You are on page 1of 10

Article

pubs.acs.org/accounts

Protein Domain Mimics as Modulators of ProteinProtein


Interactions
Published as part of the Accounts of Chemical Research special issue Chemical Biology of Peptides.
Nicholas Sawyer, Andrew M. Watkins, and Paramjit S. Arora*
Department of Chemistry, New York University, 100 Washington Square East, New York, New York 10003, United States

CONSPECTUS: Proteinprotein interactions (PPIs) are ubiquitous in biological systems and


often misregulated in disease. As such, specic PPI modulators are desirable to unravel complex PPI
pathways and expand the number of druggable targets available for therapeutic intervention.
However, the large size and relative atness of PPI interfaces make them challenging molecular
targets. This Account describes our systematic approach using secondary and tertiary protein
domain mimics (PDMs) to specically modulate PPIs.
Our strategy focuses on mimicry of regular secondary and tertiary structure elements from one of
the PPI partners to inspire rational PDM design. We have compiled three databases (HIPPDB,
SIPPDB, and DIPPDB) of secondary and tertiary structures at PPI interfaces to guide our designs
and better understand the energetics of PPI secondary and tertiary structures. Our eorts have
focused on three of the most common secondary and tertiary structures: -helices, -strands, and
helix dimers (e.g., coiled coils).
To mimic -helices, we designed the hydrogen bond surrogate (HBS) as an isosteric PDM and the
oligooxopiperazine helix mimetic (OHM) as a topographical PDM. The nucleus of the HBS approach is a peptide macrocycle in
which the N-terminal i, i + 4 main-chain hydrogen bond is replaced with a covalent carboncarbon bond. In mimicking a main-
chain hydrogen bond, the HBS approach stabilizes the -helical conformation while leaving all helical faces available for
functionalization to tune binding anity and specicity. The OHM approach, in contrast, envisions a tetrapeptide to mimic one
face of a two-turn helix. We anticipated that placement of ethylene bridges between adjacent amides constrains the tetrapeptide
backbone to mimic the i, i + 4, and i + 7 side chains on one face of an -helix.
For -strands, we developed triazolamers, a topographical PDM where the peptide bonds are replaced by triazoles. The triazoles
simultaneously stabilize the extended, zigzag conformation of -strands and transform an otherwise ideal protease substrate into a
stable molecule by replacement of the peptide bonds.
We turned to a salt bridge surrogate (SBS) approach as a means for stabilizing very short helix dimers. As with the HBS
approach, the SBS strategy replaces a noncovalent interaction with a covalent bond. Specically, we used a bis-triazole linkage
that mimics a salt bridge interaction to drive helix association and folding. Using this approach, we were able to stabilize helix
dimers that are less than half of the length required to form a coiled coil from two independent strands.
In addition to demonstrating the stabilization of desired structures, we have also shown that our designed PDMs specically
modulate target PPIs in vitro and in vivo. Examples of PPIs successfully targeted include HIF1/p300, p53/MDM2, Bcl-xL/Bak,
Ras/Sos, and HIV gp41. The PPI databases and designed PDMs created in these studies will aid development of a versatile set of
molecules to probe complex PPI functions and, potentially, PPI-based therapeutics.

INTRODUCTION
Proteinprotein interaction (PPI) modulators have vast
regularity and energetic importance of these secondary and
tertiary structures led to the hypothesis that protein domain
potential as specic probes to elucidate complex biological mimics (PDMs) that imitate interfacial PPI secondary or
pathways and as therapeutics to correct misregulated PPIs in tertiary structures would act as potent PPI inhibitors.
disease.1,2 Developing a systematic strategy to discover high To test this hypothesis, many groups have developed elegant
anity PPI modulators is an area of intensive research.3,4 Our approaches to mimic protein secondary and tertiary struc-
eorts have focused on the role and mimicry of secondary (e.g., tures.827 One common approach is to covalently stabilize
-helix and -strand) and tertiary (e.g., coiled-coil) structures isolated peptides, which are typically unstructured in solution.
to inhibit PPI interfaces (Figure 1). Estimates based on peptide Preorganization of the peptide in its active conformation may
backbone dihedral angles suggest that PPI interfaces are improve upon the anity of unstructured peptides for their
composed of roughly equal measures of regular secondary target protein. A second common approach mimics protein
structures (-helices and -strands) and nonregular structures.5 surfaces with nonpeptidic scaolds that can approximate the
Regular structural elements display side chain groups in dened
orientations, often burying large surface areas and contributing Received: March 15, 2017
signicantly to the overall PPI binding energy.6,7 The structural Published: May 31, 2017

2017 American Chemical Society 1313 DOI: 10.1021/acs.accounts.7b00130


Acc. Chem. Res. 2017, 50, 13131322
Accounts of Chemical Research Article

Figure 1. Examples of (A) -helices, (B) -strands, and (C) helix dimers at proteinprotein interaction interfaces. PDB codes: (A) 1BXL (Bcl-xL/
Bak), 1YCR (p53/MDM2); (B) 1OY3 (NF-B homodimer), 1F3U (Rap30/Rap74); (C) 2IW5 (LSD1/CoREST), 3CL3 (vFLIP/IKK).

side chain geometry of binding epitopes. Both approaches have


been successful for many protein targets, particularly in the
mimicry of -helices, -strands, and -hairpins.28
This Account reviews our eorts in (1) the construction and
analysis of an extensive PPI database and (2) the design and
evaluation of PPI inhibitors inspired by PPIs identied in the
database. We will discuss the database and design of PPI
inhibitors in reference to three specic secondary and tertiary
structures: -helices, -strands, and helix dimers. For each
structure, we highlight specic ndings from database analysis,
scaolds used for PPI modulator design, and a case study for a
biological system of interest. Though many of our PPI
modulators were designed prior to database construction, the
database continues to inform our target selections and ligand
design.

