You are on page 1of 16

Bioorganic & Medicinal Chemistry Letters 29 (2019) 2535–2550

Contents lists available at ScienceDirect

Bioorganic & Medicinal Chemistry Letters


journal homepage: www.elsevier.com/locate/bmcl

Digest

Applications of amide isosteres in medicinal chemistry T


a,⁎ a b
Shaoyi Sun , Qi Jia , Zaihui Zhang
a
Xenon Pharmaceuticals Inc., 200-3650 Gilmore Way, Burnaby, BC V5G 4W8, Canada
b
Signalchem Lifesciences Corp., 110-13210, Vanier Place, Richmond, BC V6V 2J2, Canada

ARTICLE INFO ABSTRACT

Keywords: Isosteric replacement of amide groups is a classic practice in medicinal chemistry. This digest highlights the
Amide isostere applications of most commonly employed amide isosteres in drug design aiming at improving potency and
Transition-state mimic selectivity, optimizing physicochemical and pharmacokinetic properties, eliminating or modifying toxicophores,
Heterocycle as well as providing novel intellectual property of lead compounds.
Trifluoroethylamine
Fluoroalkene
3-Aminooxetane
Boronic acid

Introduction the use of TS mimics as non-hydrolysable amide isosteres to replace a


specific scissile peptide bond is a well-known approach to overcome
Peptides and nonpeptidic small-molecule amides represent an im- one of the major drawbacks in the use of peptides as medical agents,
portant class of biologically active molecules. The potential use of these namely, their rapid degradation by peptidases.3 A TS mimic is defined
molecules as oral drugs is often compromised by their poor physico- as a functional group that can mimic the tetrahedral TS of amide bond
chemical (PC) and pharmacokinetic (PK) properties, largely due to the hydrolysis, but cannot itself be hydrolyzed by the enzyme. Some TS
metabolic instability of the amide bond toward enzymatic degradation mimics that have been successfully used in the design of protease in-
and/or the high polarity. Therefore, isosteric replacement of the amide hibitors are depicted in Fig. 1.
group has been widely adopted as a fundamental tactical approach in Saquinavir (1, Fig. 2) is the first TS peptidomimetic to be approved
medicinal chemistry for the design of new drugs.1 The success of this by the FDA for the treatment of AIDS.4 The design strategy applied by
strategy in identifying new biologically active molecules in distinct the Roche team was to focus on the HIV pol substrate fragment Leu165-
therapeutic areas has gained a significant growth. This digest highlights Ile169, which contains the scissile bond Phe167-Pro168. Optimization
most commonly employed amide isosteres in drug design and in- of this peptide was achieved by applying the hydroxyethylamine (I,
corporates sufficient details in addressing a number of aspects relative Fig. 1) TS mimic.
to the amide groups, including the importance in improving potency Following the discovery of saquinavir, seven other analogs have
and selectivity, optimizing PC and PK properties, eliminating or mod- been successfully marketed for medical use including nelfinavir (2),5
ifying toxicophores, as well as providing novel intellectual property of amprenavir (3),6 indinavir (4),7 ritonavir (5),8 lopinavir (6),9 darunavir
lead compounds. Representative examples selected in this digest are (7),10 and atazanavir (8),11 which exhibited better potency and/or PK
classified into six categories: transition-state (TS) mimics, heterocycles, properties than saquinavir. Atazanavir (8, Fig. 2) contains a hydro-
trifluoroethylamine, fluoroalkene, 3-aminooxetane, and boronic acid. xyethyl hydrazine fragment (VI, Fig. 1). The aza modification elimi-
By summarizing these examples, we hope to provide medicinal che- nated one of the chiral centers in the molecule, thereby reduced the
mists an overview of the usefulness of these amide isosteres in drug challenges associated with the synthesis of this compound.
design. Aliskiren (11, Fig. 3) represents a unique class of peptidomimetic
renin inhibitors, developed by Novartis for treatment of hypertension.
Transition-state mimics as peptide bond isosteres The design of aliskiren was focused on the modification of the large
peptide CG29287 (9) derived from the N-terminal sequence of angio-
Chemical modification of peptides to improve their biological ac- tensinogen by incorporation of the statine TS mimic (V, Fig. 1) at the
tivity and PC-PK properties has been extensively studied.2 In particular, renin cleavage site. Subsequent reduction of molecular size by


Corresponding author.
E-mail address: ssun@xenon-pharma.com (S. Sun).

https://doi.org/10.1016/j.bmcl.2019.07.033
Received 26 April 2019; Received in revised form 17 July 2019; Accepted 19 July 2019
Available online 22 July 2019
0960-894X/ © 2019 Elsevier Ltd. All rights reserved.
S. Sun, et al. Bioorganic & Medicinal Chemistry Letters 29 (2019) 2535–2550

Figure 1. Selected transition-state (TS) mimics.

truncation at both the N- and the C-terminus and incorporation of a Diazepam (16, Fig. 5), an agent increases the activity of the GABA
hydroxyethylene isostere (II, Fig. 1) led to CGP38560 (10), and further neurotransmitter in the brain, was marketed by Hoffmann-La Roche for
structure refinement afforded the drug molecule 11.12 treatment of a number of CNS disorders such as anxiety, epilepsy,
TS mimics (Fig. 1) have also been utilized in the design of less muscle spasms and alcohol withdrawal syndrome.14
peptide-like β-secretase (BACE-1) inhibitors.13 This approach produced In humans, demethylation of diazepam is effected by CYP450 en-
many potent inhibitors containing statine (e.g., compound 12),13b hy- zymes (CYP 2C9, 2C19, 3A4, 3A5, and 2B6), producing the active
droxyethylene (e.g., compound 13),13c reduced amide (e.g., compound metabolite desmethyldiazepam (structure not shown),15 which is
14),13m and hydroxyethylamine (e.g., compound 15)13n cores (Fig. 4). equipotent to diazepam.16 Due to its longer half-life (50–120 h), des-
Compound 15 is the latest brain penetrant molecule in this class re- methyldiazepam accumulates in humans as a major active component
ported by Bueno et al. with a brain to plasma ratio of 0.38 in mice. during administration.17
Unlike HIV protease and renin, BACE1 is a CNS located aspartic pro- An extensive investigation of amide isosteres derived from dia-
tease, and despite extensive efforts that have been made, none of these zepam at Upjohn resulted in the discovery of alprazolam (17), a ben-
compounds was selected for clinical development, largely due to the zodiazepine with a 1,2,4-triazole ring fused to its structure. Alprazolam
challenges to achieve optimal CNS penetration. However, information is a highly potent, short-acting benzodiazepine, which was approved as
gathered in these studies can be applied to the optimization of mole- an anxiolytic drug in 1983.18
cules of peripheral targets. Midazolam (18) is another diazepam analog, in which the amide
was modified into a imidazole ring, and is an agent used for anesthesia,
seizure suppression, procedural sedation, trouble sleeping, and severe
Heterocycles as amide bond surrogates
agitation.19
In an effort to develop more potent calcitonin gene-related peptide
Replacement of an amide group with a heterocyclic ring in a
(CGRP) receptor antagonists for treatment of migraine, Merck re-
bioactive molecule can have many important effects on physicochem-
searchers explored fused heterocyclic derivatives of their clinical can-
ical and pharmacological properties.1 This substitution introduces
didate telcagepant (19, Fig. 6). The imidazoazepane scaffold (as shown
structural rigidity, which may lead to compounds with improved po-
in structure 20) was identified as a suitable replacement for the lactam.
tency, selectivity, metabolic stability, and PK properties. Commonly
This investigation led to the identification of MK-2918 (21) as a novel
used heterocycles include triazoles, imidazoles, oxadiazoles, oxazoles,
and highly potent CGRP receptor antagonist. Compound 21 contains an
isoxazoles, imidazolidinones, triazolones and many six-membered het-
azabenzoxazinone spiropiperidine fragment, which improved the PK
erocycles, which retain the geometry of the amide bond or maintain the
properties in rhesus monkeys. The tertiary methyl ether element at-
hydrogen bond (H-bond) accepting/donating properties of the amide
tached on the imidazole ring greatly enhanced the potency and
group, and yield many successful examples.

