HIV Integrase

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

European Journal of Medicinal Chemistry 225 (2021) 113787

Contents lists available at ScienceDirect

European Journal of Medicinal Chemistry


journal homepage: http://www.elsevier.com/locate/ejmech

Advances in the development of HIV integrase strand transfer


inhibitors
Yue Wang a, Shuang-Xi Gu b, c, *, Qiuqin He a, **, Renhua Fan a
a
Department of Chemistry, Fudan University, 2005 Songhu Road, Yangpu District, Shanghai, 200438, China
b
Key Laboratory for Green Chemical Process of Ministry of Education, School of Chemical Engineering & Pharmacy, Wuhan Institute of Technology, Wuhan,
430205, China
c
Hubei Key Laboratory of Novel Reactor and Green Chemical Technology, Wuhan Institute of Technology, Wuhan, 430205, China

a r t i c l e i n f o a b s t r a c t

Article history: HIV-1 integrase (IN) is a key enzyme in viral replication that catalyzes the covalent integration of viral
Received 1 May 2021 cDNA into the host genome. Currently, five HIV-1 IN strand transfer inhibitors (INSTIs) are approved for
Received in revised form clinical use. These drugs represent an important addition to the armamentarium for antiretroviral
5 August 2021
therapy. This review briefly illustrates the development history of INSTIs. The characteristics of the
Accepted 5 August 2021
Available online 18 August 2021
currently approved INSTIs, as well as their future perspectives, are critically discussed.
© 2021 Elsevier Masson SAS. All rights reserved.
Keywords:
HIV-1 integrase
Strand transfer inhibitors
Metal chelating
Resistance

1. Introduction clinical use (Fig. 1): elvitegravir (EVG) in 2012 [8], dolutegravir
(DTG) in 2013 [9,10], bictegravir (BIC) in 2018 [11] and cabotegravir
Acquired immunodeficiency syndrome (AIDS) is a chronic, life- (CAB) in 2021 [12]. These FDA-approved drugs act by inhibiting the
threatening disease caused by the human immunodeficiency vi- IN strand transfer reaction and are referred to as IN strand transfer
rus (HIV). Despite the availability of more than 40 antiviral drugs inhibitors (INSTIs).
approved for therapeutic use, the HIV epidemic remains one of the
most serious global health threats [1]. The World Health Organi-
zation (WHO) estimated that approximately 38 million people were 2. HIV-1 IN structure and function
living with HIV-1 infection at the end of 2020, with 1.7 million new
HIV infections and 0.7 million deaths related to AIDS [2]. HIV-1 IN is a 32 kDa protein comprised of three structural do-
HIV-1 integrase (IN), along with reverse transcriptase (RT) and mains: the N-terminal domain (NTD), the catalytic core domain
protease (PR), is an essential enzyme for retroviral replication [3]. (CCD), and the C-terminal domain (CTD) [13e15]. The structure of
HIV-1 IN has no mammalian counterpart and thus represents an each isolated domain has been determined, as well as the bound
attractive target for the development of anti-HIV drugs [4,5]. In fact, structures of CCD with NTD [16] and CTD [17]. Recombinant full-
although anti-HIV drugs targeting RT and PR have been used for length IN in solution has also been reported to exist in different
decades, it was not until 2007 that IN was validated as a viable forms ranging from monomer to tetramer to higher-order species
target for HIV therapy by discovering the first IN inhibitor, ralte- [18e22].
gravir (RAL, Fig. 1) [6,7]. Since then, four additional IN inhibitors The NTD (residues 1e50) adopts a zinc finger fold containing a
have been approved by the Food and Drug Administration (FDA) for pair of highly conserved histidine and cysteine residues (HHCC
motif) [23e25]. The CTD (residues 213e288) adopts an SH3-like
fold and is implicated in DNA binding [26]. The CCD (residues
51e212) contains the catalytic triad D64, D116, E152 (DDE motif)
* Corresponding author. Department of Chemistry, Fudan University, 2005
Songhu Road, Yangpu District, Shanghai, 200438, China.
within an RNase H-like fold that comprises the active site [27,28].
** Corresponding author. The DDE motif has been proposed to bind two divalent metals, Mg
E-mail addresses: shuangxigu@163.com (S.-X. Gu), qqhe@fudan.edu.cn (Q. He). (II) or Mn (II), which are essential for catalytic activity. Also, Mg is

https://doi.org/10.1016/j.ejmech.2021.113787
0223-5234/© 2021 Elsevier Masson SAS. All rights reserved.
Y. Wang, S.-X. Gu, Q. He et al. European Journal of Medicinal Chemistry 225 (2021) 113787