RESULTS AND DISCUSSION


-Helical PDMs
-Helices are commonly observed at PPI interfaces, particularly
at homodimeric complexes, where approximately 40% of all
interface residues are in the -helical conformation.5 Given the
prevalence of -helices at PPI interfaces, we systematically
analyzed the Protein Data Bank (PDB) to construct a database
(Helical Interfaces in ProteinProtein Interactions Database,
or HIPPDB) of -helices that contribute signicantly to the
binding free energy of PPI interfaces.2931 In an attempt to
preselect for targetable PPIs, all of our databases are
constructed using cutos for the minimum number of hotspot
residues (G 1 Rosetta energy unit) in the structural
element. Analysis of HIPPDB reveals that only a small fraction
(14%) of helical PPI interfaces have their hotspots concen- Figure 2. Distribution of -helix hotspots at PPI interfaces. Hotspot
trated into binding pockets that might be blocked with frequency is plotted as a bar graph (A) and on an idealized -helix (B).
traditional small molecules. In the remaining interfaces, The data show increases in hotspot frequency every 34 residues,
corresponding to alignment of hotspots on a single -helical face
hotspots are more widely distributed. (helical faces indicated with black arcs). The color legend depicts
In designing a PPI -helix mimic, the major decision is which fractional occurrence of hot spot residues (G for alanine mutation
molecular architecture to use as a scaold for the functional > 1 kcal/mol) at each positon.
groups that will determine the molecules binding anity and
specicity. This selection process is informed by two critical
factors: (1) the spacing of hotspots with respect to each other
and (2) alignment of hotspots with respect to the helical faces. of HIPPDB helices.31 Hotspot alignment on a single helical face
Single-Face Helix Mimics. Analysis of PPI -helices led to the hypothesis that a rigid topographical mimic that
reveals a hotspot periodicity of 34 residues, corresponding reproduces the -helical spacing of functional groups, but not
to hotspot alignment on a single helical face (Figure 2). necessarily the peptide backbone, would function as a PPI
Hotspot residues are aligned in this way in approximately 60% inhibitor with similar ecacy to a helical peptide. The success
1314 DOI: 10.1021/acs.accounts.7b00130
Acc. Chem. Res. 2017, 50, 13131322
Accounts of Chemical Research Article

Figure 3. Design rationale for hydrogen bond surrogate (HBS) helices and oligooxopiperazine helix mimetics (OHMs). (A) (Left) An -helix
projects its i, i + 4, and i + 7 side chains on a single helical face. A tetrapeptide can be modeled to mimic these side chain projections (center left) and
synthetically realized using half-chair oxopiperazine rings (center). Overlay of the OHM model with an idealized -helix (center right) shows
excellent mimicry of a single helical face. The general synthetic scheme for OHMs (right): (a) oNBS-Cl, 2,4,6-collidine; (b) 2-bromoethanol,
triphenylphosphine (PPh3), diisopropylazodicarboxylate (DIAD); (c) base (e.g., 1,8-diazabicyclo(5.4.0)undec-7-ene (DBU)); (d) 2-
mercaptoethanol, DBU. (B) (Left) The helix-nucleating N-terminal hydrogen bond can be transformed into a covalent bond (center left),
specically a carboncarbon bond (center). Overlay of the HBS helix crystal structure with an -helix shows excellent mimicry. The general
synthetic scheme for HBS helices (right): (a) SPPS for secondary amine, (b) SPPS, (c) 4-pentenoic acid, N,N-diisopropylcarbodiimide (DIC); (d)
ring-closing metathesis (RCM).

of many conformationally rigid scaolds that mimic -helix single -helical face. Isolated -helical peptides, on the other
topography solidly supports this hypothesis.15,18,20,32 hand, could eectively mimic their protein counterparts.
In our laboratory, we have developed the oligooxopiperazine However, short peptides are conformationally labile, proteolyti-
helix mimetic (OHM) as a topographical helix mimetic (Figure cally unstable and generally bind with weak anity, in part
3A).22,3234 We rationalized that if the goal is to mimic a two- because of the entropic penalty for folding upon binding.36
turn -helix that displays three hotspot residues on the same In light of these challenges, dierent strategies have been
helical face, i.e. the i, i + 4, and i + 7 positions, a tetrapeptide devised to covalently constrain peptides into the -helical
may be sucient. We created a tetrapeptide model that spans conformation.20,24 We have extensively evaluated a hydrogen
the length of a two-turn -helix and positions its i, i + 1, and i + bond surrogate (HBS) approach as one such strategy (Figure
3 residues in positions equivalent to the i, i + 4, and i + 7 3B).11,3743 The crux of this approach is the replacement of a
residues of an -helix. To stabilize the requisite peptide helix-nucleating N-terminal backbone hydrogen bond with an
conformation, we used oxopiperazine rings obtained by isosteric, covalent carboncarbon bond.44,45 Covalent nuclea-
bridging neighboring amides with ethylene. Though the tion of the -helical conformation leads to helix propagation46
OHM backbone does not trace that of the -helix, side chains and enhances the proteolytic stability of HBS helices compared
are presented in i, i + 4, and i + 6/i + 7 positions. The optimal to their respective unconstrained peptide analogs.47 Impor-
-helix side chain mimicry requires the half-chair geometry tantly, by replacing a main-chain hydrogen bond, the number of
observed in the piperazine ring. OHMs are eciently side chains available for functionalization is maximized while
synthesized on solid-phase in successive dipeptide units by minimizing the size of the covalent constraint.
alkylating an o-nitrobenzenesulfonyl (oNBS) activated dipep- HBS helices are eciently synthesized by solid-phase peptide
tide under Mitsunobu conditions22,35 followed by base- synthesis (SPPS) with three modications (Figure 3B). The
catalyzed cyclization (Figure 3A). In addition to providing rst modication is the incorporation of an N-allyl amino acid
conformational stability, the ethylene bridges confer proteolytic residue at the i + 4 position. N-Allyl amino acids can be
stability by protecting the peptide bonds as tertiary amides. The synthesized in solution or on resin, typically using the oNBS
utility of the OHM scaold for specic, high anity binding protecting group to activate the amine for monoalkylation.35
(Kd = 30500 nM) has been demonstrated with various PPI The second modication for HBS synthesis is the solid-phase
targets.22,32 incorporation of a 4-pentenoic acid derivative in place of the i
Multi-Face Helix Mimics. The remaining 40% of high and i + 1 residues.11,43 The terminal alkene thus mimics the
anity PPI -helices in HIPPDB engage their protein partners carbonyl group of the i residue. Finally, the third and nal
with two or three helical faces.30 Mimicry of these helices with modication of SPPS for HBS synthesis is ring-closing
topographical -helical mimetics is at best incomplete because metathesis (RCM)48 between the alkenes at the i and i + 4
those molecules display functional groups equivalent to only a equivalent positions, forming a covalent macrocyclic -helix
1315 DOI: 10.1021/acs.accounts.7b00130
Acc. Chem. Res. 2017, 50, 13131322
Accounts of Chemical Research Article