2536
S. Sun, et al. Bioorganic & Medicinal Chemistry Letters 29 (2019) 2535–2550

Figure 2. Structures of peptide-derived TS inhibitors of HIV protease.

diminished the hERG activity.20 stability and brain penetration, the benzamide functionality was re-
In a similar approach, Elliott et al. replaced the lactam present in placed with an 1,2,4-oxadiazole ring and the isobutyl side chain was
compound 22 (Fig. 7) with heterocycles leading to a series of novel modified to a trifluoropropane group. These modifications identified
human neurokinin-1 receptor (hNK1R) antagonists. Both triazolone BMS-708163 (26), a potent, selective, metabolically stable, and orally
(23) and imidazole (24) displayed similar potency to that of 22. bioavailable γ-secretase inhibitor with IC50 values of 0.3 nM and
However, these two compounds showed different profiles; triazolone 23 0.27 nM in reduction of Aβ40 and Aβ42. BMS-708163 demonstrated
demonstrated good activities in a gerbil in vivo model and low hIkr af- favorable PK properties in both rats and dogs, and the brain to plasma
finity, while imidazole 24 was inactive in this in vivo model and showed ratio in dogs reached 2.4. BMS-708163 has been evaluated in clinical
insufficient selectivity over hIkr.21 trials as a potential treatment of Alzheimer's disease.23
1,2,4-Oxadiazole and 1,3,4-oxadiazole heterocyclic systems have A recent example where an amide bond was replaced with ox-
both planarity and dipole moment similar to the amide functionality, adizaoles was reported in the optimization of a subtype selective me-
however, they lack the H-bond donating capacity. Replacement of an tabotropic glutamate receptor subtype 7 (mGlu7) negative allosteric
amide group with an oxadiazole can lead to improvements in metabolic modulator (NAM) 27 (Fig. 9).24a While compound 27 displayed good in
stability, membrane permeability and CNS penetration 22 vitro activity, its poor metabolic stability (rat CLp = 64.2 mL/min/kg)
An example worthy of note is given by the γ-secretase inhibitor 25 limited the in vivo evaluation. Reed et al. found that both 1,2,4- and
(Fig. 8), a compound poorly stable in both rat and human microsomal 1,3,4-oxadizaole rings are effective bioisosteres of the amide bond in
incubation. Oral administration of 25 in rats revealed that the brain to 27. In combination with a replacement of the triazole substitution with
plasma ratio was very low (0.01). In an effort to improve metabolic an imidazole ring, this modification led to the most potent mGlu7 NAM

2537
S. Sun, et al. Bioorganic & Medicinal Chemistry Letters 29 (2019) 2535–2550

Figure 3. Structures of renin inhibitors 9, 10 and aliskren (11).

28 (VU6019278) in this series, which showed low predicted hepatic


clearance (rat CLhep = 27.7 mL/min/kg) and high CNS penetration (rat
Kp = 4.9, Kp,uu = 0.65).24b
The discovery of the human immunodeficiency virus type-1 (HIV-1)
integrase inhibitor raltegravir (36) illustrated the role of 1,3,4-ox-
adiazole as an effective amide isostere in modulating potency shift in
cell based antiviral assays in a medium containing 10% heat-inactivated
fetal bovine serum (FBS) or 50% normal human serum (NHS).25 As
summarized in Fig. 10, 2-pyridine carboxamide 30 was 2-fold less ac-
tive on the enzyme than 29, with a 6-fold shift in cell based assay in the
presence of 10% FBS, but a 50-fold shift in the presence of 50% NHS.
Figure 5. Structures of diazepam (16), alprazolam (17) and midazolam (18).
Addition of more nitrogen in the ring reduced the potency shift, e.g.,
pyrimidine derivative 31 was potent in all assays, while only a marginal
improvement was observed with pyridazine derivative 32. The five-

Figure 4. Structures of BACE-1 inhibitors 12–15.

2538
S. Sun, et al. Bioorganic & Medicinal Chemistry Letters 29 (2019) 2535–2550

Figure 6. Structures of telcagepant (19), compound 20 and MK-2918 (21).

membered analogs oxazole 33, imidazole 34, thiazole 35 and triazole


37 showed single-digit nanomolar activity on integrase, however, a
large potency decrease was observed in cellular assays in both 10% FBS
and 50% NHS conditions for all four analogs. Oxadiazole 36 was the
most potent compound with minimal potency shift in cell based anti-
viral assays. The good potency, selectivity, mutants, PK and safety
profiles of 36 contributed to its success in clinical development as the
first HIV-integrase inhibitor approved for the treatment of HIV-1 in-
fection in the United States in 2007.
In the case of γ-secretase inhibitor 38, replacement of the amide
bond with oxadiazoles resulted in a significant improvement in CYP3A4
liability. As shown in Fig. 11, oxadiazoles 39 and 40 maintained similar
γ-secretase inhibition to that of amide 38, along with reductions in Figure 8. Structures of compounds 25 and BMS-708163 (26).
CYP3A4 activity, suggesting that both oxadiazoles were proper amide
isosteres in this particular case. In contrast, a significant increase in analogs, where the amide portion was replaced with a series of five-
CYP3A4 inhibition was observed with oxazole 41.26 membered heterocycles. As shown in Fig. 12, oxadiazole analog 43 and
Initial SAR exploration around phosphoinositide 3-kinase δ (PI3Kδ) oxazole analog 44 showed comparable potency for PI3Kδ to amide 42,
inhibitor 42 and homology docking suggested that an internal H-bond suggesting that the H-bond donor of the amide was not necessary for
in the molecule maintains a planar binding conformation, facilitating the potency. In addition, furan analog 46 was almost equipotent to
an effective interaction of the key Trp760 residue in the biding pocket. compounds 42, 43 and 44, suggesting that the H-bond acceptor was
Building on this hypothesis, Down et al. designed a class of small size also not necessary for the potency, since furan is a weak H-bond

Figure 7. Structures of hNK1R antagonists 22–24.

