

    viewed inDe Luca et al

    tert-Butoxy-(4-phenyl-

    tives (Figure 3), inclu

    (Boehringer-Ingelheim

    viral potency, LE = 0.3

    2014; Fenwick et al., 2

    signed to identify inh

    DNA; structural studie

    LEDGF binding pocke

    core positions the tw

    Chemistry & Biology 21, September 18, 2014 mimics the binding of a peptide derived fromHIF1a (PDB: 4AJY; not shown).

    inhibitors, even though histone peptides

    bind with low affinity, speaks to the po-

    tential to gain binding energy that is not

    used by the native PPI.

    HIV Integrase

    Viruses hijack host proteins to facilitate

    their replication and survival; these host/

    pathogen complexes represent impor-

    tant clinical targets. The homotetrameric

    protein HIV integrase (IN) integrates the

    viral genome into human DNA by cata-

    lyzing 30 processing and strand transfer.The HIV drugs raltegravir and elvitegravir

    bind to the active site of IN and inhibit its

    enzymatic activity. Recently, a second

    class of IN inhibitors targeting a host/

    pathogen PPI have reached the clinic

    (Engelman et al., 2013).Figure 3. Small-Molecule Inhibitors of PPIswith Primary Epitopes(A) Crystal structure of BRD4 bromodomain (whitesurface) bound to (+)-JQ1 (green sticks; PDB:3MXF) overlaid with acetylated histone peptide(magenta cartoon; PDB: 2WP1). Clinical com-pounds or their closest published analogs areshown below.(B) Crystal structure of the dimerization interface ofHIV integrase (white and cyan surface) bound tocompound 16 (green sticks; PDB: 4NYF) withoverlaid epitope from LEDGF (magenta; PDB:2B4J). Compound 16 is a precursor to the clinicalcompound BI 224436.(C) Von-Hippel Lindau (VHL) protein (white sur-face) with bound inhibitor compound 51 (greensticks; PDB: 4B9K). The central hydroxyprolineThe human protein lens endothelial

    growth factor (LEDGF) acts as an IN

    cofactor, facilitating integration of the viral

    genome into the host chromosome, pro-

    eolytic degradation, and stimulating its

    lman et al., 2013). The cocrystal structure

    ain with a dimer of IN catalytic domain

    f helix from LEDGF pointing into the IN

    400 A2 buried surface area on IN

    05; Basse et al., 2013). LEDGF-mimicking

    roof of concept for inhibiting the LEDGF/

    ed small-molecule discovery programs

    ning, structure-aided design, and HTS (re-

    ., 2011). Themost potent PPI inhibitors are

    quinolin-3-yl)-acetic acid (tBPQA) deriva-

    ding the clinical compound BI-224436

    , Gilead), which has low nanomolar anti-

    3, and LLE = 11.5 (Table 1) (Fader et al.,

    014). tBPQAs were found in an HTS de-

    ibitors of IN-catalyzed 30 processing ofs later revealed the compound bound to

    t at the IN dimer interface. The quinolone

    o side-chain mimetics; as with BRD4

    2014 Elsevier Ltd All rights reserved 1105

  • to

    )Table 1. Properties of Selected Protein-Protein Interaction Inhibi

    PPI Target/Partnera Compound Name MW (Da

    Primary Epitope

    LFA1/ICAM1 SAR1118 (lifitegrast) 615

    cIAP/SMAC GDC-0152 499

    GDC-0917/CUDC-427 565

    AT-406/SM-406 562

    TL-32711 (birinapant) 809

    Bromodomain/histone I-BET762 (GSK 525762) 424

    compound 3e 400

    RVX-208 370

    Integrase/LEDGF BI 224436 443inhibitors, BI-224436 binds to pockets that the LEDGF turn does

    not (Figure 3).

    tBPQAs are dual inhibitors. In addition to inhibiting the LEDGF/

    IN complex, they also allosterically inhibit IN catalytic activity

    by preventing the formation of functional IN tetramer. Recent

    studies suggest that the mechanism of these inhibitors is com-

    plex and altering tetramerization may be more therapeutically

    important than inhibiting LEDGF per se (Balakrishnan et al.,

    2013; Sharma et al., 2014). BI-224436 could provide a new

    mechanism for HIV treatment that will most likely have a different

    resistance profile than active site IN inhibitors. Furthermore, its

    story highlights the fact that enzyme/protein complexes have

    multiple effects on the enzyme; small-molecule inhibitors might

    alter one or more of these functions.

    Prospects for Primary Sequence Epitopes

    Inhibitors of PPIs that consist of primary sequence epitopes have

    been discovered using multiple approaches. Fragment-based,

    biochemical, and cell-based screening have all yielded hits for

    this PPI class (Fader et al., 2014; Gehling et al., 2013; Mirguet

    VHL/HIF1a compound 51 450

    Secondary Epitope

    BCL family/BH3 ABT-263 (navitoclax) 975

    ABT-199 868

    compound 14 437

    MDM2/p53 RG7112 (Ro5045337) 728

    RG7388 (Ro5503781) 616

    MI-888f 548

    PDK1/PIF-tide PS210 380

    Menin/MLL MIV-6R 418

    MIV-2-2 416

    p300 CH1 domain/HIF1a OHM 1 495

    Tertiary Epitope

    IL-2/IL-2Ra SP4206 663

    HPV11 E1/E2 BILH 434 608

    MW, molecular weight; nd, not determined.aItalicized compounds are in active clinical development. For PPI families (e.

    IC50 is reported. When available, Kd values are reported.bLE = DG / number of heavy atoms.cLLE = pIC50 clogD.dActivity is

  • Chemistry & Biology

    ReviewSecondary Structure Epitopes: a Helix, b Sheet, orExtended PeptidesIt has been estimated that up to 40% of PPIs consist of a single

    peptide from one protein that binds into a groove on the other

    (Petsalaki and Russell, 2008). Sixty percent of PPI in the protein

    databank utilize an a helix, and 60%of these bind a single face of

    the helix (Raj et al., 2013). In general, hot spots on the target pro-

    tein are centered on two to three subpockets that bind hydro-

    phobic residues on the partner peptide. Such interfaces have

    been described as hot sequences (London et al., 2010) or

    secondary-structural epitopes, reflecting the fact that the

    key residues on the peptide are not next to each other in the pri-

    mary sequence, but that an isolated peptide can often mimic the

    binding affinity and pose of the PPI.

    BCL Family

    BCL family proteins are central effectors of apoptosis. Proapo-

    ptotic family members, including BAD and BAK, are kept in an

    inhibited state through binding to the antiapoptotic family mem-

    bers, including BCL-2, BCL-xL, and MCL1. The antiapoptotic

    members are helical proteins with an open groove that binds

    to a single helix (the BH3 domain) on the proapoptotic partner.

    Upregulation of the antiapoptotic BCL family members is an

    important mechanism through which cancer cells avoid cell

    death in the face of pathway dysregulation, radiation, and

    chemotherapy. The structure of BCL-xL bound to BAK peptide

    (16-mer, Kd = 340 nM) suggested that this PPI might be particu-

    larly druggable (Sattler et al., 1997). The binding site onBCL-xL is

    comprised of two large (200 A2) nearby pockets and three hy-drophobic residues from the BH3 peptides define the PPI hot

    spot (Oltersdorf et al., 2005). Indeed, between 2000 and 2004,

    several laboratories published small-molecule inhibitors with

    (Oltersdorf et al., 200

    available analog ABT-2

    1 and 2 clinical trials

    leukemia (CLL) (Rober

    these early clinical tria

    an important dose-lim

    Current clinical BCL

    BCL-xL. ABT-199 was

    avoid thrombocytopen

    tivity into ABT-263, a s

    with BCL-2. A trypto

    molecule sat very nea

    that position led to AB

    tightly than ABT-263 a

    compared to BCL-xL

    the BAK-binding pock

    bone structure or side

    currently in phase 1 tr

    very promising, with re

    2014). BCL-xL-selectiv

    have also been develo

    while avoiding the lym

    2014). The L-shaped B

    ingmode fromABT-19

    the peptide-binding gr

    and compound 14 are

    than ABT-737 and AB

    activity for BCL family

    weaker thanbindingaf

    teinbinding,or thecellu

    for several members o

    Chemistry & Biology 21, September 18, 2014 0.21. ABT-737 demonstrated impressive

    single-agent activity in cancer xenograft

    models, but was not orally bioavailable

    5). In 2008, Abbott disclosed the orally

    63 (Tse et al., 2008), which entered phase

    for solid tumors and chronic lymphocytic

    ts et al., 2012; Rudin et al., 2012). During

    ls, thrombocytopenia (platelet loss) was

    iting toxicity linked to BCL-xL inhibition.

    family inhibitors are selective for BCL-2 or

    designed to be selective toward BCL-2 to

    ia (Souers et al., 2013). To engineer selec-

    tripped-down scaffold was cocrystalized

    phan residue from a neighboring BCL-2

    rby to the inhibitor; linking of an indole atFigure 4. BCL-2/BCL-xL and MDM2Interfaces Contain Secondary Epitopes(A) Structure of BCL-2 (white surface) bound toABT-199 (green sticks; PDB: 4MAN) with overlaidBAX BH3 peptide (magenta cartoon; PBD: 2XA0).Chemical structures of selected BCL-2 and BCL-xL inhibitors are shown.(B) Structure of MDM2 (white surface) boundto RG7112 (green sticks; PDB: 4IPF) with over-laid p53 peptide (magenta cartoon; PDB: 1YCR).Chemical structures of compounds in clinicaltesting or their closest published analogs areshown.

    potencies in the 0.110 mM range that

    bound in the BAK-binding site (reviewed

    in Arkin and Wells, 2004).