Figure 4. In vivo modulation of hypoxia signaling by designed -helix PDMs. The center image shows the CBP CH1 domain (surface
representation) interacting with HIF1 CTAD (ribbon) (PDB 1L8C). The upper panels show the HIF1 helix (center) with hotspots mimicked
using OHM (left) and HBS (right) PDMs and Kd values for each PDM and negative controls. The Kd value for the unconstrained peptide
counterpart, Ac-ELARALDQ-NH2, is 6 M. The lower panels show reduction in mouse xenograft tumor volume after a sequence of parenteral 13
15 mg/kg injections with the HIF1-derived OHM (left) or HBS (right) PDMs.

nucleus. The RCM step has been optimized for solid-phase the cognate p300/CBP CH1 domain with submicromolar
microwave conditions and is compatible with most protecting anity.49,51
groups commonly used in Fmoc/tBu-based SPPS.11,37 As with Motivated by the success of the HBS helices, a series of
OHMs, a variety of HBS helices have been established as HIF1-based OHMs were also designed to mimic the hotspot
inhibitors of specic helix-mediated PPIs.38,47,4952 residues of one of the HIF1 CTAD helices whose hotspots lie
Case Study: Inhibition of the HIF1-p300/CBP PPI. A on a single helical face.32 Hotspot mimicry using the OHM
pertinent illustration of OHMs and HBS helices as PPI scaold results in nanomolar binding anity for the CH1
inhibitors both in vitro and in vivo is exemplied by our studies domain, suggesting that these OHMs capture the majority of
on the inhibition of complex formation between the hypoxia the high-anity elements of the HIF1 helix. HBS and OHM
inducible factor 1 (HIF1) transcription factor and the analogues lacking computationally predicted hotspot residues
transcriptional coactivator homologues p300 and CBP (Figure display negligible binding to the p300/CBP CH1 domain,
4). 32,51 The HIF1-p300/CBP PPI has been studied suggesting that both types of helix mimetics recognize their
extensively because of the central role of HIF1 in regulating target domain at the intended binding site in a sequence-
transcription in hypoxic environments, such as solid tumors. specic manner.
HIF1 engages p300/CBP primarily through its C-terminal In addition to successfully binding to p300/CBP, the
transactivation domain (CTAD).53,54 While the HIF1 CTAD designed HBS helices and OHMs also eciently downregulate
is intrinsically disordered in isolation, it folds upon binding to HIF1-mediated transcription activation in cell culture.32,49,51
the CH1 domain of p300/CBP, forming two distinct - Transcription of specic HIF target genes, such as VEGFA,
helices.55,56 Computational and experimental studies have LOX, and GLUT1,57 is downregulated upon treatment with
demonstrated an essential role for both helices in high-anity designed HBS helices or OHMs. Gene expression proling of
interaction between HIF1 and p300/CBP. treated A549 cells under hypoxic conditions revealed 2-fold
Given the importance of the HIF1 helices in p300/CBP up- or downregulation of up to 600 transcripts, including
binding, we hypothesized that mimics of either -helix would transcripts from many genes involved in the hallmarks of
be sucient to inhibit the HIF1-p300/CBP PPI. HBS helices cancer.58
based on the HIF1 CTAD sequence were designed and found The HBS and OHM mimics of the HIF1 helix also proved
to form stable -helices in solution. Each HBS helix binds to to be eective in reducing tumor growth in vivo in mouse
1316 DOI: 10.1021/acs.accounts.7b00130
Acc. Chem. Res. 2017, 50, 13131322
Accounts of Chemical Research Article

xenograft models (Figure 4).32,51 For HBS helix treatment, 786-


O human renal cell carcinoma xenografts in BALB/c mice
showed greater than 50% volume reduction over a 4-week
treatment course. Similar reduction in tumor volume was
observed in BALB/c mice with MDA-MB-231 human breast
adenocarcinoma xenografts upon similar OHM treatment.
Daily weight measurements suggested no toxicity for either
compound. Post-treatment tumor imaging with the near-
infrared tumor contrast agent IR-783 revealed reduced tumor
volume and no evidence of tumor metastasis.
The success of HBS helices in reducing the HIF1-p300/
CBP interaction in vitro and reducing tumor volume in vivo
studies prompted further investigation of the potential of
HIF1 helix mimetics to reduce other cancers associated with
p300/CBP PPIs. The CH1 domain of p300/CBP is known to
interact with numerous transcription factors, including p53.59
Figure 5. Distribution of -strand hotspots at PPI interfaces. Hotspot
In human papillomavirus-positive head and neck squamous cell frequency is plotted as a bar graph (A) and on an idealized -strand
carcinoma (HPV-positive HNSCC), the p300/CBP-p53 PPI is (B). With the exception of position i + 2, the data show a minimal
disrupted by the viral protein E6, reducing p53 stability and increase in hotspot frequency every 2 residues, corresponding to
transcriptional activity.60 It was thus hypothesized that HIF1 alignment of hotspots on a single -strand face. The color legend
helix mimetics might block the p300-E6 PPI and reactivate depicts fractional occurrence of hot spot residues (G for alanine
p53.61 Indeed, a HBS helix derived from one of the HIF1 mutation > 1 kcal/mol) at each positon.
CTAD helices reactivates p53 in HPV-positive HNSCC but not
HPV-negative HNSCC. The HBS helix also potentiates the
eects of cisplatin, suggesting that combination therapy with approaches have been reviewed recently,64,70 so we will limit
the HIF1 HBS helix and reduced cisplatin dosing may our discussion here to the triazolamer scaold developed in our
improve the toxicity prole of cisplatin treatment. Interestingly, laboratory.
a HBS helix derived from the other HIF1 CTAD helix does Triazolamers. To design conformationally and proteolyti-
not reactivate p53, suggesting partial but incomplete overlap of cally stable, single-strand -strand mimics, we developed the
triazolamer scaold. In this scaold, the peptide backbone is
the E6 and HIF1 binding sites on p300/CBP.
replaced with 1,3-substituted 1,2,3-triazoles.13,67,71 This scaold
Overall, these studies highlight the utility of HBS helices and
recapitulates peptide side chain interactions but cannot
OHMs as modulators of p300/CBP PPIs in vitro and in vivo.
hydrogen bond in the same way as a peptide backbone. The
Further studies are expected to provide insight into the
triazole moiety is synthesized from amino acid-derived alkynes
specicity of HBS helix/OHM PPI targeting.
and azides.72 Peptide backbone replacement protects these
-Strand PDMs molecules from protease degradation. Stabilization of the -
The similar frequency of -helices and -strands at strand conformation is driven by dipoledipole interactions
heterodimeric interfaces (26% and 24%, respectively) under- between triazoles, which have large dipole moments. Both
scores the essential role of -strands in PPIs.5 Thus, to solution NMR and X-ray diraction studies have conrmed
complement HIPPDB, we constructed SIPPDB (Strand zigzag triazolamer conformations that mimic the pleats
Interfaces in ProteinProtein Interactions Database), a data- observed in -strands.13,73 (Figure 6A, B). The i, i + 2 distance
base of -strands that contribute signicantly to the binding between triazolamer side chain Cs (7.9 ) is similar to the i, i
free energy of their respective PPI interfaces.62 + 2 distance in -strands (7.2 ), making this an appealing
With the exception of position i + 2, the expected periodicity scaold for the design of not only PPI modulators but also
of -strand hotspots on a single face is not as evident as the protease inhibitors.
hotspot periodicity of 34 observed for -helices (Figure 5). Case Study: Inhibition of HIV-1 Protease. Buoyed by the
PPI -strands also tend to be shorter than PPI -helices. Our promising biophysical properties of triazolamers, we conducted
survey of PPI -strands revealed that only one-quarter of these initial studies to evaluate triazolamers as PPI inhibitors.
strands engage their protein partners with hotspots on a single Preliminary investigations focused on HIV-1 protease
face. In contrast, almost one-third of PPI -strands have (HIVPR), as protease-substrate interactions can be thought of
binding energy contributions distributed equally across their 2 as transient PPIs (Figure 6C, D).67 Initial designs were based
faces. on two tetrapeptide inhibitors that adopt a strand-like
One signicant challenge with -strand PPI modulators is conformation in the protease active site.74 FRET-based activity
that -strands are ideal protease substrates.63 To overcome this assays75 revealed robust inhibition of HIV-1 protease by the
challenge and the general conformational instability of isolated designed triazolamers and demonstrate the potential of this
peptides, many approaches have been investigated to scaold for the rational design of inhibitors involving a -
simultaneously stabilize the -strand conformation, protect or strand. Further evaluation of triazolamer-based PPI inhibitors
replace the peptide backbone, and retain critical functional are in progress.
groups.64 These approaches can be roughly classied into two Helix Dimer PDMs
categories: use of nonpeptidic amino acid analogs to stabilize The results above demonstrate the advantages of secondary
individual -strands9,13,6568 or protection of individual strands structure mimics as PPI inhibitors. Nonetheless, many PPI
in the context of a -hairpin stabilized by noncovalent interfaces do not possess a single, energetically dominant -
interactions or macrocyclization (Figure 6A).12,23,69 Both helix or -strand. Instead, these interfaces often rely on tertiary
1317 DOI: 10.1021/acs.accounts.7b00130
Acc. Chem. Res. 2017, 50, 13131322
Accounts of Chemical Research Article