2539
S. Sun, et al. Bioorganic & Medicinal Chemistry Letters 29 (2019) 2535–2550

Figure 9. Structures of mGlu7 NAM 27 and VU6019278 (28).

acceptor27. In contrast, thiophene analog 47, predicted to have a non- identification of two series of novel NaV1.7 blockers as exemplified by
planar preferred conformation, was about 10-fold less potent relative to isoxazoline 4929b and isoxazole 50.29c Both compounds displayed im-
the progenitors, supporting the hypothesis that the planarity of the proved PK properties, and in vivo efficacy in rodent pain models similar
amide linker was essential for the potency. Further refinements of ox- to CDA54 (see Fig 13).
adiazole analog 42 and oxazole analog 44 by addition of extended In the optimization of ataxia telangiectasia mutated and rad3-re-
amine groups on the oxadiazole and oxazole rings identified clinical lated (ATR) kinase inhibitor VE-821 (51), initial efforts were focused on
candidate GSK2269557 (45) for the treatment of respiratory indications replacing the anilide group with a variety of fused heterocycles such as
via inhalation.28 benzimidazole, benzoxazole, benzothiazole, and indole to address the
CDA54 (48) was reported to be a potent and state-dependent so- potential liability associated with the aniline buried in the molecule.
dium channel (NaV1.7) blocker. Evaluation of 48 in human liver mi- Although the ATR potency was maintained, these structural changes
crosomes (HLM) showed rapid cleavage of the N-Me amide side chain, resulted in significant decreases in selectivity for ATR against ataxia
leading to dealkylated metabolites.29a Key SAR studies were executed telangiectasia mutated (ATM) kinase.30a Homology modeling analyses
toward identification of analogs with improved PK properties by re- suggested that phenyl substituted five-membered heterocycles could
placing this amide group with heterocycles. These efforts led to the closely mimic the shape of the anilide and fit in the binding pocket

Figure 10. Selected SAR toward raltegravir (36).

2540
S. Sun, et al. Bioorganic & Medicinal Chemistry Letters 29 (2019) 2535–2550

Figure 11. Selected SAR of γ-secretase inhibitor 38


and its heterocyclic analogs.

Figure 12. Key SAR toward PI3Kδ inhibitor GSK2269557 (45).

Figure 13. Structures of NaV1.7 inhibitors 48–50.

properly without causing steric clashes with the tyrosine gatekeeper Subsequently, the more potent isopropylsulfone 52 was taken as a
residue of the phosphoinositol 3-kinase-like kinase (PIKK) family, po- starting point, and a matched pair of 53 and 54 was evaluated. The
tentially retaining the potency and selectivity for ATR over ATM. partially saturated 4,5-dihydroisoxazole analogue 53 was found to be

2541
S. Sun, et al. Bioorganic & Medicinal Chemistry Letters 29 (2019) 2535–2550

Figure 14. Key SAR toward ATR inhibitor VX-970 (55).

much less active than the flat unsaturated isoxazole 54, suggesting that atoms are better H-bond acceptors than aromatic oxygens. Because of
a planar heterocyclic linker is necessary for optimal binding with the its good potency and superior selectivity, isoxazole 54 was selected for
enzyme. As summarized by the data in Fig. 14, aromatic five-membered exploration of a highly negatively charged area of the ATP biding site
heterocycles are generally well tolerated as an amide replacement with and ultimately led to the discovery of VX-970 (55), the first ATR in-
many compounds showing similar affinity on ATR to that of amide 52, hibitor evaluated in humans.30b
and the selectivity for ATR over the homologous kinases ATM and DNA- 1,2,3-triazole is among the most common amide bond isosteres; its
dependent protein kinase (DNA-PK) was maintained. structural features allow a good overlap with an amide, and it has better
As predicted by the negative relative conformational energies, H-bond accepting and H-bond donating capacity than an amide, how-
compounds 54, 56, 57 and 59 favor the bioactive conformation, ever, it possesses strong dipole moment.31 The 1,4-disubstituted 1,2,3-
showing low nanomolar potency against ATR. 1,2,4-oxadiazole 58 fa- triazole analog 60 was examined in this study and found to be less
vors the alternative conformation, and was the least active compound active in cell-based assay.30b Examples where replacement of amide
among this set of analogs. The preference for the alternative con- bonds with 1,2,3-triazoles had detrimental effects on potency have also
formation of 58 was predicted by the positive relative conformational been reported by Doiron et al. in a series of cystic fibrosis transmem-
energy, and is in agreement with the observation that aromatic nitrogen brane conductance regulator (CFTR) modulators.32

2542
S. Sun, et al. Bioorganic & Medicinal Chemistry Letters 29 (2019) 2535–2550

Figure 15. Structures of PPARδ modulators 61 and 62.

Figure 16. Structures of SCD1 inhibitors 63–66.

Figure 17. Structures of GSM modulators 67–69.

An X-ray structure of the PPARδ modulator 61 (Fig. 15) bound to compounds displayed significantly better PPARδ activity and se-
the ligand binding domain revealed that the amide portion exists in the lectivity, e.g., the corresponding imidazole analog of 61 (structure not
thermodynamically disfavored cis-amide orientation.33a This observa- shown) had an EC50 of 1 nM for PPARδ with greater than 1200-fold
tion promoted the authors to replace the amide bond with five-mem- selectivity over PPARα. The further optimized compound 62 was a
bered heterocycles to secure the bioactive conformation. Among a potent and selective PPARδ modulator with good PK properties. Analog
number of heterocyclic derivatives evaluated, the imidazole series of 62 altered the expression of PPARδ target genes and improved fatty