Fig. 1. The FDA approved INSTIs.

considered to be the metal cofactor of HIV-1 IN under physiological polyhydroxylated aromatic compounds [40e42], and DNA com-
conditions [29]. plexes [43] have been reported to act as IN inhibitors, most of these
Over the past decade, intasome structures from several retro- compounds only showed inhibitory activity in in vitro assays. The
viruses have been determined by X-ray crystallography [30e33] or discovery of the b-diketoacid (DKA) class as HIV-1 IN inhibitors in
single-particle cryogenic electron microscopy (cryo-EM) [34e38]. the cell-based assay was a major advance in validating IN as a
The pioneering studies conducted on the prototype foamy virus therapeutically antiretroviral drug target. DKAs, exemplified by L-
(PFV) intasome have provided initial snapshots of the structure and 731988 [44] (1, Fig. 3), is chemically characterized by a diketo acid
function of the retroviral intasome. Furthermore, the crystal moiety linked to an aromatic ring. The diketo acid moiety (g-ke-
structures of PFV intasome-drug complexes have revealed the tone, enolizable a-ketone, and carboxylic acid) was believed to
mechanistic details of the drug action [30]. However, PFV and HIV-1 interact with two Mg2þ ions and the end of viral DNA, blocking the
INs share only ~15% sequence identity, and many of the sites where ST reaction (Fig. 2). S-1360 [45] (3, Fig. 3), 5-ClTEP [46] (3, Fig. 3)
drug-resistance mutations occur in HIV-1 IN are not conserved in and L-870810 [47,48] (4, Fig. 3) also belong to the DKA family as
PFV IN Refs. [36,37]. To better understand the interaction of INSTIs they can be regarded as diketo acid bioisosteric analogues. All the
with HIV-1 intasomes, the structures of primate immunodeficiency bioisosteres have three functional groups that mimic g-ketone,
virus intasomes including HIV-1 and red-capped mangabey simian enolizable a-ketone, and carboxylic acid and possess a coplanar
immunodeficiency virus (SIVrcm) in both unbound and INSTI- conformation [49]. S-1360 was the first IN inhibitor to enter clinical
bound forms were determined by cryo-EM to near-atomic resolu- trials, but it failed in the phase II clinical trial owing to its rapid
tion [36,37]. metabolism [50].
The biochemical, genetic, and complementation studies indicate
that HIV-1 IN functions as a multimer [18e20,22]. Minimally, a
tetramer is required for the concerted integration of a pair of vDNA 3.2. First-generation INSTIs
ends into the target DNA. The HIV-1 IN tetramer stabilizes through
its interaction with two vDNA ends [19,22]. The integration pro- RAL, derived from the evolution of DKAs in the HCV polymerase
moted by IN is comprised of two distinct steps. Initially, the enzyme program by Merck Research Laboratories, was approved for clinical
recognizes the long terminal repeats (LTR) termini of the viral use as the first HIV-1 IN inhibitor in 2007 [6,7]. RAL is orally
dsDNA. It selectively cleaves the terminal dinucleotides from the 30 - administrated in 400 mg twice daily. It inhibited IN-catalyzed ST
ends of vDNA in the cytoplasm of an infected cell (30 -processing with an IC50 value of 10 nM and blocked the spread of HIV infection
reaction). Subsequently, IN bound to the processed vDNA ends in cell culture with an EC95 value of 33 nM [51]. It retains strong
enters the nucleus, where the joining reaction catalyzed by IN activity against diverse clinical isolates resistant to other classes of
connects 30 -OH ends of vDNA to the host DNA (strand transfer re- drugs, including NRTIs, NNRTIs, and PIs. The approval of RAL
action, ST) [4,5].