Figure 6. Triazolamers as -strand mimetics. NMR (A) and X-ray crystallographic (B) data support a zigzag geometry for triazolamers that mimics
-strands. The solid and dashed arrows in panel (A) indicate strong and weak ROESY cross peaks, respectively, in the NMR spectrum. (C) The
interaction of the tetrapeptide inhibitor L-400,417 (stick representation) with HIV-1 protease (HIVPR) dimer (green cartoon) was used to guide
triazolamer HIVPR inhibitor design (overlay on right with L-400,417 in green and a triazolamer design in purple). (D) Examples of a triazolamer
inhibitor (left) and negative control (right), showing IC50s for HIVPR protease inhibition .

structures to drive complex formation, where hotspots are


distributed over multiple secondary structure motifs.3
One of the simplest and best studied examples of tertiary
structure is the coiled coil.7678 Coiled coils form a helical
dimer and can occur in either parallel or antiparallel orientation.
We constructed DIPPDB (Dimeric Interfaces in Protein
Protein Interactions Database) to systematically examine helix
dimers at PPI interfaces.79 To construct the database, we
applied two separate lters to HIPPDB. The rst lter
examined the geometry between pairs of interface helices,
retaining only helix pairs with (1) an antiparallel or parallel
orientation and (2) an interhelical distance compatible with
typical coiled coil hydrophobic cores. The second lter
examined the energetics of the PPI interface helices and
retained only those helix pairs where both helices contributed
signicantly to the calculated binding energy. The overlap Figure 7. Distribution of helix dimer hotspots at PPI interfaces. The
between these two ltered sets was rened to yield DIPPDB. frequency of particular spacings between the rst and last hotspots is
On average, the rst and last hotspot residues for DIPPDB plotted as a bar graph (A) and on an idealized coiled coil (B). The
helix dimers are separated by approximately 13 residues (or two spacing between rst and last hotspots is usually shorter than 3
helical turns, see Figure 7).79 This average distance is larger heptads. The color legend depicts fractional occurrence of hot spot
than the analogous hotspot separation observed in isolated PPI residues (G for alanine mutation > 1 kcal/mol) at each spacing.
-helices (approximately 7 residues or one helical turn).
Minimal Coiled Coil Mimetics Using the Salt Bridge
Nonetheless, this spacing represents a relatively compact
Surrogate (SBS) Approach. In light of these ndings, we
hotspot distribution compared with the minimum length of
embarked upon a series of studies to identify chemical
approximately 21 residues required to form a stable coiled coil strategies for stabilizing minimal coiled coil mimetics as PPI
in solution from two separate peptide chains.80,81 inhibitors (Figure 8A).26 Ultimately, we obtained the desired
We also analyzed the chemistry of the dimerization interface stability for an 18-residue helix dimer (9 residues per helix) by
between PPI helix dimers. Interestingly, we observed that these employing azidealkyne click chemistry to covalently link
interfaces often deviate from idealized hydrophobic coiled coil positions that typically form salt bridge interactions. This salt
packing. This deviation from idealized packing may represent bridge surrogate (SBS) approach is analogous to the HBS
functional constraints imposed by the PPI interface, which can approach, in that it replaces a stabilizing noncovalent
involve helix dimerization interface residues.79 interaction with a covalent interaction. To synthesize SBS
1318 DOI: 10.1021/acs.accounts.7b00130
Acc. Chem. Res. 2017, 50, 13131322
Accounts of Chemical Research Article

Figure 8. Salt bridge surrogate (SBS) helix dimers as PPI inhibitors. (A) Dierent bis-triazole SBS linkages (middle left) were synthesized as part of
an idealized coiled coil (top left) and monitored by circular dichroism (CD, bottom left), demonstrating that a propargyl ether-azidolysine linkage
(green) favored helix formation. A helical wheel diagram and structural model based on NMR restraints are shown on the right. (B) The SBS
approach was applied to the NHR2 coiled coil (top left). Though the SBS alone did not stabilize the helix dimer as monitored by CD, combining the
SBS with optimization of the helix dimer interface yields a minimal helix dimer with signicant helicity and native-like binding anity for N2B
(NHR2-N2B Kd = 356 M). Binding anity was further enhanced with a disulde bridge.