2543
S. Sun, et al. Bioorganic & Medicinal Chemistry Letters 29 (2019) 2535–2550

inhibitors, exemplified by 6434a and 65,34b which demonstrated sig-


nificant efficacy in reducing the plasma C16:1/C16:0 triglycerides de-
saturation index in rats. Furthermore, the amide portion at C5-position
in compound 64 was replaced with a 1,2,4-triazole moiety. However,
compound 66 was found to be about 16-fold less active than 64
(Fig. 16).34c
E-2012 (67, Fig. 17) represents a new class of γ-secretase mod-
ulators (GSMs) with impressive potency in reduction of Aβ42 in cells.
The clinical development of 67 was suspended due to the finding of rat
lenticular toxicity. Sun et al. modified the structure of 67, and identified
oxadiazine as an effective replacement of the lactam in 67. The es-
sential design was built on an analysis of related patent literature that
the middle part of the molecule can be highly variable, and a bulky
group was well tolerated at the benzylic position (as shown in 68),
which inspired the team to build a ring system from the position of the
carbonyl oxygen to the benzylic position. An imino group would be able
to form such a ring while still maintaining the H-bond accepting fea-
tures of the carbonyl oxygen. However, it is too basic, thus may lead to
Pgp efflux and hERG liabilities. Therefore, a series of less basic hy-
droxyamidines, including oxadiazolines and oxadiazines, were pre-
pared. Oxadiazine 69 was one of the most potent GSM, which de-
monstrated significant efficacy in reduction of Aβ42 in rats (3 mg/kg,
orally) and cynomolgous monkeys (30 mg/kg, orally). The corre-
sponding plasma exposure of 69 in monkeys was ∼6 μM (4 h post‐-
dosing), and the average brain/plasma ratio was determined to be
1.4.35
In the optimization of a series of hepatitis C virus (HCV) NS5A in-
hibitors with dual GT-1a/-1b inhibitory activity at Bristol-Myers
Squibb, the potential genotoxic issue associated with the aniline present
in compound 70 was addressed by fusing the amide on the phenyl to
form a benzimidazole ring that mimics the H-bond donating and ac-
cepting properties of the amide moiety. Subsequent modification of the
central core of 71 by removal of the alkyne linker and extension of the
benzimidazole ring to a phenyl-imidazolyl fragment effectively restored
Figure 18. Key SAR toward HCV NS5A inhibitor daclatasvir (73). the GT-1a inhibitory potency and finally, replacement of the R-phe-
nylglycine fragment in 72 with L-valine along with a chiral inversion
acid oxidation when tested in mice and in patient-derived muscle improved PK properties and led to the discovery of daclatasvir (73,
myoblasts. These studies support the hypothesis that selective PPARδ Fig. 18)36 Daclatasvir is the first HCV NS5A replication complex in-
modulators may offer benefits as a therapeutic target for Duchenne hibitor marketed to treat HCV patients.
muscular dystrophy.33b Researchers at Merck have identified amide 74 (Fig. 19) as a potent
In the quest for novel and potent stearoyl-CoA desaturase-1 (SCD1) transient receptor potential vanilloid 1 (TRPV1) antagonist. Evaluation
inhibitors, researchers at Xenon and Novartis discovered that replace- of 74 in rats (5 mpk) revealed poor oral absorption with low hepatic
ment of the amide moiety at the C2-position in compound 63 with portal vein (HPV) and systemic exposures (AUChpv = 0.08 μM h,
heterocycles resulted in substantial improvements in potency and AUCsys = 0.04 μM h, respectively). To improve the PK properties of 74,
physicochemical properties. Both imidazolidinone and triazolone rings a number of fused heterocyclic derivatives were examined and yielded
revealed to be ideal replacements for the amide group. This modifica- benzimidazoles and indazolones as promising carboxamide isosteres,
tion provided a series of novel, potent, and metabolically stable SCD1 which retained good in vitro potency on hTRPV1. While benzimidazole

Figure 19. Structures of TRPV1 inhibitors 74–76.

2544
S. Sun, et al. Bioorganic & Medicinal Chemistry Letters 29 (2019) 2535–2550

Figure 20. Structures of TYK2 inhibitors


77–79.

Figure 21. Structures of compounds 80, 81 and BMS-770767 (82).

Figure 22. Structures of p38α MAPK inhibitors 83–85.

Figure 23. Structures of KV1.3 inhibitor 86 and KV1.5 inhibitors 87 and 88.

75 showed a modest improvement in PK (AUChpv = 1.48 μM h, inhibitors, Liang et al. investigated structurally restricted analogs de-
AUCsys = 1.25 μM h, respectively), indazolone 76 demonstrated a sig- rived from lead 77 (Fig. 20), in which the rotatable benzamide bond
nificantly enhanced PK profile in rats (AUChpv = 6.2 μM h, was cyclized on the pyridine ring to form fused heterocycles. This
AUCsys = 3.0 μM h, respectively). Compound 76 was also found to have modification led to several active scaffolds including imidazopyridine,
low plasma clearance of 8 mL/min/kg with a half-life of 3.6 h (iv, oxazolopyridine, thiazolopyridine, and pyrazolopyridine. Imidazopyr-
1 mpk).37 idine 78 had an IC50 value of 1.7 nM with much better selectivity over
In an effort to discover potent and selective tyrosine kinase 2 (TYK2) Janus kinase 2 (JAK2). Optimization of 78 was carried out by replacing

2545
S. Sun, et al. Bioorganic & Medicinal Chemistry Letters 29 (2019) 2535–2550

Figure 24. Structures of CHK1 inhibitor 89–91.

Figure 25. Structures of HLE inhibitors 92–94.

JAK2, and demonstrated efficacy in an imiquimod-induced psoriasis


model in mice.38
Compound 80 (Fig. 21) was one of the highly potent and selective
inhibitors of human 11β-hydroxysteroid dehydrogenase type 1 (11β-
HSD-1) identified by scientists at Bristol-Myers Squibb.39a To discover
structurally distinguished 11β-HSD-1 inhibitors, the team investigated
a number of heterocycles as potential H-bond accepting surrogates to
Figure 26. Structures of PLG 95 and PLG-mimetic 96. mimic the amide portion in 80, and identified 1,2,4-triazolopyridine
(shown in compound 81) as an effective replacement of the picolina-
mide core. Optimization of 81 was focused on improving metabolic
the cyclopropyl amide portion with an amino pyrimidine group, and
stability while maintaining the potency on 11β-HSD-1 and ultimately
consecutively appending a cyano group on the dichlorophenyl ring to
produced a series of oxygen-linked 11β-HSD-1 inhibitors, represented
improve PK property. Compound 79 potently inhibited the TYK2 en-
by clinical candidate BMS-770767 (82), which has been evaluated in
zyme and the IL-23 pathway in cells, exhibited high selectivity over
humans as a potential treatment of type 2 diabetes.39b,c

Figure 27. Structures of Cat. K inhibitors 97, 98, odanacatib (99) and 100.