    The major preclinical breakthrough

    was the 2005 disclosure of ABT-737,

    which was discovered through NMR-

    based fragment discovery, structure-

    based design, and medicinal chemistry

    (Oltersdorf et al., 2005). ABT-737 binds

    to BCL-2, BCL-xL, and BCL-W with sub-

    nanomolar affinity, but it is also a large

    compound (MW = 813 Da), with LE =T-199, which binds BCL-2 10-fold more

    nd shows 5,000-fold selectivity for BCL-2

    . Like its precursors, ABT-199 accesses

    ets, but does not mimic the peptide back-

    -chain orientation (Figure 4). ABT-199 is

    ials for CLL, and preliminary results look

    sponse rates of >80% (Cancer Discovery,

    e inhibitors (e.g., compound 14; Figure 4)

    ped, with the aim of treating solid tumors

    phocytic effects of BCL-2 (Koehler et al.,

    CL-xL-selective series has a distinct bind-

    9 and appears to be buriedmore deeply in

    oove. It is noteworthy that both ABT-199

    more ligand efficient and have better LLE

    T-263 (Table 1). Interestingly, cell-based

    inhibitors is generally 100- to 1,000-fold

    finity. Thismaybedue topermeability, pro-

    lar context ofBCL-2-xLandseems tohold

    f this target class. Thus, very-high-affinity

    2014 Elsevier Ltd All rights reserved 1107

  • compounds have been required (Souers et al., 2013). While it is

    still early days in the clinical life of BCL family antagonists, the

    impressive story of their discovery highlights the realities of drug

    discovery for a novel and challenging target; often, several itera-

    tions of clinical candidates are needed to strike the ideal balance

    of efficacy, safety, and dosing regimen.

    MDM2

    The transcription factor p53 is a master tumor suppressor that

    regulates cell-fate decisions such as senescence, cell-cycle

    arrest, and apoptosis. p53 or its regulators are mutated or

    altered in most cancers, resulting in reduction or loss of p53

    function. The ubiquitin E3 ligase responsible for p53 degradation

    is MDM2 (murine double minute 2) and the MDM2/MDMx heter-

    odimer (Khoo et al., 2014). This system is overexpressed or

    amplified in several cancers with wild-type p53, and inhibition

    of p53 ubiquitylation by MDM2 should lead to increased p53

    levels and activity. The MDM2/p53 complex can be minimized

    to an a helix from the N-terminal transactivation domain of p53

    that binds into a pocket on the surface of the N-terminal domain

    of MDM2 (Kussie et al., 1996). The MDM2/p53 interface is more

    compact than that found for BCL-xL/BH3 peptide: the helix is

    only two turns, and the three critical hot-spot residues (Phe19,

    Trp23, and Leu26) point toward a deep pocket in the center of

    the peptide-binding groove. The isolated peptide binds with

    500 nM affinity to the N domain of MDM2.The breakthrough for inhibiting p53/MDM2 came from the dis-

    covery of the Nutlins, identified in a large high-throughput screen

    (Vassilev et al., 2004). The optimized,100 nM inhibitor containsa central imidazole ring with four substitutions. Three hydropho-

    bic substitutionsmimic the binding of the three key residues from

    p53, and the fourth substituent binds on the protein surface.

    Soon after the discovery of the Nutlins, several labs described

    compounds with different cores but similar molecular-recogni-

    tion features (Fry, 2012; Khoo et al., 2014). A general feature is

    a tripod shape that pushes at least two hydrophobic elements

    deep into the binding site; these are fairly rigid structures that

    use a central ring to enforce the side-chain geometries. Some

    compounds also bind further along the p53-binding groove.

    Compounds also differ in their ability to bindMDMX, whichmight

    be important for efficacy and overcoming resistance in some

    contexts (Khoo et al., 2014). The key observation from these

    diverse MDM2 ligands is that many different scaffolds (including

    peptides) can direct the hydrophobic groups into the p53-bind-

    ing pocket. (Fry, 2012). Thus, it is not critical to mimic precisely

    the peptide backbone or the peptides side-chain orientation.

    One of the puzzling aspects ofMDM2 inhibitor design has been

    the role of the fourth substituent. Modifications to this group alter

    the compounds physicochemical characteristics and have a

    modest effect on binding to the N-terminal domain of MDM2.

    However, changes often have a disproportionate effect on cell-

    based potency (Ding et al., 2013; Zhao et al., 2013). Recent

    data suggest that parts ofMDM2outside thep53-bindingdomain

    may play a role in the binding interaction. For instance, MDM2mutations that render cells resistant to Nutlins often occur

    outside the p53-binding domain (Wei et al., 2013). The direct

    demonstration of a structural interaction has yet to be made,

    but thegeneral point rings true: inminimizing thePPI for biochem-

    ical and structural studies, we are likely to miss long-range inter-

    actions that the compounds will face when they enter the cell.

    1108 Chemistry & Biology 21, September 18, 2014 2014 Elsevier LtThree scaffolds have been developed into experimental thera-

    peutics. Roche initially took RG7112 (Figure 4) into phase 1 trials

    (Tovar et al., 2013) and is currently testing the more potent and

    ligand-efficient analog RG7388 (Ding et al., 2013). The spirooxin-

    dole class of compounds, exemplified by MI-888, was discov-

    eredby theWang lab atUniversity ofMichigan and is beingdevel-

    oped by Sanofi (MI-773, structure not disclosed) (Zhao et al.,

    2013). Roche recently disclosed Ro8994, a 5 nM (95% bioavail-

    able) analog based on a spiroindolinone scaffold that combines

    features of RG7388 andMI-888 (Zhang et al., 2014). Finally, Daii-

    chi Sankyo is developing an undisclosed MDM2 inhibitor based

    on a dihydroimidazothiazole scaffold (Miyazaki et al., 2013).

    These clinical-stage inhibitors show impressive tumor regres-

    sions in xenograft models and selective killing of cells containing

    wild-type p53 versus those with mutant p53 (Tovar et al., 2013).

    Their drug-like properties are also promising, with LE in the range

    of 0.270.35, LLE in the range of 4.36.7, and very high oral bio-

    availabilities despite molecular weights in the range of 550730.

    MDM2 inhibitors are being developed for single-agent use and

    in combination with existing drugs. Early trials for RG7112 have

    shown pharmacodynamic activity, including 3- to 5-fold in-

    creases in levels of p53 and p53 target genes (Ray-Coquard

    et al., 2012), and potential therapeutic activity in acute leukemia

    (M. Andreeff et al., 2012, Am. Soc. Hematol., conference). The

    biological challenges facing clinical positioning of these com-

    pounds has been recently discussed (Khoo et al., 2014). Despite

    decades of work on p53, its precise roles in survival and death

    pathways remain to be elucidated; these clinical MDM2 antago-

    nists are poised to address some of these critical issues.

    Selected Emerging Examples of Secondary Structure

    Epitope Inhibitors

    Diverse methods have been used to develop secondary struc-

    tural mimetics and prospects for developing inhibitors of this

    class of PPI are good. One instructive example is the N lobe of

    the kinase PDK1. This site binds to a 10-residue peptide called

    a PDK1-interacting fragment (PIF-tide) from substrate proteins.