3. HIV-1 INSTIs

3.1. Earlier HIV-1 INSTIs

Since the cofactor Mg2þ plays a vital role in HIV-1 catalytic


functions, the initial investigations on HIV-1 IN inhibitors were
mainly focused on screening the compounds with the structural
elements of a bidentate metal-binding pharmacophore. Although
compounds with various structural features, such as peptides [39], Fig. 2. Originally proposed chelating mechanism for DKAs in IN catalytic site.

2
Y. Wang, S.-X. Gu, Q. He et al. European Journal of Medicinal Chemistry 225 (2021) 113787

confirmed that the coplanar three specifically positioned hetero-


atoms coordinate the two Mg2þ (Fig. 4a). The p-F-benzyl group
displaces and substitutes for the 3’ terminal adenosine of the pro-
cessed vDNA, forming a p-stacking interaction with the penulti-
mate cytidine residue. The terminal oxadiazole ring makes a well-
oriented p-stacking interaction with Y212 (corresponding to Y143
in HIV-1 IN). This contact is likely important and unique for RAL,
which could be the reason for the Y143 mutation on the drug po-
tency. In the crystal structures, although PFV IN residues Q217 and
N224 (corresponding to Q148 and N155 in HIV-1 IN, respectively)
do not directly contact RAL, the mutations of these two residues
may induce conformational changes within the catalytic pocket
and result in a decreased binding efficacy.
EVG, the second IN inhibitor, was approved in 2012 by FDA [8]. It
has a reduced dosing frequency (once-daily, 150 mg), but it requires
a pharmacokinetic (PK) booster to increase EVG systemic exposure
[54]. The structural feature of EVG is the 4-quinolone-3-carboxylic
acid scaffold. The coplanar monoketo acid motif in 4-quinolone-3-
carboxylic acid could be an alternative to the diketo acid motif [8].
EVG showed a significant IC50 value of 7.2 nM against HIV-1 IN and
an EC50 value of 0.9 nM in cell-based assay [8]. The binding mode of
EVG in the active site of PFV is similar to that of RAL (Fig. 4b). The
coplanar monoketo acid motif binds two Mg2þ ions. The halobenzyl
moiety displaces 3’ adenine of bound viral DNA and forms a well-
oriented p-stacking interaction with the penultimate cytidine
residue of the processed strand [30]. The reduced susceptibility to
EVG was associated with T66I, E92Q, S147G, Q148 H/K/R and
N155H, suggesting a significant cross-resistance between ralte-
gravir and EVG cross-resistance [55].

3.3. Second-generation INSTIs


Fig. 3. The structures of the DKAs and its bioisosteric analogues. The three functional
groups in DKAs are indicated in blue. The three heteroatoms chelating with two Mg2þ An intramolecular tethering design based on carbamoyl pyr-
ions are indicated with dashed lines in orange. idone explored by Shionogi company led to the discovery of DTG as
the second-generation HIV-1 IN inhibitor [56]. DTG differs from the
first-generation INSTIs, where its chelating motif is located on a
provided a second chance to patients who were left with almost no tricyclic scaffold. In addition, the linker that connects the central
treatment alternatives after the failure of highly active antiretro- chelating moiety to the halogenated benzyl group is longer than
viral therapy (HAART) at that time. However, clinical resistance to either in RAL or EVG [57]. It was approved for medical use in 2013
RAL was observed. HIV-1 mutations that give rise to raltegravir with a once-daily, unboosted dose (50 mg) [58]. Compared to the
resistance map to IN involve three primary genetic pathways: first-generation INSTIs, DTG had a distinct resistance profile. It
Q148 H/R/K, N155H and less frequently, Y143C [52]. G140S is an exhibited in vitro wild-type (WT) or near WT activity against site-
important compensatory mutation for Q148 mutations and the directed mutants with the single primary resistance mutations
double mutant G140S/Q148H was observed in clinical trials as one observed in vivo at Q148, at N155, and at Y143 [59]. It displayed
of the most prevalent profiles of RAL resistance [53]. The cocrystal remarkably potent in vitro activity against the troublesome Q148K
structures of full-length IN from PFV intasome with RAL [30] mutant (IC50 ¼ 0.68 nM). Its long plasma terminal half-life and slow

Fig. 4. Crystallographic RAL (a, from PDB 3OYA) and EVG (b, from PDB 3L2U) in the active site of PFV intasome.