2546
S. Sun, et al. Bioorganic & Medicinal Chemistry Letters 29 (2019) 2535–2550

a-carbon to the adjacent proline methylene, still maintaining the cri-


tical H-bond interactions with Val-216, and a phenyl substituent on the
pyridine ring retains the hydrophobic interaction with the enzyme.
While compound 93 was highly active in vitro with a Ki of 4.5 nM, it was
not orally active at a dose of 20 mg/kg in an acute hemorrhagic assay,
largely due to its poor physical properties, e.g, low solubility (1.8 μg/
mL in saline). Subsequently, extensive optimization was conducted on
93, and eventually led to a series of orally active pyrimidone analogs.
Compound 94 showed excellent bioavailability in rats (62%) and dogs
(88%), and demonstrated an ED50 of 7.5 mg/kg in the acute hemor-
rhagic model (Fig. 25).44
Applying a similar strategy, Saitton et al. designed a 2,3,4-sub-
stitued pyridine derivative 96 (Fig. 26) by mimicking the naturally
occurring Pro-Leu-Gly-NH2 (PLG) tripeptide 95. The PLG-mimetic 96
Figure 28. Structures of ABCE1 inhibitors 101–104. was found to be more potent than the natural PLG in enhancing the
maximum response of the dopamine agonist N-propylapomorphine
Phthalazinylbenzamide 83 (Fig. 22) represents a class of potent and (NPA) at the human D2 receptor in a cell based assay with a maximum
selective inhibitors of p38a mitogen-activated protein kinase (MAPK). response of 146% at 10 μM (115% for PLG).45
PK studies of 83 in rats revealed a suboptimal profile (Cl = 1.7 L/h/kg,
t½ = 1.4 h).40a A cocrystal structure of 83 with p38a revealed that the Trifluoroethylamine as an amide isostere
carbonyl oxygen engages the NH of Asp168. Based on this observation,
the carbonyl group in 83 was fused on the phenyl ring to form a ben- Replacement of an amide with a trifluoroethylamine group was
zoisoxazole ring. The nitrogen lone pair of the benzoisoxazole isostere originally proposed by Zanda in consideration of both geometry and
was projected to mimic the carbonyl H-bonding interaction with Asp electronic effects.46 The electron withdrawing effect of the CF3 group
168. Analog 84 was found to maintain good potency and selectivity that reduces the basicity of the amine, and the electronegative fluorine atom
compound 83 had. Optimization of 84 was effected by replacement of has some mimicry of the carbonyl oxygen of the amide. While re-
the bottom phenyl substitute with a morpholine ring as well as re- placement of an amide group with the trifluoroethylamine moiety
placement of the cyclopropyl group with a methyl. This effort led to the eliminates metabolic instability to proteases and amidases, this mod-
discovery of compound 85 with equal potency on p38a and lacking of ification leads to synthetic complexities due to the introduction of a
CYP3A4 activity. Analog 85 showed better PK properties than 83 chiral center in the molecule.
(Cl = 1.4 L/h/kg, t½ = 2.8 h) and exhibited efficacy (ED50 = 0.05 mg/ One successful application of trifluoroethylamine as an amide iso-
kg) in the rat collagen induced arthritis (CIA) model.40b stere has been found in the design of cathepsin K (Cat K) inhibitor
Benzamide 86 (Fig. 23) was originally identified as a voltage-gated odanacatib (99), a compound advanced into phase III clinical evalua-
potassium channel (KV1.x) inhibitor by Merck.41 The main application tion as a potential treatment of osteoporosis in patients.47 As shown in
of this class of compounds was to block KV1.3 as potential treatments of Fig. 27, the cathepsin K inhibitor 97 was moderately selective for ca-
various diseases involving the immune system. Researchers at Bristol- thepsin K over the related enzyme cathepsin L. Replacement of the
Myers Squibb elaborated this scaffold utilizing conformational con- arylamide group in 97 with a trifluoroethylamine element resulted in
strains and amide isosteres by fusing the amide bond into an amino 98 with enhancements in both potency and selectivity. The finely op-
heterocyclic ring, while maintaining the H-bond features. These efforts timized product odanacatib is a non-basic molecule with the fluorinated
led to a class of novel, potent, and selective KV1.5 inhibitors, which valine blocking hydroxylation and the cyclopropyl moiety slowing the
were useful for treatment of IKur-associated disorders, including atrial amide hydrolysis.
fibrillation, exemplified by indazole 87 and pseudosaccharin amine While the CF3 group provided an ideal balance between potency,
88.42 selectivity and metabolic stability, odanacatib is a highly crystalline
An X-ray crystal structure of checkpoint 1 kinase (CHK1) inhibitor molecule with low water solubility, which translates into low oral
89 (Fig. 24) revealed that an internal H-bond between the amide CO bioavailability in preclinical species. To address this liability associated
and the urea NH groups constrains the conformation in a six-membered with odanacatib, the less lipophilic difluoroethylamine analogue 100
ring. Accordingly, 89 was rigidified into thienopyridine and thieno- (log D7 is about 3 log units lower) was designed with anticipation to
pyridazine class of inhibitors. It was found that these structurally dis- improve the solubility. It was also expected that the increased basicity
tinct analogs, e.g., compounds 90 and 91, maintained comparable in- (pKa is about one log unit higher) would enable the formation of che-
hibitory activity on CHK1 with IC50 values of 3 nM and 1 nM, mically stable salts. As anticipated, compound 100 displayed significant
respectively.43 improvement in oral bioavailability in rats and dogs, and the potency
Six-membered heterocycles pyridone and pyrimidone have been and selectivity were comparable to odanacatib. However, PK studies in
reported as amide surrogates in the design of a series of non-peptidic rats of several chemically stable salts of 100, e.g., hydrochloric acid
human leukocyte elastase (HLE) inhibitors. With the assistance of a salt, sulfuric acid salt, methanesulfonic acid salt, and p-toluenesulfonic
modeling study that docked inhibitor 92 into the active site of HLE, acid salt, showed no advantage over the neutral form.48
Brown et al. employed a substituted pyridone to build a bridge from the The trifluoroethylamine component was also employed in the

Figure 29. Calculated dipole moments of the amide


bond, fluoroalkene, trifluoromethylalkene50 and
chloroalkene51.

2547
S. Sun, et al. Bioorganic & Medicinal Chemistry Letters 29 (2019) 2535–2550

Figure 30. Structures of leu-enkephalin (105) and peptidomimetics (106–109).

required to overcome the CYP 2D6 inhibitory liability, compound 104


was 18–fold more potent than 102 and demonstrated efficacy in re-
ducing brain Aβx42 levels in mice.49 This example provided an alter-
native approach to mitigate an aniline structural alert.

Fluoroalkene as an amide isostere

The fluoroalkene moiety is considered more lipophilic than an


amide, which makes it an interesting metabolically stable amide iso-
Figure 31. Structure of HCV NS5A inhibitor 110. stere for enhancing membrane permeability. Besides the capability of
mitigating the alkene structural alert, the high electronegative fluorine
design of a series of BACE-1 inhibitors, where a methylene linker was atom mimics the carbonyl oxygen of the amide and provides a fluor-
placed between the amino and phenyl group to address an aniline oalkene with a substantial dipole moment (1.4 D) (see Fig. 29). How-
structural alert present in a class of literature BACE1 inhibitors. The ever, this motif is unable to mimic many important features of an amide
fundamental design was guided by an X-ray cocrystal structure of a group, e.g., it is neither a good H-bond donor nor a good H-bond ac-
related analog 101 and the recognition of a non-planar benzylamine ceptor.
moiety was able to adopt an orientation orthogonal to the plane of the Leu-enkephalin (105, Fig. 30) was a potent δ-opioid receptor ago-
fluorophenyl P1 ring to optimally engage the carbonyl oxygen of nist, but failed to produce analgesia in vivo due to poor PK properties,
Gly230 of the enzyme. SAR (see Fig. 28) revealed that the pKa of the e.g., rapid cleavage of the Tyr1-Gly2 bond by aminopeptidase N in
benzylic amine plays an important role in modulating the enzymatic human plasma (t½ = 0.69 min), and of the Gly3-Phe4 bond by angio-
activity. For example, difluoroethyl analog 103 (pKa of 5.2 for 103 vs tensin-converting enzyme at the blood–brain barrier (t½ = 130 min).52
3.8 for 102) was about four-fold less potent than trifluoroethyl analog Replacement of the Tyr1-Gly2 amide bond with a fluoroalkene was
102, along with an increase in MDR efflux (Er of 3.1 for 103 vs 1.2 for reported by Altman and colleagues. Although the effects of activating δ-
102). More basic analogs, e.g, the trifluoroethyl group was replaced and μ-opioid receptors were 60- and 45-fold weaker than 105, analog
with a 3,3,3-trifluoropropyl (pKa of 6.0) or a cyclopropylmethyl group 106 exhibited excellent stability in rat plasma (76% remaining of 106
(pKa of 6.3), were much less active. The CF3-cyclopropyl group pro- at 4 h vs t½ < 5 min for 105) and human plasma (68% remaining of
vided an ideal balance of basicity (pKa of 2.9) and geometry to fill the 106 at 4 h vs t½ = 12 min for 105). In contrast, trifluoethylamine de-
binding pocket properly. Although further structural optimization was rivative 107 (both S and R isomers) had no activity at δ-opioid and μ-
opioid receptors at 10 μM concetration.53

Figure 32. Structure of pentapeptide 111 and peptidomimetic 112.