    Binding serves to colocalize the proteins and to allosterically acti-

    vate PDK1 kinase activity (Gold et al., 2006). Fragments that

    mimic thePIF-tide havebeen found throughvirtual screening (En-

    gel et al., 2006), NMR (Stockman et al., 2009), and tethering

    (Sadowsky et al., 2011). For tethering, six cysteine residues

    were introduced around the PIF-tide binding site and were

    screened for binding to a library of disulfide-containing frag-

    ments. Remarkably, fragments tethered at the same cysteine

    residues could be either activators or inhibitors of PDK-1 kinase

    activity. Cocrystal structures showed that activators pushed

    the C helix toward the active site and better organized the cata-

    lytic apparatus, whereas inhibitors shifted the C helix outward

    into a less competent state (Figure 5). Thus, small changes in

    the small-molecule ligand can have positive or negative effects

    onenzymeallostery. Virtual screening, followedbychemical opti-

    mization, provided cell active inhibitors of the PDK1/PIF-tide

    interaction. PS210 (LE=0.35)mimics two hot-spot phenylalanine

    Chemistry & Biology

    Reviewresidues on the PIF-tide (Busschots et al., 2012). In biochemical

    assays, this compound activates the kinase toward peptide sub-

    strates, asdoes thePIF-tide. Paradoxically, in cells, a cell-perme-

    able version of PS210 inhibits phosphorylation of the subset of

    PDK1 substrates that require binding at the PIF-tide site. We

    recently identified cell active compounds from an HTS that

    d All rights reserved

  • Chemistry & Biology

    Reviewmonitored displacement of an optimized PIF-tide (unpublished

    data). These compounds serveas additional examples of peptide

    mimicry by small molecules and provide ligand efficient scaffolds

    that may be further optimized.

    Multiple approaches have also led to inhibitors of the transcrip-

    tion factor mixed-lineage leukemia (MLL) protein at two different

    domains. MLL binds to the scaffolding protein menin, and this

    interfacecanbeminimized toan11-merpeptide from theN-termi-

    nal domain ofMLL.HTS followedby structural biology andmedic-

    inal chemistry led to MIV2-2 and MIV-6, potent, ligand-efficient,

    and cell-active inhibitors that closely mimic the MLL peptide hot

    spots (Figure 5 and Table 1) (He et al., 2014; Shi et al., 2012).

    MLLalsobinds to theKIXdomainof theacetyltransferaseCBPus-

    ing an a helix in the C-terminal domain of MLL. Mapp and

    colleagues have designed synthetic molecules that bind to the

    KIX domain at the MLL site (Buhrlage et al., 2009). They have

    also used tethering to identify disulfide-bound compounds that

    bind in the MLL site and allowed the first X-ray structure of the

    KIX domain (Wang et al., 2013b). Their ultimate goal is to develop

    bifunctional molecules that activate transcription by bringing

    together DNA-binding proteins and transcriptional coactivators.

    Prospects for Secondary Structure Epitope Design

    Design approaches have focused on both the binding pocket

    and the partner peptide. When focusing on the binding pocket,

    (Figure 5 and Table 1),

    and inhibits tumor grow

    Computational appro

    residues have also se

    2014). This type of epi

    peptide approaches (

    Boersma et al., 2012;

    of PPIs with secondar

    low LE; however, the

    and the PDK1, menin

    more favorable ranges

    Tertiary Structural EBinding SitesInterleukin-17

    The larger, shallower i

    discovery remain chall

    disclosed the develop

    hibits binding of interle

    et al., 2012), which

    hormone/receptor int

    undisclosed structure

    a LE of 0.24. We eand preclinical data su

    Chemistry & Biology 21, September 18, 2014 interfaces often show small motions that

    create deeper binding pockets (Johnson

    and Karanicolas, 2013; Wells and

    McClendon, 2007). Pocket-based design

    is justified by the observation that inhibi-

    tors such as ABT-199 and RG7112 mimic

    the function of hot-spot residues but

    rarely mimic trajectories seen in the pep-

    tide structures. Others have focused on

    strategies to mimic the peptide partner

    directly. For example, Lao et al. designed

    an a-helical mimetic of the HIF1a transac-

    tivation domain that binds to the cysteine-

    histidine rich 1 (CH1) domain on CBP/

    p300. The optimized compound OHM 1

    is based on a bis-oxopiperazine scaffold;

    OHM 1 binds CH1 with a Kd of 500 nM

    downregulates hypoxia-inducible genes,

    th in a xenograft model (Lao et al., 2014).

    aches to systematically replace peptideFigure 5. Preclinical Examples of Inhibitorsof PPIs with Secondary Epitopes(A) Crystal structures of PDK-1, showing the PIF-tide binding site bound with Tethered activator(JS30, green sticks; PDB: 3OTU) or inhibitor (1F8,magenta sticks; PDB: 3ORX). JS30 pushes the Chelix into a catalytically competent conformation.The chemical structure of noncovalent PIF-tideinhibitor PS210 is shown.(B) Crystal structure of menin (white surface)shown bound to MIV-6R (green sticks; PDB:4OG8) with MLL peptide overlaid (magentacartoon; PDB: 3U85). Structures of MLL inhibitorsMIV-6R and MIV 2-2 are shown.(C) Structure of OHM 1, which binds to CBP at theHIF1a peptide-binding site.

    virtual methods should allow for side-

    chain flexibility, since hot spots at theseen some promising success (Guo et al.,

    tope is also highly amenable to stabilized

    Azzarito et al., 2013; Bernal et al., 2010;

    Chang et al., 2013). Historically, inhibitors

    y-structural epitopes were large and had

    second-generation clinical compounds

    , and HIF1a lead compounds all fall into

    for LE and LLE.

    pitopes: Discontinuous

    nterfaces that originally discouraged drug

    enging. Recently, Ensemble Therapeutics

    ment of a macrocyclic compound that in-

    ukin-17 (IL-17) to its receptor (Livingston

    represents a breakthrough in targeting

    erfaces with synthetic macrocyles. The

    has 3 nM affinity, a MW of 750 Da, andagerly await publication of the structural

    rrounding this program.

    2014 Elsevier Ltd All rights reserved 1109

  • IL-2 and HPV-11

    Two examples of PPI inhibitors highlight the dynamic nature

    of globular PPIs. A series of compounds, exemplified by

    SP4206, binds to IL-2 at the receptor-alpha interface (Figure 6)

    (reviewed in Wilson and Arkin, 2011). These compounds were

    identified serendipitously and optimized through fragment-

    based methods, including tethering. IL-2 presents a relatively

    flat surface in the apo- and IL-2Ra bound states but shows

    pockets upon binding SP4206. SP4206 binds at the receptor-binding hot spot and mimics the receptors hot-spot residues

    and charge distribution, even though the residues on are sepa-

    rated over multiple regions of the receptors structure. The

    conformational changes seen across the SP4206-binding sur-

    face include side-chain rotations and loop rearrangements,

    larger changes than those observed for the peptide epitopes

    described above. Molecular dynamics simulations suggest that

    these structural adaptations are better thought of as part of the

    protein conformational ensemble rather than induced fits.

    BILH 434 (Figure 6) is another example of an inhibitor of a ter-

    tiary structural epitope (Wang et al., 2004). Human papilloma

    virus-11 (HPV11) requires the replication initiation factor E1 heli-

    case to bind to the E2 transcription factor at specific DNA sites.

    BILH 434 was identified in a screen for inhibitors of the E1/E2/

    DNA complex and binds to the transactivation domain (TAD) of

    E2 protein with a Kd of 40 nM (LE = 0.25). The X-ray structure

    of the compound/E2 complex shows the compound nestled

    into the elbow of the E2 TAD. Several side chains in this protein

    interface change conformations to create a deep pocket for BILH

    434 to bind. Again, changes to the structure were unanticipated

    and were crucial for creating a binding site at an otherwise flat

    surface. Importantly, both the IL-2 and E2 TAD discoveries relied

    heavily on biophysical validation of the binding mode that was

    confirmed by X-ray crystallography.

    Prospects for Tertiary Epitope Binders

    This class is probably the most challenging, and there are only

    a handful of successful programs that have broken into the nano-

    molar level of potency with modest LEs (typically 0.24 kcal/HA).

    Allosteric MechanisFunctional screening

    tously led to allosteric

    BIO8898, an inhibitor

    conformation of the trim

    vianet al., 2011). Simila

    and disrupt trimer form

    component complexe

    gray box screening,

    in vitro. This approach

    HSP70 chaperone netw

    and how these HSP7

    effects on folding and

    et al., 2013; Wang et a

    dipitous) allosteric mo

    forward for the most d

    How Has Our Under10 Years?The discovery of small

    tinues to be challengin

    strate that it is possible

    being sought, but it is l

    a target class analogo

    extremely diverse in t

    the binding affinities an

    are (conservatively) 10

    1110 Chemistry & Biology 21, September 18, 2014 2014 Elsevier Ltd All rights reservedms of Inhibitionfor PPI inhibitors has also serendipi-

    mechanisms of inhibition. For instance,

    of the trimeric ligand CD40L, alters the

    er, thus inhibiting its binding toCD40 (Sil-

    rly, compoundshavebeen found that bind

    ation for TNF (He et al., 2005). For multi-

    s, one way to harness this serendipity is

    where the protein network is assembled

    has yielded interesting modulators of the

    ork. Depending on the biological context

    0 modulators bind, they have opposite

    thus utility in different disease states (LiFigure 6. Small-Molecule Inhibitors of PPIswith Tertiary Epitopes(A) Structure of IL-2 (white surface) bound toSP4206 (green sticks; PDB: 1PY2).(B) Structure of transactivation domain of E2 pro-tein fromHPV11 (white surface) bound to BILH 434(green sticks; PDB: 1R6N).