3
Y. Wang, S.-X. Gu, Q. He et al. European Journal of Medicinal Chemistry 225 (2021) 113787

dissociation rate from the IN complex (dissociative t1/2 ¼ 71 h, 3.4. Other leading INIs
dissociation was 8 times and 26 times slower than RAL and EVG,
respectively) suggested that it should present a high barrier to The success of the drugs in targeting INST has sparked signifi-
resistance development [60]. In 2019, WHO recommended DTG- cant interest in developing the next-generation INSTIs. The main
containing regimens as the preferred first- and second-line treat- focus is identifying the new chemical classes based on the metal-
ment for all patients with HIV [61]. chelating mechanism with an improved genetic barrier to resis-
SAR studies on DTG revealed that the modification of oxaza- tance and enhanced patient compliance.
bridging combined with introducing an additional fluoro atom on A series of bicyclic inhibitors 5e7 (Fig. 6) were designed using a
the 2,4-difluorophenyl ring would significantly improve the anti- hydroxylamide group as a metal-chelating component [69]. A
viral potency against G140S/Q148H, leading to the discovery of a halobenzylamide functionality was attached through a linker, and
potent and unboosted inhibitor BIC [62]. BIC potently inhibited its carboxamide carbonyl group is not part of the key metal-
G140S/Q148H with an EC50 value of 3.1 nM and displayed a sta- chelating triad. Compound 5 showed less active in the ST inhibi-
tistically improved resistance profile compared to RAL, EVG and tory potency than the corresponding compounds 6 and 7 (Table 1).
DTG against a set panel of patient-derived INSTI-resistant viral As shown in Table 2, in cellular assays, inhibitors 6 and 7 displayed
isolates [63]. Additionally, BIC exhibited a substantially longer EC50 values at nanomolar concentrations against WT virus. More-
dissociation half-life from HIV-1 IN-DNA complexes than DTG [64]. over, 6c and 7c retained strong potency against a panel of IN mu-
The most recently approved CAB is also structurally similar to tants, including Y143R, N155H, G140S/Q148H, G118R, and E138K/
DTG [12]. Although the potency of CAB is lower than DTG and BIC, it Q148K (Table 3). Further structural modification at the 4、6、7-
has the unique feature of a long half-life (close to 40 days) [65]. It positions of compound 7 through incorporating a N-containing
can be formulated as a nanoparticle injection so that it can be given functional group at the 4-position led to 8 (Fig. 6), resulting into
at prolonged intervals [65]. The combination of CAB and rilpivirine improved potency against vectors harboring the major forms of
(a NNRTI) was approved by FDA in Jan 2021 as the first long-acting drug-resistant IN mutants [70e72]. The most potent 8a exhibited a
injection for HIV-1 treatment. Such a combination of two long- superior antiviral activity against a broad range of INSTI-resistant
acting injectable agents is called cabenuva. It can be adminis- mutants compared to DTG (only the selected data are shown in
trated once a month and replace the currently used daily pills to Table 4; for complete data, please check Ref 72). Compared to DTG,
treat HIV-1 infection, marking a significant change from the it also exhibited stronger activity against several triple mutants
methods followed in the past [66]. such as E138K/G140A/Q148K, L74M/G140A/Q148R, and L74 M/
Both BIC and CAB displayed strong inhibitory activity against the G140C/Q148R. The cryo-EM structures of HIV-1 with bound 8a
RAL-resistant IN mutants Y143R and N155H, the EVG-resistant IN revealed the key contacts in the substrate binding region (Fig. 7).
mutants T66I and E92Q, and the DTG-resistant IN mutant H51Y and The amino group at the 4-position of 8a forms an intramolecular
E138K/R263K with EC50 values of <5 nM. However, only BIC hydrogen bond with the halobenzylamide oxygen, stabilizes the
potently inhibited the well-known RAL-resistant IN double mutant planar conformation, and facilitates metal chelation. Moreover, the
G140S/Q148H and the DTG-resistant IN mutant G118R, R263K, and electron and/or inductive effects on the aromatic core enhance the
H51Y/R263K with EC50 values of 5 nM [67]. metal coordination strength. The long chain of the hydroxyl group
The structural analysis of DTG and BIC-bound in the active site of points directs the bulk solvent, displaces the loosely bound water,
SIVrcm has been investigated recently (Fig. 5) [37]. The three cen- and provides an additional enthalpic gain. The crystallographic 8a
tral electronegative heteroatoms and the halogenated benzyl in the active site of HIV-1 intasome suggests the importance of the
moiety make interactions with SIVrcm similar to those found in interactions with the b4-a2 loop [36]. The inhibitor that binds
RAL-PFV intasome crystals. Importantly, DTG and BIC intimately within the envelope defined by both the viral and host DNA sub-
contact backbone N117 and G118 from the b4-a2 loop, a crucial strates assists 8a in retaining broad efficacy.
feature of the second-generation INSTIs. Although the second- Systematic optimization of N-methylpyrimidinone carbox-
generation INIs have been developed to overcome cross- amides led to the discovery of BMS-707035 (9, Fig. 8) which
resistance with the first-generation drugs, several newly identi- incorporated with an embedded morpholine heterocycle [73].
fied resistance mutations such as G118R, R263K, and S153Y have BMS-707035 showed an excellent IC50 value of 3 nM against ST
been selected for the second-generation INIs [68]. Furthermore, reaction and an EC50 value of 17 nM in cell-based assays in the
recent studies suggested that the mutations emerging from the presence of human serum albumin. BMS-707035 displayed low
Q148 pathway can confer cross resistance to the second-generation clearance in the rat, dog and monkey. It also demonstrated good
INIs. safety margins in toxicity studies using dogs and rats. However,