helix dimers, an azidolysine residue is incorporated at a salt PDMs that target specic PPIs with high anity and specicity.
bridge position in one peptide by SPPS. Propargyl ether is then Our computational databases (HIPPDB, SIPPDB, and
coupled under copper(I)-catalyzed alkyneazide cycloaddition DIPPDB) provide a foundation for understanding the
(CuAAC) conditions.72,82 The remaining free alkyne is then energetics of secondary and tertiary structures at PPI interfaces.
coupled to a Fmoc-protected azidolysine amide, which is used The overarching theme of our experimental approach is the
for further SPPS. introduction of new covalent bonds to drive peptides into
Case Study: The NHR2/N2B Interaction. Having specic bioactive conformations. In the HBS and SBS
established the SBS approach for stabilizing an idealized 18- approaches, the new covalent bonds stabilize helical structure
residue helix dimer, we examined the use of this method to by replacing native hydrogen bonding or ionic interactions. For
constrain a biological helix dimer, namely a minimal coiled coil
triazolamers and OHMs, new covalent bonds drive the
derived from the Nervy homology 2 (NHR2) protein (Figure
formation of topographical -strand and -helix scaolds,
8B). NHR2 binds to the NHR2-binding (N2B) motif of E-
proteins through an antiparallel coiled coil and plays an respectively. We have successfully translated information from
essential role in leukemogenesis.83 Adding the SBS cross-link to several of our database entries into eective PPI inhibitors that
an otherwise native NHR2 sequence did not result in stable target specic PPIs in vitro. While some of the PDMs described
helix formation or detectable binding to the isolated N2B motif. herein have shown in vivo potential, potent engagement of
Optimization of the helix dimerization interface was required to intracellular targets remains a key challenge. Interplay between
obtain a stable helix dimer that displayed a native-like binding computational PPI analysis and evaluation of synthetic PPI
anity to the N2B motif. Further stabilization of the dimer inhibitors will allow for the development of a versatile set of
interface by a cysteine-mediated disulde bridge further molecules to probe complex PPI functions and to inspire the
improved both helicity and binding anity for the N2B target. design of new PPI-based therapeutics.


Taken together, these experiments highlight the combination of
the SBS approach with hydrophobic core optimization to create AUTHOR INFORMATION
potent helix dimer PPI inhibitors. Importantly, the SBS
approach is sequence independent and provides access to Corresponding Author
helix dimer mimetics that are less than half the size of the *E-mail: arora@nyu.edu.
smallest stable coiled coil,81 improving the prospects of these ORCID
compounds for intracellular applications.

Paramjit S. Arora: 0000-0001-5315-401X


CONCLUSIONS Present Address

In summary, we have described a systematic strategy that A.M.W.: Department of Biochemistry, Stanford University,
combines computational and experimental studies to generate 279 Campus Drive, Stanford, CA 94305.
1319 DOI: 10.1021/acs.accounts.7b00130
Acc. Chem. Res. 2017, 50, 13131322
Accounts of Chemical Research Article

Author Contributions (3) Arkin, M. R.; Tang, Y.; Wells, J. A. Small-molecule inhibitors of
protein-protein interactions: progressing toward the reality. Chem. Biol.
The manuscript was written through contributions of all
2014, 21, 11021114.
authors. All authors have given approval to the nal version of (4) Raj, M.; Bullock, B. N.; Arora, P. S. Plucking the high hanging
the manuscript. fruit: A systematic approach for targeting protein-protein interactions.
Notes Bioorg. Med. Chem. 2013, 21, 40514057.
(5) Guharoy, M.; Chakrabarti, P. Secondary structure based analysis
The authors declare the following competing nancial
and classification of biological interfaces: identification of binding
interest(s): P.S.A. is a co-inventor on patent applications on motifs in protein-protein interactions. Bioinformatics 2007, 23, 1909
secondary and tertiary structure mimics described in this paper. 1918.
He is also cofounder of Inthera Bioscience, which is pursuing (6) Clackson, T.; Wells, J. A. A hot spot of binding energy in a
therapeutic applications of HBS and OHM technologies. hormone-receptor interface. Science 1995, 267, 383386.
(7) Nooren, I. M. A.; Thornton, J. M. Diversity of protein-protein
Biographies
interactions. EMBO J. 2003, 22, 34863492.
Nicholas Sawyer received a B.S. in Biochemistry from Rutgers (8) Appella, D. H.; Christianson, L. A.; Klein, D. A.; Powell, D. R.;
University in 2010 and a Ph.D. in Molecular Biophysics and Huang, X.; Barchi, J. J.; Gellman, S. H. Residue-based control of helix
Biochemistry from Yale University in 2016. He is currently a shape in -peptide oligomers. Nature 1997, 387, 381384.
postdoctoral fellow at New York University, where he is investigating (9) Nowick, J. S.; Chung, D. M.; Maitra, K.; Maitra, S.; Stigers, K. D.;
Sun, Y. An unnatural amino acid that mimics a tripeptide -strand and
the use of the HBS approach to induce protein structure.
forms -sheetlike hydrogen-bonded dimers. J. Am. Chem. Soc. 2000,
Andrew M. Watkins received a B.A. in Chemistry from Harvard 122, 76547661.
University in 2011 and a Ph.D. in Chemical Biology from New York (10) Schafmeister, C. E.; Po, J.; Verdine, G. L. An all-hydrocarbon
University in 2016. He is currently a postdoctoral fellow at Stanford cross-linking system for enhancing the helicity and metabolic stability
University, where he is developing new algorithms for the structure of peptides. J. Am. Chem. Soc. 2000, 122, 58915892.
prediction and design of RNA. (11) Chapman, R. N.; Dimartino, G.; Arora, P. S. A highly stable
short -helix constrained by a main-chain hydrogen-bond surrogate. J.
Paramjit S. Arora obtained a B.S. in Chemistry from UC Berkeley Am. Chem. Soc. 2004, 126, 1225212253.
where he rst fell in love with organic chemistry while exploring (12) Fasan, R.; Dias, R. L. A.; Moehle, K.; Zerbe, O.; Vrijbloed, J. W.;
uorescent dyes in Richard Mathies group. He received a Ph.D. in Obrecht, D.; Robinson, J. A. Using a -hairpin to mimic an -helix:
Chemistry at UC Irvine under the mentorship of James Nowick, where Cyclic peptidomimetic inhibitors of the p53-HDM2 protein-protein
his interest in peptidomimetics was sparked by the exquisite fold interaction. Angew. Chem., Int. Ed. 2004, 43, 21092112.
adopted by urea scaolds. He learned the fundamentals of (13) Angelo, N. G.; Arora, P. S. Nonpeptidic foldamers from amino
biomolecular recognition as a postdoctoral fellow at Caltech under acids: synthesis and characterization of 1,3-substituted triazole
the tutelage of Peter Dervan before joining the faculty of New York oligomers. J. Am. Chem. Soc. 2005, 127, 1713417135.
University. His research eorts build upon principles of folding and (14) Phillips, S. T.; Piersanti, G.; Bartlett, P. A. Quantifying amino
recognition to modulate proteinprotein interactions. acid conformational preferences and side-chain-side-chain interactions
in -hairpins. Proc. Natl. Acad. Sci. U. S. A. 2005, 102, 1373713742.