    These interfaces tend to be much more

    dynamic than the primary and secondary

    class epitopes and thus less amenable

    to virtual screening. Small molecules

    can, however, find deeper pockets in

    these interfaces and thus bind with higher

    LE than their protein partner counterpart.

    The small molecules also bind hot spots

    predicted from alanine scanning. This

    class is probably themost prevalent class

    for extracellular PPIs. Strong validation

    of the biology will certainly make this

    class temptingif high-hangingfruit.

    For example, four groups have recently reported allosteric and

    orthosteric PPI inhibitors of Ras, a very challenging cancer target

    with no apparent deep pockets in the interface. Interestingly,

    three groups took fragment-based approaches, using NMR

    (Maurer et al., 2012; Sun et al., 2012) and tethering (Ostrem

    et al., 2013), and one group identified inhibitors through virtual

    screening (Shima et al., 2013). Finally, PPI-binding peptides

    discovered from phage display could provide another avenue

    toward peptide-based therapies (DeLano et al., 2000).

    Chemistry & Biology

    Reviewl., 2013a). Directed (as opposed to seren-

    dulation of PPIs provides potential way

    ifficult targets (Nussinov and Tsai, 2014).

    standing of PPIs Evolved in the Past

    , drug-like molecules that inhibit PPIs con-

    g, but the diverse success stories demon-

    for some systems. General strategies are

    ikely that PPI is not properly thought of as

    us to class I GPCRs or kinases. PPIs are

    he size and shapes of the interfaces and

    d dynamics of assembly. Given that there

    0,000 PPIs (Venkatesan et al., 2009), our

  • current database of PPI inhibitors is small and biased. Neverthe-

    less, here is what we know so far.

    Analyses of PPIs in which crystal structures are available for

    both the protein-protein and protein-small-molecule reveal

    someprovocative themes (JohnsonandKaranicolas, 2013;Smith

    and Gestwicki, 2012; Wells and McClendon, 2007). First, PPI

    target proteins tend to have extended binding grooves that may

    be divided into three or more (visible) subpockets (Fuller et al.,

    2009). Hot-spot residues are centered in or around these pockets

    and are complementary on both sides of the interface. Second,

    small conformational changes in thebinding site present a deeper

    pocket to small molecules than to the partner protein. These

    pockets are also more readily formed at the PPI interfaces than

    elsewhere on the protein surfaces (Johnson and Karanicolas,

    2013). The consensus is that ligand-binding surfaces (including

    PPIs) haveahigherpropensity forbinding (tocognate ligands, sol-

    vent molecules, inhibitors) than the rest of the protein surface.

    Third, PPIs with small-molecule inhibitors tend to have small,

    high-affinity interfaces and include a hot segment that can reca-

    pitulate the binding of the partner protein. The observation that

    high-affinity complexes havemore small-molecule ligands (Smith

    andGestwicki, 2012) could imply that they are inherently sticky.

    Fortunately, however, there are many examples where peptides,

    mutated proteins, and small moleculessuch as bromodomain

    inhibitorsbind to hot spot sites with higher affinities than the

    native PPI. The preponderance of inhibitors for primary- and sec-

    ondary-structural epitopes is probably due to their inherent

    druggability, the fact that there are many such interfaces (Petsa-

    laki and Russell, 2008), and the availability of design approaches.

    In general, success seems to follow the simplicity of the epitopes

    to mimic: primary is easier than secondary is easier than tertiary.

    PPI inhibitors bind at the hot spot and tend tomimic the types of

    binding interactions (e.g., hydrophobic)madeby thepartner.How-

    ever, they generally use different functional groups and approach

    thebinding site fromdifferent orientations than in thePPI. To reach

    each subpocket in an extended interface, molecules often have

    three distinct pharmacophores arrayed on a scaffold that serves

    to orient them. Hence, orthosteric inhibitors have tended to be

    larger than 500Da and to bemore 3D than the average compound

    inanHTS library (Basseetal., 2013; Labbe etal., 2013). Toaddress

    the potential lack of PPI-inhibitor-like compounds in HTS libraries,

    investigators have proposed design of libraries that mimic the

    shapes of current PPI inhibitors or protein secondary structures.

    To date, more traction has been gained through virtual screening

    and the use of nontraditional libraries, including larger, more natu-

    ral-product-like macrocycles on the one hand and fragment-

    based lead discovery on the other. Fragment screening has

    been a particularly successful strategy, perhaps because the

    pocket sizes at PPIs are close to fragment size and the bias in

    chemical composition is lower. Experimental and virtual fragment

    screens have even be used to define the druggability of a PPI (Haj-

    duk and Greer, 2007; Kozakov et al., 2011).

    On the other hand, second-generation clinical candidates are

    Chemistry & Biology

    Reviewgenerally smaller andmore efficient than their precursors, and so

    perhaps the inherency of undruglikeness of PPI inhibitors has

    been overstated (Table 1), particularly for small interfaces. Con-

    cerns for LE and LLE originate from the observation that larger

    molecules are more difficult to tune for in vivo properties. More

    important than molecular weight per se is finding the balance

    Chemistry & Biology 2of hydrophobicity and electrostatics that provides tight, specific

    binding to the protein target and acceptable pharmacokinetics

    and pharmacodynamics. These issues are by no means specific

    to PPI drug discovery.

    Screening methodologies have also become more sophisti-

    cated in the past 10 years. Foremost, there is an expanded

    use of multiple drug discovery technologies, including an inte-

    grated use of biochemistry, biophysics, structural biology, and

    computational studies. Initial PPI inhibitors have been discov-

    ered using a wide variety of methods; we conclude that the

    screening format itself is less crucial than the compounds going

    into the screen and the postscreening evaluation of hits. True

    positives should be validated by demonstrating that the inhibitor

    binds to a single site on the PPI target using biophysical and

    structural methods. The importance of mechanism of action

    over potency at the screening stage cannot be overstated.

    What We Expect to See More of in the Near FutureTwenty years ago, PPIs were considered impossible fruit to

    reach. Today, there are dozens of targets for which early stage

    compounds have been discovered, and nearly a dozen com-

    pounds have been in the clinic. Clearly, much progress has

    been made, and this has generated the courage to move

    forward, even with enthusiasm.

    Ten years from now, how far will we have come? In terms of

    methodology, new computational methods will allow better un-

    derstanding of allostery and conformational ensembles that are

    grist for improved virtual screening algorithms. There will be

    a continued emphasis on fragment-based design, and also a

    resurgence in natural products and natural-product-like mole-

    cules, since biology has evolved these molecules for binding at

    protein interfaces. Popular target classes will include epigenetic

    and transcription factor complexes, signal transduction net-

    works, and protein quality control systems, reflecting the need

    for chemical tools to dissect these complex protein networks.

    Building on the function of natural products, there will also be

    greater emphasis on synthetic stabilizers of PPIs (Milroy et al.,

    2014). Advances in cell biologywill allow faster discovery of com-

    pounds mechanism of action and more translatable cell-based

    assays. Importantly, the PPI inhibitors in clinical trials today will

    be drugs benefitting patients. To us the writing is on the wall:

    PPIs are a challenging but tractable class for small-molecule dis-

    covery if the biology is compelling and the will is strong.

    ACKNOWLEDGMENTS

    We thank our colleagues in the PPI inhibitor field for their beautiful work andapologize to those whose work we could not describe due to space consider-ations. We also thank Jason Gestwicki for his thoughtful edits. This work wassupported by NIH grants AG046526 and CA136779.

    REFERENCES

    Arkin, M.R., andWells, J.A. (2004). Small-molecule inhibitors of protein-proteininteractions: progressing towards the dream. Nat. Rev. Drug Discov. 3,

    301317.

    Azzarito, V., Long, K., Murphy, N.S., and Wilson, A.J. (2013). Inhibition ofa-helix-mediated protein-protein interactions using designed molecules.Nat. Chem. 5, 161173.

    Bai, L., Smith, D.C., and Wang, S. (2014). Small-molecule SMAC mimetics asnew cancer therapeutics. Pharmacol. Ther. 144, 8295.

    1, September 18, 2014 2014 Elsevier Ltd All rights reserved 1111

  • Balakrishnan, M., Yant, S.R., Tsai, L., OSullivan, C., Bam, R.A., Tsai, A., Nied-ziela-Majka, A., Stray, K.M., Sakowicz, R., and Cihlar, T. (2013). Non-catalyticsite HIV-1 integrase inhibitors disrupt core maturation and induce a reversetranscription block in target cells. PLoS ONE 8, e74163.