Fig. 5. Binding modes of DTG (a, from PDB 6RWN) and BIC (b, from PDB 6RWM) in the active site of SIVrcm intasome.

4
Y. Wang, S.-X. Gu, Q. He et al. European Journal of Medicinal Chemistry 225 (2021) 113787

Fig. 6. The design of bicyclic inhibitors 5e8 with a hydroxylamide functionality.

Table 1
Strand transfer inhibitory potency of bicyclic inhibitors 5e7.

No. R IC50 (nM)


0 0
5a 3 -Cl-4 -F 530
5b 30 ,40 -diF 780
5c 20 ,40 -diF 1800
6a 30 -Cl-40 -F 140
6b 30 ,40 -diF 170
6c 20 ,40 -diF 410
7a 30 -Cl-40 -F 270
7b 30 ,40 -diF 400
7c 20 ,40 -diF 550

Table 2
Antiviral potencies of bicyclic inhibitors 6e7 in cellular assays.

No. R EC50 (nM) SIc


a b b b
WT Y143R N155H G140S/Q148H
Fig. 7. Binding mode of compound 8a in the active site of HIV-1 intasome (from PDB
6a 30 -Cl-40 -F 38 34 90 N/Ad >6579 6PUY).
6b 30 ,40 -diF 62 40 2200 N/A >4032
6c 20 ,40 -diF 5.1 4.9 134 438 >49020
7a 30 -Cl-40 -F 35 54 148 489 2914
BMS-707035 was inactive against G140S/Q148H mutant virus with
7b 30 ,40 -diF 20 45 189 507 9600
7c 20 ,40 -diF 6.2 11 31 308 22097 an EC50 value of >8 mM.
a
The mutations associated with the Q148 pathway would cause a
Cells were infected with the lentiviral vector harboring WT IN, and the values
are indicated in EC50.
structural shift of the position of Mg2þ [74]. To enhance the binding
b
Cells were infected with the viral vectors carrying IN mutations, and the values capacity of IN inhibitors, the amide moiety of BMS-707035 was
are indicated in EC50. further replaced with azole heterocycles, generating a set of het-
c
Selectivity index calculated as the ratio of CC50 to EC50. erocycle amide isosteres in the pyrimidinones (10, Fig. 8) [74]. The
d
Not available.
utilizaiton of azole heterocylces would afford a more open geom-
etry and exhibit greater tolerance of the altered active site of Q148
Table 3 mutants. Among the azole heterocycle derivatives, imidazole
Fold change (FC) of 6c and 7c compared with that of RAL. replacement (10a, Table 5) provided the most substantial
improvement in capacity for resistance coverage. It showed potent
No. EC50 (nM)a
activity in cellular assays with EC50 values of 33 nM, 320 nM, and
a
WT Y143Rb N155Hb G140S/Q148Hb G118Rb E138K/Q148Kb
2 nM against Q148R, G140S/Q148H, and N155H mutant virus,
RAL 4 41  38  475  9 375  respectively (Table 5). The binding modes of compounds 9 and 10a
6c 5.1 1 26  86  N/Ab N/A in the active site of the homology model of WT IN and G140S/
7c 6.2 2 5 50  6 32 
Q148H IN, respectively, revealed that the mutation of glutamine
a
Cells were infected with the lentiviral vector harboring WT IN, and the values are 148 to histidine leads to a more open pocket that is no longer well
indicated in EC50. bCells were infected with viral vectors carrying IN mutations, and
occupied by 9. Meanwhile, a more extended geometry of 10 oc-
the indicated values correspond to the fold-change (FC) in EC50 relative to WT. bNot
available. cupies much of the increased space. Combining the wider geometry
of azoles and its effect on Mg2þ binding coupled with better steric