ACKNOWLEDGMENTS
The work described here has been supported by the National
(15) Yin, H.; Lee, G. I.; Sedey, K. A.; Kutzki, O.; Park, H. S.; Orner,
B. P.; Ernst, J. T.; Wang, H. G.; Sebti, S. M.; Hamilton, A. D.
Terphenyl-based Bak BH3 -helical proteomimetics as low-molecular-
Institutes of Health (R01GM073943) and the National Science weight antagonists of Bcl-xL. J. Am. Chem. Soc. 2005, 127, 10191
Foundation (CHE1151554 and CHE0848410). N.S. is 10196.
(16) Goodman, C. M.; Choi, S.; Shandler, S.; DeGrado, W. F.
supported by a Ruth L. Kirschstein National Research Service
Foldamers as versatile frameworks for the design and evolution of
Award (NRSA F32GM120853) postdoctoral fellowship from function. Nat. Chem. Biol. 2007, 3, 25262.
NIGMS. A.M.W. has been supported by NYU Deans (17) Yoo, B.; Kirshenbaum, K. Peptoid architectures: elaboration,
Dissertation Fellowship and Stanford Deans Postdoctoral actuation, and application. Curr. Opin. Chem. Biol. 2008, 12, 71421.
Fellowship. (18) Horne, W. S.; Johnson, L. M.; Ketas, T. J.; Klasse, P. J.; Lu, M.;

DEDICATION
We dedicate this Account to Neville Kallenbach on the
Moore, J. P.; Gellman, S. H. Structural and biological mimicry of
protein surface recognition by alpha/beta-peptide foldamers. Proc.
Natl. Acad. Sci. U. S. A. 2009, 106, 147516.
(19) Ko, E.; Liu, J.; Burgess, K. Minimalist and universal
occasion of his 80th Birthday.

peptidomimetics. Chem. Soc. Rev. 2011, 40, 44114421.


(20) Azzarito, V.; Long, K.; Murphy, N. S.; Wilson, A. J. Inhibition of
ABBREVIATIONS alpha-helix-mediated protein-protein interactions using designed
HBS, hydrogen bond surrogate; OHM, oligooxopiperazine molecules. Nat. Chem. 2013, 5, 161173.
helix mimetic; PPI, proteinprotein interaction; PDM, protein (21) De Araujo, A. D.; Hoang, H. N.; Kok, W. M.; Diness, F.; Gupta,
P.; Hill, T. A.; Driver, R. W.; Price, D. A.; Liras, S.; Fairlie, D. P.
domain mimic; PDB, Protein Data Bank; SBS, salt bridge
Comparative -helicity of cyclic pentapeptides in water. Angew. Chem.
surrogate; SPPS, solid phase peptide synthesis

2014, 126, 70857089.


(22) Lao, B. B.; Drew, K.; Guarracino, D. A.; Brewer, T. F.; Heindel,
REFERENCES D. W.; Bonneau, R.; Arora, P. S. Rational design of topographical helix
(1) Milroy, L. G.; Grossmann, T. N.; Hennig, S.; Brunsveld, L.; mimics as potent inhibitors of protein-protein interactions. J. Am.
Ottmann, C. Modulators of protein-protein interactions. Chem. Rev. Chem. Soc. 2014, 136, 78777888.
2014, 114, 46954748. (23) Lingard, H.; Han, J. T.; Thompson, A. L.; Leung, I. K. H.; Scott,
(2) Modell, A. E.; Blosser, S. L.; Arora, P. S. Systematic targeting of R. T. W.; Thompson, S.; Hamilton, A. D. Diphenylacetylene-linked
protein-protein interactions. Trends Pharmacol. Sci. 2016, 37, 702 peptide strands induce bidirectional -sheet formation. Angew. Chem.,
713. Int. Ed. 2014, 53, 36503653.