    Basse, M.J., Betzi, S., Bourgeas, R., Bouzidi, S., Chetrit, B., Hamon, V., Mor-elli, X., and Roche, P. (2013). 2P2Idb: a structural database dedicated toorthosteric modulation of protein-protein interactions. Nucleic Acids Res. 41(Database issue), D824D827.

    Benetatos, C.A., Mitsuuchi, Y., Burns, J.M., Neiman, E.M., Condon, S.M., Yu,G., Seipel, M.E., Kapoor, G.S., Laporte, M.G., Rippin, S.R., et al. (2014). Biri-napant (TL32711), a bivalent SMACmimetic, targets TRAF2-associated cIAPs,abrogates TNF-induced NF-kB activation, and is active in patient-derivedxenograft models. Mol. Cancer Ther. 13, 867879.

    Bernal, F., Wade, M., Godes, M., Davis, T.N., Whitehead, D.G., Kung, A.L.,Wahl, G.M., and Walensky, L.D. (2010). A stapled p53 helix overcomesHDMX-mediated suppression of p53. Cancer Cell 18, 411422.

    Boersma, M.D., Haase, H.S., Peterson-Kaufman, K.J., Lee, E.F., Clarke, O.B.,Colman, P.M., Smith, B.J., Horne, W.S., Fairlie, W.D., and Gellman, S.H.(2012). Evaluation of diverse a/b-backbone patterns for functional a-helix mim-icry: analogues of the Bim BH3 domain. J. Am. Chem. Soc. 134, 315323.

    Buckley, D.L., Gustafson, J.L., Van Molle, I., Roth, A.G., Tae, H.S., Gareiss,P.C., Jorgensen, W.L., Ciulli, A., and Crews, C.M. (2012a). Small-molecule in-hibitors of the interaction between the E3 ligase VHL and HIF1a. Angew.Chem. Int. Ed. Engl. 51, 1146311467.

    Buckley, D.L., Van Molle, I., Gareiss, P.C., Tae, H.S., Michel, J., Noblin, D.J.,Jorgensen, W.L., Ciulli, A., and Crews, C.M. (2012b). Targeting the von Hip-pel-Lindau E3 ubiquitin ligase using small molecules to disrupt the VHL/HIF-1a interaction. J. Am. Chem. Soc. 134, 44654468.

    Buhrlage, S.J., Bates, C.A., Rowe, S.P., Minter, A.R., Brennan, B.B., Majmu-dar, C.Y., Wemmer, D.E., Al-Hashimi, H., and Mapp, A.K. (2009). Amphipathicsmall molecules mimic the binding mode and function of endogenous tran-scription factors. ACS Chem. Biol. 4, 335344.

    Busschots, K., Lopez-Garcia, L.A., Lammi, C., Stroba, A., Zeuzem, S., Piiper,A., Alzari, P.M., Neimanis, S., Arencibia, J.M., Engel, M., et al. (2012). Sub-strate-selective inhibition of protein kinase PDK1 by small compounds thatbind to the PIF-pocket allosteric docking site. Chem. Biol. 19, 11521163.

    Cai, Q., Sun, H., Peng, Y., Lu, J., Nikolovska-Coleska, Z., McEachern, D., Liu,L., Qiu, S., Yang, C.Y., Miller, R., et al. (2011). A potent and orally active antag-onist (SM-406/AT-406) of multiple inhibitor of apoptosis proteins (IAPs) in clin-ical development for cancer treatment. J. Med. Chem. 54, 27142726.

    Cancer Discovery (2014). BCL-2 inhibitor yields high response in CLL and SLL.Cancer Discov. 4, OF5.

    Chang,Y.S.,Graves,B.,Guerlavais, V., Tovar,C., Packman,K., To,K.H.,Olson,K.A., Kesavan, K., Gangurde, P., Mukherjee, A., et al. (2013). Stapled a-helicalpeptide drug development: a potent dual inhibitor of MDM2 and MDMX forp53-dependent cancer therapy. Proc. Natl. Acad. Sci. USA 110, E3445E3454.

    Cherepanov, P., Ambrosio, A.L., Rahman, S., Ellenberger, T., and Engelman,A. (2005). Structural basis for the recognition between HIV-1 integrase andtranscriptional coactivator p75. Proc. Natl. Acad. Sci. USA 102, 1730817313.

    Clackson, T., andWells, J.A. (1995). A hot spot of binding energy in a hormone-receptor interface. Science 267, 383386.

    Condon, S.M., Mitsuuchi, Y., Deng, Y., LaPorte, M.G., Rippin, S.R., Haimowitz,T., Alexander, M.D., Kumar, P.T., Hendi, M.S., Lee, Y.H., et al. (2014). Birina-pant, a smac-mimetic with improved tolerability for the treatment of solidtumors and hematological malignancies. J. Med. Chem. 57, 36663677.

    De Luca, L., Ferro, S., Morreale, F., De Grazia, S., and Chimirri, A. (2011). In-hibitors of the interactions between HIV-1 IN and the cofactor LEDGF/p75.ChemMedChem 6, 11841191.

    DeLano, W.L., Ultsch, M.H., de Vos, A.M., and Wells, J.A. (2000). Convergentsolutions to binding at a protein-protein interface. Science 287, 12791283.Ding, Q., Zhang, Z., Liu, J.J., Jiang, N., Zhang, J., Ross, T.M., Chu, X.J., Bart-kovitz, D., Podlaski, F., Janson, C., et al. (2013). Discovery of RG7388, a potentand selective p53-MDM2 inhibitor in clinical development. J. Med. Chem. 56,59795983.

    Dubrez, L., Berthelet, J., and Glorian, V. (2013). IAP proteins as targets for drugdevelopment in oncology. Onco. Targets Ther. 9, 12851304.

    1112 Chemistry & Biology 21, September 18, 2014 2014 Elsevier LtDueber, E.C., Schoeffler, A.J., Lingel, A., Elliott, J.M., Fedorova, A.V.,Giannetti, A.M., Zobel, K., Maurer, B., Varfolomeev, E., Wu, P., et al. (2011).Antagonists induce a conformational change in cIAP1 that promotes autoubi-quitination. Science 334, 376380.

    Engel, M., Hindie, V., Lopez-Garcia, L.A., Stroba, A., Schaeffer, F., Adrian, I.,Imig, J., Idrissova, L., Nastainczyk, W., Zeuzem, S., et al. (2006). Allostericactivation of the protein kinase PDK1 with low molecular weight compounds.EMBO J. 25, 54695480.

    Engelman, A., Kessl, J.J., and Kvaratskhelia, M. (2013). Allosteric inhibition ofHIV-1 integrase activity. Curr. Opin. Chem. Biol. 17, 339345.

    Erickson, R.I., Tarrant, J., Cain, G., Lewin-Koh, S.C., Dybdal, N., Wong, H.,Blackwood, E., West, K., Steigerwalt, R., Mamounas, M., et al. (2013). Toxicityprofile of small-molecule IAP antagonist GDC-0152 is linked to TNF-a pharma-cology. Toxicol. Sci. 131, 247258.

    Erlanson, D.A., Wells, J.A., and Braisted, A.C. (2004). Tethering: fragment-based drug discovery. Annu. Rev. Biophys. Biomol. Struct. 33, 199223.

    Fader, L.D., Malenfant, E., Parisien, M., Carson, R., Bilodeau, F., Landry, S.,Pesant, M., Brochu, C., Morin, S., Chabot, C., et al. (2014). Discovery of BI224436, a Noncatalytic Site Integrase Inhibitor (NCINI) of HIV-1. ACS Med.Chem. Lett. 5, 422427.

    Fenwick, C., Amad, M., Bailey, M.D., Bethell, R., Bos, M., Bonneau, P.,Cordingley, M., Coulombe, R., Duan, J., Edwards, P., et al. (2014). PreclinicalProfile of BI 224436, a Novel HIV-1 Non-Catalytic-Site Integrase Inhibitor.Antimicrob. Agents Chemother. 58, 32333244.

    Filippakopoulos, P., and Knapp, S. (2014). Targeting bromodomains: epige-netic readers of lysine acetylation. Nat. Rev. Drug Discov. 13, 337356.

    Filippakopoulos, P., Qi, J., Picaud, S., Shen, Y., Smith, W.B., Fedorov, O.,Morse, E.M., Keates, T., Hickman, T.T., Felletar, I., et al. (2010). Selectiveinhibition of BET bromodomains. Nature 468, 10671073.

    Flygare, J.A., Beresini, M., Budha, N., Chan, H., Chan, I.T., Cheeti, S., Cohen,F., Deshayes, K., Doerner, K., Eckhardt, S.G., et al. (2012). Discovery ofa potent small-molecule antagonist of inhibitor of apoptosis (IAP) proteinsand clinical candidate for the treatment of cancer (GDC-0152). J. Med.Chem. 55, 41014113.