2548
S. Sun, et al. Bioorganic & Medicinal Chemistry Letters 29 (2019) 2535–2550

trifluoromethylalkene as an amide isostere appears to be less ex-


plored.59

3-Aminooxetane as an amide isostere

The heterocyclic 3-aminooxetane unit as an amide isostere was in-


dependently conceived by Carreira60 and Shipman.61 Replacement of
amides with this non-hydrolyzable motif could, in principle, led to
metabolically stable mimetics. While the oxetane group is able to re-
duce the basicity of the amine by about 3 pKa units, it is a much weaker
H-bond acceptor than an amide carbonyl group,60a which translates
into suboptimal mimicry of an amide.
Carreira et al. modified the natural leu-enkephalin (105, see Fig. 33)
wherein four amide bonds were successively replaced by the 3-ami-
nooxetane structure.60d In this study, leu-enkephalin was found to ra-
pidly degrade in human serum with a half-life of about 10 min. The 3-
aminooxetane analogs 113 and 114 were stable in human serum,
however, lost binding affinity at the rat δ-opioid receptor. Analog 115
showed a slightly increased half-life of about 15 min with a consider-
able affinity of 157 nM. Analog 116 was about two times more stable
Figure 33. Structures of leu-enkephalin analogs 113–116. with an affinity comparable to leu-enkephalin (Ki of 9.2 nM for 105 in
this test). In addition, 116 exhibited analgesic activity in an in vivo
mouse assay. These results demonstrated that 3-aminooxetane can be a
valuable replacement of an amide, however, the effect on biological
activity and metabolic stability of replacing an amide bond by the 3-
aminooxetane moiety in a bioactive molecule is highly context depen-
dent.

Boronic acid as an amide replacement


Figure 34. Structures cefalotin (117) and vaborbactam (118).
There have been considerable efforts in the design of boronic acid
3 4 transition-state inhibitors, utilizing the boronic moiety as a replacement
Replacement of the Gly -Phe bond with a fluoroalkene was ac-
of the reactive β-lactam ring present in β-lactam antibiotics.62 These
complished by Nadon et al. Analog 108 displayed high binding affinity
novel inhibitors are anticipated to have fewer resistances because they
to δ-opioid receptor, suggesting that a H-bond donor at this site was not
are able to inhibit klebsiella pneumoniae carbapenemase (KPC) and
essential, the fluoroalkene isostere and the Gly3-Phe4 amide bond act as
other β-lactamases. The rationale behind these designs relies on the
a H-bond acceptor. The weak activity of the olefin 109 further high-
formation of a reversible covalent adduct between the active site serine
lighted the role of the fluorine atom for effective mimicry of the H-bond
in a carbapenemase and the boronic moiety, which mimics the tetra-
accepting feature of the amide.54
hedral TS during the serine carbapenemase-catalyzed lactam hydro-
Another example is the fluoroalkene containing analog of dacla-
lysis.
tasvir 110 (Fig. 31), which exhibited picomolar activity against HCV
Vaborbactam (118, Fig. 34) a boron-containing mimic of cefalotin
genotype 1b replicon. However, 110 was much less active on genotype
(117), is a potent inhibitor of serine carbapenemases, particularly, the
1a replicon than daclatasvir (structure 73, Fig. 18), suggesting this
KPC-producing strains, with no inhibition of mammalian serine pro-
genotype is highly sensitive to structural changes.55
teases.62e Clinical trials of vaborbactam were conducted in combination
In addition, fluoroalkene has also been applied in the design of di-
with meropenem by Rempex Pharmaceuticals, and eventually the
peptidyl peptidase IV inhibitors as a strategy to improve chemical and
combination drug vabomere was approved in 2017 by FDA for treat-
metabolic stability.56
ment of complicated urinary tract infections, including pyelone-
Waelchli and colleagues found that both a fluoroalkene and a
phritis.62g
chloroalkene were effective dipeptide isosteres in a series of human
parathyroid hormone agonists.57 Recently, the chloroalkene structure
was introduced into a cyclic pentapeptide 111 in the design of con- Conclusion
formationally rigid peptidomimetics. The chloroalkene containing
analog 112 (Fig. 32) showed about 20-fold higher activity against The design and application of amide isosteres have cultivated suc-
human dermal fibroblast (HDF) attachment to vitronectin compared to cess toward solving a range of problems in drug discovery. However,
peptide 111.58 one cannot overestimate the impact that an amide isostere might have
However, since the chlorine atom is less electronegative than a on a bioactive molecule. Isosteres are typically less than exact mimics;
fluorine (3.16 on the Pauling electronegativity scale compared to 3.98 an effective isostere for one specific circumstance might not translate
for F) and chloroalkene is more lipophilic than fluoroalkene (cLog P is into another case. The effect on biological activity and /or PK proper-
about 0.5 units higher), it is not clear if the chlorine atom could ef- ties of replacing an amide group with a certain isostere in a bioactive
fectively block the metabolic epoxidation of the alkene. molecule is highly context dependent, and could be beneficial or de-
Wipf has proposed that a trifluoromethyl group is a better electro- leterious, e.g., size, shape, electronic distribution, polarity, lipophili-
static mimic of the carbonyl oxygen of an amide than the fluorine of city, H-bond and pKa, potentially playing key roles in the interactions
fluoroalkene based on the fact that the dipole moment of tri- between the ligand and its receptor. While lead optimization is chal-
fluoromethylalkene (2.3D) is closer to that of an amide (3.6D) (see lenging, often no single method is able to solve every problem. The
Fig. 29). Despite the attractiveness of this hypothesis, the use of examples summarized here provide medicinal chemists with a resource
to solve the issues arising from their optimization of bioactive amides.