1320 DOI: 10.1021/acs.accounts.7b00130


Acc. Chem. Res. 2017, 50, 13131322
Accounts of Chemical Research Article

(24) Henchey, L. K.; Jochim, A. L.; Arora, P. S. Contemporary (47) Wang, D.; Liao, W.; Arora, P. S. Enhanced metabolic stability
strategies for the stabilization of peptides in the alpha-helical and protein-binding properties of artificial helices derived from a
conformation. Curr. Opin. Chem. Biol. 2008, 12, 692697. hydrogen-bond surrogate: application to Bcl-xL. Angew. Chem., Int. Ed.
(25) Moon, H.; Lim, H.-S. Synthesis and screening of small-molecule 2005, 44, 65256529.
-helix mimetic libraries targeting proteinprotein interactions. Curr. (48) Miller, S. J.; Grubbs, R. H. Synthesis of conformationally
Opin. Chem. Biol. 2015, 24, 3847. restricted amino acids and peptides employing olefin metathesis. J. Am.
(26) Wuo, M. G.; Mahon, A. B.; Arora, P. S. An effective strategy for Chem. Soc. 1995, 117, 58555856.
stabilizing minimal coiled coil mimetics. J. Am. Chem. Soc. 2015, 137, (49) Henchey, L. K.; Kushal, S.; Dubey, R.; Chapman, R. N.;
1161811621. Olenyuk, B. Z.; Arora, P. S. Inhibition of hypoxia inducible factor 1-
(27) Checco, J. W.; Kreitler, D. F.; Thomas, N. C.; Belair, D. G.; transcription coactivator interaction by a hydrogen bond surrogate -
Rettko, N. J.; Murphy, W. L.; Forest, K. T.; Gellman, S. H. Targeting helix. J. Am. Chem. Soc. 2010, 132, 941943.
diverse proteinprotein interaction interfaces with /-peptides (50) Henchey, L. K.; Porter, J. R.; Ghosh, I.; Arora, P. S. High
derived from the Z-domain scaffold. Proc. Natl. Acad. Sci. U. S. A. specificity in protein recognition by hydrogen-bond-surrogate -
2015, 112, 45524557. helices: selective inhibition of the p53/MDM2 complex. ChemBio-
(28) Pelay-Gimeno, M.; Glas, A.; Koch, O.; Grossmann, T. N.
Chem 2010, 11, 21042107.
Structure-Based Design of Inhibitors of ProteinProtein Interactions: (51) Kushal, S.; Lao, B. B.; Henchey, L. K.; Dubey, R.; Mesallati, H.;
Mimicking Peptide Binding Epitopes. Angew. Chem., Int. Ed. 2015, 54,
Traaseth, N. J.; Olenyuk, B. Z.; Arora, P. S. Protein domain mimetics
88968927.
as in vivo modulators of hypoxia-inducible factor signaling. Proc. Natl.
(29) Jochim, A. L.; Arora, P. S. Systematic analysis of helical protein
interfaces reveals targets for synthetic inhibitors. ACS Chem. Biol. Acad. Sci. U. S. A. 2013, 110, 156027.
2010, 5, 919923. (52) Douse, C. H.; Maas, S. J.; Thomas, J. C.; Garnett, J. A.; Sun, Y.;
(30) Bullock, B. N.; Jochim, A. L.; Arora, P. S. Assessing helical Cota, E.; Tate, E. W. Crystal structures of stapled and hydrogen bond
protein interfaces for inhibitor design. J. Am. Chem. Soc. 2011, 133, surrogate peptides targeting a fully buried protein-helix interaction.
1422014223. ACS Chem. Biol. 2014, 9, 22042209.
(31) Bergey, C. M.; Watkins, A. M.; Arora, P. S. HippDB: A database (53) Jiang, B. H.; Zheng, J. Z.; Leung, S. W.; Roe, R.; Semenza, G. L.
of readily targeted helical protein-protein interactions. Bioinformatics Transactivation and inhibitory domains of hypoxia-inducible factor 1.
2013, 29, 28062807. J. Biol. Chem. 1997, 272, 1925319260.
(32) Lao, B. B.; Grishagin, I.; Mesallati, H.; Brewer, T. F.; Olenyuk, (54) Pugh, C. W.; ORourke, J. F.; Nagao, M.; Gleadle, J. M.;
B. Z.; Arora, P. S. In vivo modulation of hypoxia-inducible signaling by Ratcliffe, P. J. Activation of hypoxia-inducible factor-1; definition of
topographical helix mimetics. Proc. Natl. Acad. Sci. U. S. A. 2014, 111, regulatory domains within the subunit. J. Biol. Chem. 1997, 272,
75317536. 1120511214.
(33) Tosovska, P.; Arora, P. S. Oligooxopiperazines as nonpeptidic - (55) Dames, S. A.; Martinez-Yamout, M.; De Guzman, R. N.; Dyson,
helix mimetics. Org. Lett. 2010, 12, 15881591. H. J.; Wright, P. E. Structural basis for Hif-1/CBP recognition in the
(34) Lao, B. B.; Arora, P. S. Oligooxopiperazines as topographical cellular hypoxic response. Proc. Natl. Acad. Sci. U. S. A. 2002, 99,
helix mimetics. Top. Heterocycl. Chem. 2016, 49, 124. 52715276.
(35) Kan, T.; Fukuyama, T. Ns strategies: a highly versatile synthetic (56) Freedman, S. J.; Sun, Z.-Y. Y. J.; Poy, F.; Kung, A. L.; Livingston,
method for amines. Chem. Commun. 2004, 353359. D. M.; Wagner, G.; Eck, M. J. Structural basis for recruitment of CBP/
(36) Houk, K. N.; Leach, A. G.; Kim, S. P.; Zhang, X. Binding p300 by hypoxia-inducible factor-1. Proc. Natl. Acad. Sci. U. S. A.
affinities of host-guest, protein-ligand, and protein-transition-state 2002, 99, 53675372.
complexes. Angew. Chem., Int. Ed. 2003, 42, 48724897. (57) Liu, W.; Shen, S. M.; Zhao, X. Y.; Chen, G. Q. Targeted genes
(37) Chapman, R. N.; Arora, P. S. Optimized synthesis of hydrogen- and interacting proteins of hypoxia inducible factor-1. Int. J. Biochem.
bond surrogate helices: surprising effects of microwave heating on the Mol. Biol. 2012, 3, 165178.
activity of Grubbs catalysts. Org. Lett. 2006, 8, 58255828. (58) Hanahan, D.; Weinberg, R. A. Hallmarks of cancer: The next
(38) Liu, J.; Wang, D.; Zheng, Q.; Lu, M.; Arora, P. S. Atomic generation. Cell 2011, 144, 646674.
structure of a short -helix stabilized by a main chain hydrogen-bond (59) Wright, P. E.; Dyson, H. J. Intrinsically disordered proteins in
surrogate. J. Am. Chem. Soc. 2008, 130, 43344337. cellular signalling and regulation. Nat. Rev. Mol. Cell Biol. 2014, 16,
(39) Patgiri, A.; Witten, M. R.; Arora, P. S. Solid phase synthesis of 1829.
hydrogen bond surrogate derived -helices: resolving the case of a (60) Patel, D.; Huang, S. M.; Baglia, L. A.; McCance, D. J. The E6
difficult amide coupling. Org. Biomol. Chem. 2010, 8, 17731776. protein of human papillomavirus type 16 binds to and inhibits co-
(40) Patgiri, A.; Menzenski, M. Z.; Mahon, A. B.; Arora, P. S. Solid-
activation by CBP and p300. EMBO J. 1999, 18, 50615072.
phase synthesis of short -helices stabilized by the hydrogen bond
(61) Xie, X.; Piao, L.; Bullock, B. N.; Smith, A.; Su, T.; Zhang, M.;
surrogate approach. Nat. Protoc. 2010, 5, 18571865.
Teknos, T. N.; Arora, P. S.; Pan, Q. Targeting HPV16 E6-p300
(41) Mahon, A. B.; Arora, P. S. Design, synthesis and protein-
interaction reactivates p53 and inhibits the tumorigenicity of HPV-
targeting properties of thioether-linked hydrogen bond surrogate
helices. Chem. Commun. 2012, 48, 14161418. positive head and neck squamous cell carcinoma. Oncogene 2014, 33,
(42) Miller, S. E.; Kallenbach, N. R.; Arora, P. S. Reversible -helix 10371046.
formation controlled by a hydrogen bond surrogate. Tetrahedron 2012, (62) Watkins, A. M.; Arora, P. S. Anatomy of strands at protein
68, 44344437. protein interfaces. ACS Chem. Biol. 2014, 9, 17471754.
(43) Joy, S. T.; Arora, P. S. An optimal hydrogen-bond surrogate for (63) Tyndall, J. D. A.; Nall, T.; Fairlie, D. P. Proteases universally
-helices. Chem. Commun. 2016, 52, 57385741. recognize beta strands in their active sites. Chem. Rev. 2005, 105, 973
(44) Patgiri, A.; Jochim, A. L.; Arora, P. S. A hydrogen bond 1000.
surrogate approach for stabilization of short peptide sequences in (64) Loughlin, W. A.; Tyndall, J. D.; Glenn, M. P.; Hill, T. A.; Fairlie,
alpha-helical conformation. Acc. Chem. Res. 2008, 41, 1289300. D. P. Update 1 of: Beta-strand mimetics. Chem. Rev. 2010, 110,
(45) Cabezas, E.; Satterthwait, A. C. The hydrogen bond mimic PR3269.
approach: Solid-phase synthesis of a peptide stabilized as an alpha- (65) Smith, A. B.; Guzman, M. C.; Sprengeler, P. A.; Keenan, T. P.;
helix with a hydrazone link. J. Am. Chem. Soc. 1999, 121, 38623875. Holcomb, R. C.; Wood, J. L.; Carroll, P. J.; Hirschmann, R. De-Novo
(46) Zimm, B. H.; Bragg, J. K. Theory of the phase transition Design, Synthesis, and X-Ray Crystal-Structures of Pyrrolinone-Based
between helix and random coil in polypeptide chains. J. Chem. Phys. Beta-Strand Peptidomimetics. J. Am. Chem. Soc. 1994, 116, 9947
1959, 31, 526535. 9962.