    Ford, M.L., and Larsen, C.P. (2009). Translating costimulation blockade to theclinic: lessons learned from three pathways. Immunol. Rev. 229, 294306.

    Fry, D.C. (2012). Small-molecule inhibitors of protein-protein interactions: howto mimic a protein partner. Curr. Pharm. Des. 18, 46794684.

    Fulda, S., and Vucic, D. (2012). Targeting IAP proteins for therapeutic interven-tion in cancer. Nat. Rev. Drug Discov. 11, 109124.

    Fuller, J.C., Burgoyne, N.J., and Jackson, R.M. (2009). Predicting druggablebinding sites at the protein-protein interface. Drug Discov. Today 14, 155161.

    Gadek, T.R., Burdick, D.J., McDowell, R.S., Stanley, M.S., Marsters, J.C., Jr.,Paris, K.J., Oare, D.A., Reynolds, M.E., Ladner, C., Zioncheck, K.A., et al.(2002). Generation of an LFA-1 antagonist by the transfer of the ICAM-1 immu-noregulatory epitope to a small molecule. Science 295, 10861089.

    Gavenonis, J., Sheneman, B.A., Siegert, T.R., Eshelman, M.R., and Kritzer,J.A. (2014). Comprehensive analysis of loops at protein-protein interfacesfor macrocycle design. Nat. Chem. Biol. 10, 716722.

    Gehling, V.S., Hewitt, M.C., Vaswani, R.G., Leblanc, Y., Cote, A., Nasveschuk,C.G., Taylor, A.M., Harmange, J.C., Audia, J.E., Pardo, E., et al. (2013). Discov-ery, Design, and Optimization of Isoxazole Azepine BET Inhibitors. ACS Med.Chem. Lett. 4, 835840.

    Gold, M.G., Barford, D., and Komander, D. (2006). Lining the pockets ofkinases and phosphatases. Curr. Opin. Struct. Biol. 16, 693701.

    Guckian, K.M., and Scott, D.M. (2013). Inhibition of LFA-1/ICAM interaction forthe treatment of autoimmune diseases. In Protein-Protein Interactions in Drug

    Chemistry & Biology

    ReviewDiscovery, A. Domling, ed. (Weinheim: Wiley-VCH Verlag).

    Guo, W., Wisniewski, J.A., and Ji, H. (2014). Hot spot-based design of small-molecule inhibitors for protein-protein interactions. Bioorg. Med. Chem. Lett.24, 25462554.

    Hajduk, P.J., and Greer, J. (2007). A decade of fragment-based drug design:strategic advances and lessons learned. Nat. Rev. Drug Discov. 6, 211219.

    d All rights reserved

  • Hann, M.M., Leach, A.R., and Harper, G. (2001). Molecular complexity and itsimpact on the probability of finding leads for drug discovery. J. Chem. Inf.Comput. Sci. 41, 856864.

    Hartman, G.D., Egbertson, M.S., Halczenko, W., Laswell, W.L., Duggan, M.E.,Smith, R.L., Naylor, A.M., Manno, P.D., Lynch, R.J., Zhang, G., et al. (1992).Non-peptide fibrinogen receptor antagonists. 1. Discovery and design of exo-site inhibitors. J. Med. Chem. 35, 46404642.

    He, M.M., Smith, A.S., Oslob, J.D., Flanagan, W.M., Braisted, A.C., Whitty, A.,Cancilla, M.T., Wang, J., Lugovskoy, A.A., Yoburn, J.C., et al. (2005). Small-molecule inhibition of TNF-alpha. Science 310, 10221025.

    He, S., Senter, T.J., Pollock, J., Han, C., Upadhyay, S.K., Purohit, T., Gogliotti,R.D., Lindsley, C.W., Cierpicki, T., Stauffer, S.R., and Grembecka, J. (2014).High-affinity small-molecule inhibitors of the menin-mixed lineage leukemia(MLL) interaction closely mimic a natural protein-protein interaction. J. Med.Chem. 57, 15431556.

    Higueruelo, A.P., Schreyer, A., Bickerton, G.R., Pitt, W.R., Groom, C.R., andBlundell, T.L. (2009). Atomic interactions and profile of small molecules dis-rupting protein-protein interfaces: the TIMBAL database. Chem. Biol. DrugDes. 74, 457467.

    Hopkins, A.L., Keseru, G.M., Leeson, P.D., Rees, D.C., and Reynolds, C.H.(2014). The role of ligand efficiency metrics in drug discovery. Nat. Rev.Drug Discov. 13, 105121.

    Huang, Y., Rich, R.L., Myszka, D.G., and Wu, H. (2003). Requirement of boththe second and third BIR domains for the relief of X-linked inhibitorof apoptosis protein (XIAP)-mediated caspase inhibition by Smac. J. Biol.Chem. 278, 4951749522.

    Hwang, H., Vreven, T., Janin, J., and Weng, Z. (2010). Protein-protein dockingbenchmark version 4.0. Proteins 78, 31113114.

    Infante, J.R., Dees, E.C., Olszanski, A.J., Dhuria, S.V., Sen, S., Cameron, S.,and Cohen, R.B. (2014). Phase I dose-escalation study of LCL161, an oral in-hibitor of apoptosis proteins inhibitor, in patients with advanced solid tumors.J. Clin. Oncol. Published online August 11, 2014. http://dx.doi.org/10.1200/JCO.2013.52.3993.

    Johnson, D.K., and Karanicolas, J. (2013). Druggable protein interaction sitesare more predisposed to surface pocket formation than the rest of the proteinsurface. PLoS Comput. Biol. 9, e1002951.

    Keating, S.M., Clark, K.R., Stefanich, L.D., Arellano, F., Edwards, C.P., Bodary,S.C., Spencer, S.A., Gadek, T.R., Marsters, J.C., Jr., and Beresini, M.H. (2006).Competition between intercellular adhesion molecule-1 and a small-moleculeantagonist for a common binding site on the alphal subunit of lymphocytefunction-associated antigen-1. Protein Sci. 15, 290303.

    Khoo, K.H., Verma, C.S., and Lane, D.P. (2014). Drugging the p53 pathway: un-derstanding the route to clinical efficacy. Nat. Rev. Drug Discov. 13, 217236.

    Koehler, M.F., Bergeron, P., Choo, E.F., Lau, K., Ndubaku, C., Dudley, D.,Gibbons, P., Sleebs, B.E., Rye, C.S., Nikolakopoulos, G., et al. (2014). Struc-ture-guided rescaffolding of selective antagonists of BCL-XL. ACS Med.Chem. Lett. 5, 662667.

    Kozakov, D., Hall, D.R., Chuang, G.Y., Cencic, R., Brenke, R., Grove, L.E.,Beglov, D., Pelletier, J., Whitty, A., and Vajda, S. (2011). Structural conserva-tion of druggable hot spots in protein-protein interfaces. Proc. Natl. Acad.Sci. USA 108, 1352813533.

    Kussie, P.H., Gorina, S., Marechal, V., Elenbaas, B., Moreau, J., Levine, A.J.,and Pavletich, N.P. (1996). Structure of the MDM2 oncoprotein bound to thep53 tumor suppressor transactivation domain. Science 274, 948953.

    Labbe,C.M., Laconde,G.,Kuenemann,M.A., Villoutreix,B.O., andSperandio,O.(2013). iPPI-DB: a manually curated and interactive database of small non-pep-tide inhibitors of protein-protein interactions. Drug Discov. Today 18, 958968.

    Lao, B.B., Grishagin, I., Mesallati, H., Brewer, T.F., Olenyuk, B.Z., and Arora,P.S. (2014). In vivomodulation of hypoxia-inducible signaling by topographical

    Chemistry & Biology

    Reviewhelix mimetics. Proc. Natl. Acad. Sci. USA 111, 75317536.

    Lessard, J.A., and Crabtree, G.R. (2010). Chromatin regulatory mechanismsin pluripotency. Annu. Rev. Cell Dev. Biol. 26, 503532.

    Li, X., Srinivasan, S.R., Connarn, J., Ahmad, A., Young, Z.T., Kabza, A.M., Zui-derweg, E.R., Sun, D., and Gestwicki, J.E. (2013). Analogs of the AllostericHeat Shock Protein 70 (Hsp70) Inhibitor, MKT-077, as Anti-Cancer Agents.

    Chemistry & Biology 2ACS Med. Chem. Lett. 4. Published online September 17, 2013. http://dx.doi.org/10.1021/ml400204n.