2549
S. Sun, et al. Bioorganic & Medicinal Chemistry Letters 29 (2019) 2535–2550

Acknowledgments (b) Reed CW, Washecheck JP, Quitlag MC, et al. Bioorg Med Chem Lett.
2019;29:1211.
25. Summa V, Petrocchi A, Bonelli F, et al. J Med Chem. 2008;51:5843.
We thank Dr. Steven S. Wesolowski, Dr. Christoph M. Dehnhardt 26. McBriar MD, Clader JW, Chu I, et al. Bioorg Med Chem Lett. 2008;18:215.
and Dr. James R. Empfield for review of the manuscript. We also ap- 27. Laurence C, Brameld KA, Graton J, et al. J Med Chem. 2009;52:4073.
preciate Dr. Peter R. Bernstein and the reviewers for suggestions during 28. Down K, Amour A, Baldwin IR, et al. J Med Chem. 2015;58:7381.
29. (a) Shao PP, Ok D, Fisher MH, et al. Bioorg Med Chem Lett. 2005;15:1901
revision of this manuscript. (b) Shao PP, Ye F, Weber AE, et al. Bioorg Med Chem Lett. 2009;19:5329
(c) Shao PP, Ye F, Weber AE, et al. Bioorg Med Chem Lett. 2009;19:5334.
References 30. (a) Charrier JD, Durrant SJ, Golec JM, et al. J Med Chem. 2011;54:2320
(b) Knegtel R, Charrier JD, Durrant SJ, et al. J Med Chem. 2019;62:5547.
31. (a) Tron GC, Pirali T, Billington RA, et al. Med Res Rev. 2008;28:278
1. (a) Patani GA, LaVoie EJ. Chem Rev. 1996;96:3147 (b) Bonandi E, Christodoulou MS, Fumagalli G, et al. Drug Discovery Today.
(b) Meanwell NA. J Med Chem. 2011;54:2529. 2017;22:157.
2. (a) Leung D, Abbenante G, Fairlie DP. J Med Chem. 2000;43:305 32. Doiron JE, Le CA, Ody BK, et al. Chem Eur J. 2019;25:3662.
(b) Adessi C, Soto C. Curr Med Chem. 2002;9:963 33. (a) Lagu B, Kluge AF, Fredenburg RA, et al. Bioorg Med Chem Lett. 2017;27:5230
(c) Venkatesan N, Kim BH. Curr Med Chem. 2002;9:2243 (b) Lagu B, Kluge AF, Tozzo E, et al. ACS Med Chem Lett. 2018;9:935.
(d) Hanessian S, Auzzas L. Acc Chem Res. 2008;41:1241 34. (a) Sun S, Zhang Z, Kodumuru V, et al. Bioorg. Med. Chem. Lett. 2014;24:520
(e) Seebach D, Gardiner J. Acc Chem Res. 2008;41:1366 (b) Sun S, Zhang Z, Pokrovskaia N, et al. Bioorg. Med. Chem. 2014;24:520
(f) Gentilucci L, De Marco P, Cerisoli L. Curr Pharm Des. 2010;16:3185 (c) Chowdhury S, Dales N, Fonarev J, et al. World patent application WO 2009/
(g) Avan I, Hall CD, Katritzky AR. Chem Soc Rev. 2014;43:3575. 103739, August 27, 2009.
3. (a) De Clercq E. J Med Chem. 1995;38:2491 35. Sun Z, Asberom T, Bara T, et al. J Med Chem. 2012;55:489.
(b) Bursavich MG, Rich DH. J Med Chem. 2002;45:541 36. (a) Belema M, Nguyen VN, Romine JL, et al. J Med Chem. 2014;57:1995
(c) Tsantrizos YS. Acc Chem Res. 2008;41:1252. (b) Belema M, Nguyen VN, Romine C, et al. J Med Chem. 2014;57:2013.
4. Roberts NA, Martin JA, Kinchington D, et al. Science. 1990;248:358. 37. Fletcher SR, McIver E, Lewis S, et al. Bioorg Med Chem Lett. 2006;16:2872.
5. Patick AK, Boritzki TJ, Bloom LA. Antimicrob Agents Chemother. 1997;41:2159. 38. Liang J, Abbema AV, Balazs M, et al. Bioorg Med Chem Lett. 2017;27:4370.
6. Kim EE, Baker CT, Dwyer MD, et al. J Am Chem Soc. 1995;117:1181. 39. (a) Wang H, Ruan Z, Li J, et al. Bioorg. Med. Chem. Lett. 2008;18:3168
7. Dorsey BD, Levin RB, McDaniel SL, et al. J Med Chem. 1994;37:3443. (b) Wang H, Robl JA, Hamann LG, et al. Bioorg. Med. Chem. Lett. 2011;21:4146
8. Kempf DJ, Sham HL, Marsh KC, et al. J Med Chem. 1998;41:602. (c) Robl JA, Wang H, Li JJ, et al. Abstracts of papers, 244th ACS national Meeting &
9. Sham HL, Kempf DJ, Molla A, et al. Antimicrob Agents Chemother. 1998;42:3218. Exposition, Philadelphia, PA, United States of America, August 19-23, 2012;
10. Ghosh A, Chapsal BD, Weber IT, Mitsuya H. Acc Chem Res. 2008;41:78. American Chemical Society; Washington, DC, 2012; MEDI-217.
11. Bold G, Fässler A, Capraro H, et al. J Med Chem. 1998;41:3387. 40. (a) Herberich B, Cao G, Chakrabarti P, et al. J Med Chem. 2008;51:6271
12. (a) Bühlmayer P, Caselli A, Fuhrer W, et al. J Med Chem. 1988;31:1839 (b) Pettus LH, Xu S, Cao G, et al. J Med Chem. 2008;51:6280.
(b) Rasetti V, Cohen NC, Rüeger H, et al. Bioorg Med Chem Lett. 1996;6:1589 41. Miao S, Bao J, Garcia ML, et al. Bioorg Med Chem Lett. 2003;13:1161.
(c) Göschke R, Cohen NC, Wood JM, Maibaum J. Bioorg Med Chem Lett. 42. (a) Johnson JA, Xu N, Jeon Y, et al. Bioorg Med Chem Lett. 2014;24:3018
1997;7:2735 (b) Lloyd J, Finlay HJ, Kover A, et al. Bioorg Med Chem Lett. 2015;25:4983.
(d) Göschke R, Stutz S, Rasetti V, et al. J Med Chem. 2007;50:4818 43. Zhao L, Zhang Y, Dai C, et al. Bioorg Med Chem Lett. 2010;20:7216.
(f) Maibaum J, Stutz S, Göschke R, et al. J Med Chem. 2007;50:4832 44. Brown FJ, Andisik DW, Bernstein PR, et al. J Med Chem. 1994;37:1259.
(g) Maibaum J, Feldman DL. Annu Rep Med Chem. 2009;44:105. 45. Saitton S, Del Tredici AL, Nina M, et al. J Med Chem. 2004;47:6595.
13. (a) Ghosh AK, Shin D, Downs D, et al. J Am Chem Soc. 2000;122:3522 46. (a) Volonterio A, Bravo P, Zanda M. Org Lett. 2000;2:1827
(b) Hom RK, Fang LR, Mamo S, et al. J Med Chem. 2003;46:1779 (b) Molteni M, Volonterio A, Zanda M. Org Lett. 2003;5:3887