1321 DOI: 10.1021/acs.accounts.7b00130


Acc. Chem. Res. 2017, 50, 13131322
Accounts of Chemical Research Article

(66) Hammond, M. C.; Harris, B. Z.; Lim, W. A.; Bartlett, P. A. Beta


strand peptidomimetics as potent PDZ domain ligands. Chem. Biol.
2006, 13, 12471251.
(67) Jochim, A. L.; Miller, S. E.; Angelo, N. G.; Arora, P. S.
Evaluation of triazolamers as active site inhibitors of HIV-1 protease.
Bioorg. Med. Chem. Lett. 2009, 19, 60236026.
(68) Kang, C. W.; Sarnowski, M. P.; Ranatunga, S.; Wojtas, L.;
Metcalf, R. S.; Guida, W. C.; Del Valle, J. R. beta-Strand mimics based
on tetrahydropyridazinedione (tpd) peptide stitching. Chem. Commun.
2015, 51, 1625962.
(69) Haque, T. S.; Gellman, S. H. Insights on beta-hairpin stability in
aqueous solution from peptides with enforced type I and type II beta-
turns. J. Am. Chem. Soc. 1997, 119, 23032304.
(70) Craik, D. J.; Fairlie, D. P.; Liras, S.; Price, D. The Future of
Peptide-based Drugs. Chem. Biol. Drug Des. 2013, 81, 136147.
(71) Angelo, N. G.; Arora, P. S. Solution- and solid-phase synthesis of
triazole oligomers that display protein-like functionality. J. Org. Chem.
2007, 72, 79637967.
(72) Angell, Y. L.; Burgess, K. Peptidomimetics via copper-catalyzed
azidealkyne cycloadditions. Chem. Soc. Rev. 2007, 36, 16741689.
(73) Sola, J.; Bolte, M.; Alfonso, I. Conformational promiscuity in
triazolamers derived from quaternary amino acids mimics peptide
behaviour. Org. Biomol. Chem. 2015, 13, 10797801.
(74) Wlodawer, A.; Erickson, J. W. Structure-based inhibitors of HIV-
1 protease. Annu. Rev. Biochem. 1993, 62, 543585.
(75) Matayoshi, E. D.; Wang, G. T.; Krafft, G. A.; Erickson, J. Novel
Fluorogenic Substrates for Assaying Retroviral Proteases by Resonance
Energy Transfer. Science 1990, 247, 954958.
(76) Crick, F. H. C. The packing of -helices: simple coiled-coils.
Acta Crystallogr. 1953, 6, 689697.
(77) Burkhard, P.; Stetefeld, J.; Strelkov, S. V. Coiled coils: a highly
versatile protein folding motif. Trends Cell Biol. 2001, 11, 8288.
(78) Woolfson, D. N. The design of coiled-coil structures and
assemblies. Adv. Protein Chem. 2005, 70, 79112.
(79) Watkins, A. M.; Wuo, M. G.; Arora, P. S. Protein-protein
interactions mediated by helical tertiary structure motifs. J. Am. Chem.
Soc. 2015, 137, 1162211630.
(80) Su, J. Y.; Hodges, R. S.; Kay, C. M. Effect of chain length on the
formation and stability of synthetic -helical coiled coils. Biochemistry
1994, 33, 1550110.
(81) Burkhard, P.; Meier, M.; Lustig, A. Design of a minimal protein
oligomerization domain by a structural approach. Protein Sci. 2000, 9,
22942301.
(82) Torres, O.; Yuksel, D.; Bernardina, M.; Kumar, K.; Bong, D.
Peptide tertiary structure nucleation by side-chain crosslinking with
metal complexation and double Click cycloaddition. ChemBioChem
2008, 9, 17011705.
(83) Sun, X.-J.; Wang, Z.; Wang, L.; Jiang, Y.; Kost, N.; Soong, T. D.;
Chen, W.-Y.; Tang, Z.; Nakadai, T.; Elemento, O.; Fischle, W.;
Melnick, A.; Patel, D. J.; Nimer, S. D.; Roeder, R. G. A stable
transcription factor complex nucleated by oligomeric AML1-ETO
controls leukaemogenesis. Nature 2013, 500, 937.

1322 DOI: 10.1021/acs.accounts.7b00130


Acc. Chem. Res. 2017, 50, 13131322

You might also like