    Liu, Z., Sun, C., Olejniczak, E.T., Meadows, R.P., Betz, S.F., Oost, T., Herr-mann, J., Wu, J.C., and Fesik, S.W. (2000). Structural basis for binding ofSmac/DIABLO to the XIAP BIR3 domain. Nature 408, 10041008.

    Livingston, D.J., Alexander, S., Bond, J., Briggs, T., Fraley, A., Hale, S.,Landsman, T., Martinelli, R., Shortsleeves, K., and Terrett, N. (2012). Identifica-tion and characterization of synthetic small molecule macrocycle antago-nists of human IL17A. http://www.ensemblediscovery.com/news/pdfs/ACR%20Poster%20DJL-v1-1%20(2).pdf.

    London, N., Raveh, B., Movshovitz-Attias, D., and Schueler-Furman, O. (2010).Can self-inhibitory peptides be derived from the interfaces of globular protein-protein interactions? Proteins 78, 31403149.

    London, N., Raveh, B., and Schueler-Furman, O. (2013). Druggable protein-protein interactionsfrom hot spots to hot segments. Curr. Opin. Chem.Biol. 17, 952959.

    Maurer, T., Garrenton, L.S., Oh, A., Pitts, K., Anderson, D.J., Skelton, N.J.,Fauber, B.P., Pan, B., Malek, S., Stokoe, D., et al. (2012). Small-molecule li-gands bind to a distinct pocket in Ras and inhibit SOS-mediated nucleotide ex-change activity. Proc. Natl. Acad. Sci. USA 109, 52995304.

    McLure, K.G., Gesner, E.M., Tsujikawa, L., Kharenko, O.A., Attwell, S., Cam-peau, E., Wasiak, S., Stein, A., White, A., Fontano, E., et al. (2013). RVX-208,an inducer of ApoA-I in humans, is a BET bromodomain antagonist. PLoSONE 8, e83190.

    Milroy, L.G., Grossmann, T.N., Hennig, S., Brunsveld, L., and Ottmann, C.(2014).Modulators of protein-protein interactions.Chem.Rev.114, 46954748.

    Mirguet, O., Gosmini, R., Toum, J., Clement, C.A., Barnathan, M., Brusq, J.M.,Mordaunt, J.E., Grimes, R.M., Crowe,M., Pineau, O., et al. (2013). Discovery ofepigenetic regulator I-BET762: lead optimization to afford a clinical candidateinhibitor of the BET bromodomains. J. Med. Chem. 56, 75017515.

    Miyazaki, M., Naito, H., Sugimoto, Y., Kawato, H., Okayama, T., Shimizu, H.,Miyazaki, M., Kitagawa, M., Seki, T., Fukutake, S., et al. (2013). Lead optimiza-tion of novel p53-MDM2 interaction inhibitors possessing dihydroimidazothia-zole scaffold. Bioorg. Med. Chem. Lett. 23, 728732.

    Ndubaku, C., Varfolomeev, E., Wang, L., Zobel, K., Lau, K., Elliott, L.O.,Maurer, B., Fedorova, A.V., Dynek, J.N., Koehler, M., et al. (2009). Antagonismof c-IAP and XIAP proteins is required for efficient induction of cell deathby small-molecule IAP antagonists. ACS Chem. Biol. 4, 557566.

    Nussinov, R., and Tsai, C.J. (2014). Unraveling structural mechanisms ofallosteric drug action. Trends Pharmacol. Sci. 35, 256264.

    Oltersdorf, T., Elmore, S.W., Shoemaker, A.R., Armstrong, R.C., Augeri, D.J.,Belli, B.A., Bruncko, M., Deckwerth, T.L., Dinges, J., Hajduk, P.J., et al.(2005). An inhibitor of Bcl-2 family proteins induces regression of solidtumours. Nature 435, 677681.

    Oost, T.K., Sun, C., Armstrong, R.C., Al-Assaad, A.S., Betz, S.F., Deckwerth,T.L., Ding, H., Elmore, S.W., Meadows, R.P., Olejniczak, E.T., et al. (2004).Discovery of potent antagonists of the antiapoptotic protein XIAP for thetreatment of cancer. J. Med. Chem. 47, 44174426.

    Ostrem, J.M., Peters, U., Sos, M.L., Wells, J.A., and Shokat, K.M. (2013).K-Ras(G12C) inhibitors allosterically control GTP affinity and effector interac-tions. Nature 503, 548551.

    Perkins, J.R., Diboun, I., Dessailly, B.H., Lees, J.G., and Orengo, C. (2010).Transient protein-protein interactions: structural, functional, and networkproperties. Structure 18, 12331243.

    Petsalaki, E., and Russell, R.B. (2008). Peptide-mediated interactions inbiological systems: new discoveries and applications. Curr. Opin. Biotechnol.19, 344350.

    Raj, M., Bullock, B.N., and Arora, P.S. (2013). Plucking the high hanging fruit:

    a systematic approach for targeting protein-protein interactions. Bioorg. Med.Chem. 21, 40514057.

    Ray-Coquard, I., Blay, J.Y., Italiano, A., Le Cesne, A., Penel, N., Zhi, J., Heil, F.,Rueger, R., Graves, B., Ding, M., et al. (2012). Effect of the MDM2 antagonistRG7112 on the P53 pathway in patients with MDM2-amplified, well-differenti-ated or dedifferentiated liposarcoma: an exploratory proof-of-mechanismstudy. Lancet Oncol. 13, 11331140.

    1, September 18, 2014 2014 Elsevier Ltd All rights reserved 1113

  • Roberts, A.W., Seymour, J.F., Brown, J.R., Wierda, W.G., Kipps, T.J., Khaw,S.L., Carney, D.A., He, S.Z., Huang, D.C., Xiong, H., et al. (2012). Substantialsusceptibility of chronic lymphocytic leukemia to BCL2 inhibition: results ofa phase I study of navitoclax in patients with relapsed or refractory disease.J. Clin. Oncol. 30, 488496.

    Rohn, J. (2012). Tensha therapeutics. Nat. Biotechnol. 30, 305.

    Rudin, C.M., Hann, C.L., Garon, E.B., Ribeiro de Oliveira, M., Bonomi, P.D.,Camidge, D.R., Chu, Q., Giaccone, G., Khaira, D., Ramalingam, S.S., et al.(2012). Phase II study of single-agent navitoclax (ABT-263) and biomarkercorrelates in patients with relapsed small cell lung cancer. Clin. Cancer Res.18, 31633169.

    Sadowsky, J.D., Burlingame, M.A., Wolan, D.W., McClendon, C.L., Jacobson,M.P., and Wells, J.A. (2011). Turning a protein kinase on or off from a singleallosteric site via disulfide trapping. Proc. Natl. Acad. Sci. USA 108, 60566061.

    Sattler, M., Liang, H., Nettesheim, D., Meadows, R.P., Harlan, J.E., Eberstadt,M., Yoon, H.S., Shuker, S.B., Chang, B.S., Minn, A.J., et al. (1997). Structure ofBcl-xL-Bak peptide complex: recognition between regulators of apoptosis.Science 275, 983986.

    Schaumberg, D.A., Sullivan, D.A., Buring, J.E., and Dana, M.R. (2003).Prevalence of dry eye syndrome among US women. Am. J. Ophthalmol.136, 318326.

    Schwab, N., Ulzheimer, J.C., Fox, R.J., Schneider-Hohendorf, T., Kieseier,B.C., Monoranu, C.M., Staugaitis, S.M., Welch, W., Jilek, S., Du Pasquier,R.A., et al. (2012). Fatal PML associated with efalizumab therapy: insightsinto integrin aLb2 in JC virus control. Neurology 78, 458467, discussion 465.

    Sharma, A., Slaughter, A., Jena, N., Feng, L., Kessl, J.J., Fadel, H.J., Malani,N., Male, F., Wu, L., Poeschla, E., et al. (2014). A new class of multimerizationselective inhibitors of HIV-1 integrase. PLoS Pathog. 10, e1004171.

    Shi, A., Murai, M.J., He, S., Lund, G., Hartley, T., Purohit, T., Reddy, G.,Chruszcz, M., Grembecka, J., and Cierpicki, T. (2012). Structural insightsinto inhibition of the bivalent menin-MLL interaction by small molecules inleukemia. Blood 120, 44614469.

    Shima, F., Yoshikawa, Y., Ye, M., Araki, M., Matsumoto, S., Liao, J., Hu, L.,Sugimoto, T., Ijiri, Y., Takeda, A., et al. (2013). In silico discovery of small-mole-cule Ras inhibitors that display antitumor activity by blocking the Ras-effectorinteraction. Proc. Natl. Acad. Sci. USA 110, 81828187.

    Shimaoka, M., Salas, A., Yang, W., Weitz-Schmidt, G., and Springer, T.A.(2003). Small molecule integrin antagonists that bind to the beta2 subunitI-like domain and activate signals in one direction and block them in the other.Immunity 19, 391402.