(c) Hom RK, Gailunas AF, Mamo S, et al. J Med Chem. 2004;47:158 (c) Zanda M, New J. Chem. 2004;28:1401
(d) Stachel SJ, Coburn CA, Steele TG, et al. J Med Chem. 2004;47:6447 (d) Sani M, Volonterio A, Zanda M. ChemMedChem. 2007;2:1693.
(e) Maillard MC, Hom RK, Benson TE, et al. J Med Chem. 2007;50:776 47. (a) Black WC, Bayly CI, Davis DE, et al. Bioorg Med Chem Lett. 2005;15:4741
(f) Kortum SW, Benson TE, Bienkowski MJ, et al. Bioorg Med Chem Lett. (b) Gauthier JY, Chauret N, Cromlish W, et al. Bioorg Med Chem Lett. 2008;18:923.
2007;17:3378 48. Isabel E, Mellon C, Boyd MJ, et al. Bioorg Med Chem Lett. 2011;21:920.
(g) Iserloh U, Wu Y, Cumming JN, et al. Bioorg Med Chem Lett. 2008;18:418 49. Butler CR, Ogilvie K, Martinez-Alsina L, et al. J Med Chem. 2017;60:386.
(h) Cumming JN, Le TX, Babu S. Bioorg Med Chem Lett. 2008;18:3236 50. Wipf P, Henninger TC, Geib SJ. J Org Chem. 1998;63:6088.
(i) Cumming J, Babu S, Huang Y, et al. Bioorg Med Chem Lett. 2010;20:2837 51. Bogers MT. J Am Chem Soc. 1947;69:1243.
(j) Kaller MR, Harried SS, Albrecht B, et al. J Med Chem. 2012;55:886 52. (a) Weinberger SB, Martinez JL. J Pharmacol Exp Ther. 1988;247:129
(k) Weiss MM, Williamson T, Babu-Khan S, et al. J Med Chem. 2012;55:9009 (b) Thompson SE, Audus KL. Peptides. 1994;15:109.
(l) Dineen TA, Weiss MM, Williamson T, et al. J Med Chem. 2012;55:9025 53. (a) Karad SN, Pal M, Crowley RS, et al. ChemMedChem. 2017;12:571
(m) Ghosh AK, Rao KV, Yadav ND, et al. J Med Chem. 2012;55:9195 (b) Altman RA, Sharma KK, Rajewski LG, et al. ACS Chem Neurosci. 2018;9:1735.
(n) Bueno AB, Agejas J, Broughton H, et al. J Med Chem. 2017;60:9807. 54. Nadon J, Rochon K, Grastilleur S, et al. ACS Chem Neurosci. 2017;8:40.
14. Calcaterra NE, Barrow JC. ACS Chem Neurosci. 2014;5:253. 55. Chang W, Mosley RT, Bansal S, et al. J. Bioorg. Med. Chem. Lett. 2012;22:2938.
15. (a) Zingales IA. J. Chromatogr. 1973;75:55 56. (a) Lin J, Toscano PJ, Welch JT. Proc Nat Acad Sci USA. 1998;95:14020
(b) Ono S, Hatanaka T, Miyazawa S, et al. Xenobiotica. 1996;26:1155 (b) Zhao K, Lim DS, Funaki T, Welch JT. Bioorg Med Chem. 2003;11:207
(c) Greenblatt DJ, Divoll MK, Soong MH, et al. J. Clin. Pharmacol. 1988;28:853 (c) Van der Veken P, Senten K, Kertèsz I, et al. J Med Chem. 2005;48:1768
(d) Andersson T, Miners JO, Veronese ME, Birkett DJ. Br. J. Clin. Pharmacol. (d) Edmondson SD, Wei L, Xu J, et al. Bioorg Med Chem Lett. 2008;18:2409.
1994;38:131 57. Waelchli R, Gamse R, Bauer W, et al. Bioorg Med Chem Lett. 1996;6:1151.
(e) Hooper WD, Watt JA, Mckinnon GE, Reilly PEB. Eur. J. Drug Metab. 58. Kobayakawa T, Matsuzaki Y, Hozumi K, et al. ACS Med Chem Lett. 2018;9:6.
Pharmacokinet. 1992;17:51. 59. Xiao J, Weisblum B, Wipf P. J Am Chem Soc. 2005;127:5742.
16. Braestrup C, Squires RF. Eur J Pharmacol. 1978;48:263. 60. (a) Wuitschik G, Carreira EM, Wagner B, et al. J Med Chem. 2010;53:3227
17. (a) Hillesta L, Hansen T, Melsom H, Drivenes A. Clin. Pharmacol. Ther. 1974;16:479 (b) Burkhard JA, Wuitschik G, Plancher J, et al. Org Lett. 2013;15:4312
(b) Hillesta L, Hansen T, Melsom H. Clin. Pharmacol. Ther. 1974;16:485. (c) McLaughlin M, Yazaki R, Fessard TC, Carreira EM. Org Lett. 2014;16:4070
18. Hester Jr JB, Rudzik AD, Kamdar BV. J Med Chem. 1971;14:1078. (d) Möller GP, Müller S, Wolfstädter BT, et al. Org Lett. 2017;19:2510.
19. Walser A, Fryer RI, Benjamin L. US Patent 4166185, issued to Hoffmann-LaRoche 61. (a) Powell NH, Clarkson GJ, Notman R, et al. Chem Commun. 2014;50:8797
Aug 28, 1979. (b) Beadle JD, Knuhtsen A, Hoose A, et al. Org Lett. 2017;19:3303.
20. Paone DV, Nguyen DN, Shaw AW, et al. Bioorg Med Chem Lett. 2011;21:2683. 62. (a) Ness S, Martin R, Kindler AM, et al. Biochemistry. 2000;39:5312
21. Elliott JM, Carlson EJ, Chicchi GG, et al. Bioorg Med Chem Lett. 2006;16:2929. (b) Morandi F, Caselli E, Morandi S, et al. J Am Chem Soc. 2003;125:685
22. (a) Boström J, Hogner A, Llinàs A, et al. J Med Chem. 2012;55:1817 (c) Morandi S, Morandi F, Caselli E, et al. Bioorg Med Chem. 2008;16:1195
(b) Ladduwahetty T, Baker R, Cascieri MA, et al. J Med Chem. 1996;39:2907 (d) Winkler ML, Rodkey EA, Taracila MA, et al. J Med Chem. 2013;56:1084
(c) Borg S, Vollinga RC, Labarre M, et al. J Med Chem. 1999;42:4331 (e) Hecker SJ, Reddy KR, Totrov M, et al. J Med Chem. 2015;58:3682
(d) Rajapakse HA, Nantermet PG, Selnick HG, et al. J Med Chem. 2006;49:7270 (f) Caselli E, Romagnoli C, Vahabi R, et al. J Med Chem. 2015;58:5445
(e) Nakajima K, Chatelain R, Clairmont KB, et al. J Med Chem. 2017;60:4657. (g) Burgos RM, Biagi MJ, Rodvold KA, Danziger LH. Expert Opin Drug Metab Toxicol.
23. Gillman KW, Starrett Jr JE, Parker MF, et al. ACS Med Chem Lett. 2010;1:120. 2018;14:1007.
24. (a) Reed CW, McGowan KM, Spearing PK, et al. ACS Med Chem Lett. 2017;8:1326

2550

You might also like