    Shire (2014). Shire plans to submit a new drug application to FDA for lifitegrastfor dry eye disease in adults. http://www.shire.com/shireplc/en/investors/investorsnews/irshirenews?id=951.

    Silvian, L.F., Friedman, J.E., Strauch, K., Cachero, T.G., Day, E.S., Qian, F.,Cunningham, B., Fung, A., Sun, L., Shipps, G.W., et al. (2011). Small moleculeinhibition of the TNF family cytokine CD40 ligand through a subunit fracturemechanism. ACS Chem. Biol. 6, 636647.

    Smith, M.C., and Gestwicki, J.E. (2012). Features of protein-protein interac-tions that translate into potent inhibitors: topology, surface area and affinity.Expert Rev. Mol. Med. 14, e16.

    Souers, A.J., Leverson, J.D., Boghaert, E.R., Ackler, S.L., Catron, N.D., Chen,J., Dayton, B.D., Ding, H., Enschede, S.H., Fairbrother, W.J., et al. (2013). ABT-199, a potent and selective BCL-2 inhibitor, achieves antitumor activity whilesparing platelets. Nat. Med. 19, 202208.

    Stockman, B.J., Kothe, M., Kohls, D., Weibley, L., Connolly, B.J., Sheils,A.L., Cao, Q., Cheng, A.C., Yang, L., Kamath, A.V., et al. (2009). Identifi-cation of allosteric PIF-pocket ligands for PDK1 using NMR-based frag-ment screening and 1H-15N TROSY experiments. Chem. Biol. Drug Des.73, 179188.Sun, Q., Burke, J.P., Phan, J., Burns, M.C., Olejniczak, E.T., Waterson, A.G.,Lee, T., Rossanese, O.W., and Fesik, S.W. (2012). Discovery of small mole-cules that bind to K-Ras and inhibit Sos-mediated activation. Angew. Chem.Int. Ed. Engl. 51, 61406143.

    Sun, H., Lu, J., Liu, L., Yang, C.Y., and Wang, S. (2014). Potent and selectivesmall-molecule inhibitors of cIAP1/2 proteins reveal that the binding of Smac

    1114 Chemistry & Biology 21, September 18, 2014 2014 Elsevier Ltmimetics to XIAP BIR3 is not required for their effective induction of cell deathin tumor cells. ACS Chem. Biol. 9, 9941002.

    Tovar, C., Graves, B., Packman, K., Filipovic, Z., Higgins, B., Xia, M., Tardell,C., Garrido, R., Lee, E., Kolinsky, K., et al. (2013). MDM2 small-moleculeantagonist RG7112 activates p53 signaling and regresses human tumors inpreclinical cancer models. Cancer Res. 73, 25872597.

    Tse, C., Shoemaker, A.R., Adickes, J., Anderson, M.G., Chen, J., Jin, S., John-son, E.F.,Marsh, K.C.,Mitten,M.J., Nimmer, P., et al. (2008). ABT-263: a potentand orally bioavailable Bcl-2 family inhibitor. Cancer Res. 68, 34213428.

    Vassilev, L.T., Vu, B.T., Graves, B., Carvajal, D., Podlaski, F., Filipovic, Z.,Kong, N., Kammlott, U., Lukacs, C., Klein, C., et al. (2004). In vivo activationof the p53 pathway by small-molecule antagonists of MDM2. Science 303,844848.

    Venkatesan, K., Rual, J.F., Vazquez, A., Stelzl, U., Lemmens, I., Hirozane-Kish-ikawa, T., Hao, T., Zenkner, M., Xin, X., Goh, K.I., et al. (2009). An empiricalframework for binary interactome mapping. Nat. Methods 6, 8390.

    Vidler, L.R., Brown, N., Knapp, S., and Hoelder, S. (2012). Druggability analysisand structural classification of bromodomain acetyl-lysine binding sites.J. Med. Chem. 55, 73467359.

    Vollmuth, F., Blankenfeldt, W., and Geyer, M. (2009). Structures of the dualbromodomains of the P-TEFb-activating protein Brd4 at atomic resolution.J. Biol. Chem. 284, 3654736556.

    Wang, Y., Coulombe, R., Cameron, D.R., Thauvette, L., Massariol, M.J.,Amon, L.M., Fink, D., Titolo, S., Welchner, E., Yoakim, C., et al. (2004).Crystal structure of the E2 transactivation domain of human papillomavirustype 11 bound to a protein interaction inhibitor. J. Biol. Chem. 279, 69766985.

    Wang, A.M., Miyata, Y., Klinedinst, S., Peng, H.M., Chua, J.P., Komiyama, T.,Li, X., Morishima, Y., Merry, D.E., Pratt, W.B., et al. (2013a). Activation ofHsp70 reduces neurotoxicity by promoting polyglutamine protein degrada-tion. Nat. Chem. Biol. 9, 112118.

    Wang, N., Majmudar, C.Y., Pomerantz, W.C., Gagnon, J.K., Sadowsky, J.D.,Meagher, J.L., Johnson, T.K., Stuckey, J.A., Brooks, C.L., 3rd, Wells, J.A.,and Mapp, A.K. (2013b). Ordering a dynamic protein via a small-moleculestabilizer. J. Am. Chem. Soc. 135, 33633366.

    Wei, S.J., Joseph, T., Sim, A.Y., Yurlova, L., Zolghadr, K., Lane, D., Verma, C.,and Ghadessy, F. (2013). In vitro selection of mutant HDM2 resistant to Nutlininhibition. PLoS ONE 8, e62564.

    Wells, J.A., and McClendon, C.L. (2007). Reaching for high-hanging fruit indrug discovery at protein-protein interfaces. Nature 450, 10011009.

    Wilson, C.G., and Arkin, M.R. (2011). Small-molecule inhibitors of IL-2/IL-2R:lessons learned and applied. Curr. Top. Microbiol. Immunol. 348, 2559.

    Winter, A., Higueruelo, A.P., Marsh, M., Sigurdardottir, A., Pitt, W.R., and Blun-dell, T.L. (2012). Biophysical and computational fragment-based approachesto targeting protein-protein interactions: applications in structure-guideddrug discovery. Q. Rev. Biophys. 45, 383426.

    Wong, H., Gould, S.E., Budha, N., Darbonne, W.C., Kadel, E.E., 3rd, La, H.,Alicke, B., Halladay, J.S., Erickson, R., Portera, C., et al. (2013). Learningand confirming with preclinical studies: modeling and simulation in the dis-covery of GDC-0917, an inhibitor of apoptosis proteins antagonist. DrugMetab. Dispos. 41, 21042113.

    Yang, C., and Novack, D.V. (2013). Anti-cancer IAP antagonists promote bonemetastasis: a cautionary tale. J. Bone Miner. Metab. 31, 496506.

    Zhang, Z., Ding, Q., Liu, J.J., Zhang, J., Jiang, N., Chu, X.J., Bartkovitz, D., Luk,K.C., Janson, C., Tovar, C., et al. (2014). Discovery of potent and selectivespiroindolinone MDM2 inhibitor, RO8994, for cancer therapy. Bioorg. Med.Chem. 22, 40014009.

    Zhao, Y., Yu, S., Sun, W., Liu, L., Lu, J., McEachern, D., Shargary, S., Bernard,

    Chemistry & Biology

    ReviewD., Li, X., Zhao, T., et al. (2013). A potent small-molecule inhibitor of theMDM2-p53 interaction (MI-888) achieved complete and durable tumor regression inmice. J. Med. Chem. 56, 55535561.

    Zhong, M., Gadek, T.R., Bui, M., Shen, W., Burnier, J., Barr, K.J., Hanan, E.J.,Oslob, J.D., Yu, C.H., Zhu, J., et al. (2012). Discovery and development ofpotent LFA-1/ICAM-1 antagonist SAR 1118 as an ophthalmic solution fortreating dry eye. ACS Med. Chem. Lett. 3, 203206.

    d All rights reserved

    Small-Molecule Inhibitors of Protein-Protein Interactions: Progressing toward the RealityPrimary Peptide Epitopes: Short, Continuous, Linear PeptidesLFA1IAPsBromodomainsHIV IntegraseProspects for Primary Sequence Epitopes

    Secondary Structure Epitopes: Helix, Sheet, or Extended PeptidesBCL FamilyMDM2Selected Emerging Examples of Secondary Structure Epitope InhibitorsProspects for Secondary Structure Epitope Design

    Tertiary Structural Epitopes: Discontinuous Binding SitesInterleukin-17IL-2 and HPV-11Prospects for Tertiary Epitope Binders

    Allosteric Mechanisms of InhibitionHow Has Our Understanding of PPIs Evolved in the Past 10 Years?What We Expect to See More of in the Near FutureAcknowledgmentsReferences