occupation of the mutant active site contributed to improving the
Table 4 resistance coverage of 10.
Comparison of antiviral potencies of 8a relative to DTG. The screening of proprietary collection of HIV-1 proper INIs
No. EC50 (nM)a against WT and G140S/Q148H mutant viruses sponsored by Bristol-
L74 M M50I G140S/Q148H G148H/N155H E138K/Q148K
Myers Squibb Research and Development led to the identification
of bridged tricyclic pyrimidinone series [75]. The [3.2.2]-bridged
DTG 2.2 2.1 3.9 4.0 25
tricyclic analog 11 (Fig. 9) exhibited an EC50 value of 0.7 nM
8a 1.6 1.3 0.9 1.1 16
against WT HIV-1 and 3 nM against the G140S/Q148H mutant virus.
a
Selected EC50 values obtained from cells infected with the indicated IN mutants.

5
Y. Wang, S.-X. Gu, Q. He et al. European Journal of Medicinal Chemistry 225 (2021) 113787

Fig. 8. The replacement of amide moiety of 9 with azole heterocycles. Calculated “bite angel” are indicated with dashed lines in blue.

Table 5 displayed strong IN inhibitory potency (IC50 ¼ 0.7 nM) with


The inhibitory activity of 10a compared with 9. reasonable small serum shifts. Furthermore, it showed a minor shift
No. het EC50 (nM)a against the mutations with a 1.4-fold and 2.8-fold loss in potency
against Q148K and G140S/Q148H, respectively. This agent had a
WT Q148R G140S/Q148H N155H
good t1/2, low i.v. clearance, but poor oral bioavailability in rats and
9 3 300 >8000 140 dogs.
10a 6 33 320 2
A novel bicyclic template tetrahydronaphthyridine combining
the structural features from DKAs was designed by Merck & Co., Inc.
A carboxylic amide para to the phenolic oxygen was considered to
a
Antiviral activity assayed in an HIV infectivity cellular assay. be an important structural element since similar substitution in the
1,6-naphthyridine series, exemplified by RAL, greatly improved the
metabolic stability. An intensive exploration of the structure-
activity relationship (SAR) of the tetrahydronaphthyridine system
has led to the discovery of MK-0536 (13, Fig. 11) [77,78]. It exhibited
an excellent antiviral activity in a viral kinetic assay and maintained
good potency against the mutations E92Q, Y143R, Q148R, Q148K,
and N155H, with only a maximum 3-fold change against these
mutants (Table 6). It is believed that the potency against a panel of
mutant virus improves significantly with the substitution of the
steric bulk group at the N-2 position of the tetrahydronaphthyr-
idine system. Additionally, MK-0536 showed a reasonable pre-
Fig. 9. The structure of [3.2.2]-bridged tricyclic analog 11.
clinical PK profile in rat (CLp ¼ 2.3 mL/min/kg, %F ¼ 29) and dog
(CLp ¼ 2.2 mL/min/kg, %F ¼ 44).
However, this inhibitor was not considered for further develop- To better understand the influence of the structural diversity of
ment due to unfavorable human dosing projections predicted by the tetrahydronaphthyridine system, a third ring was introduced
the pharmacokinetic (PK) profiles in the preclinical species. using MK-0536 as a starting point [79]. The resulting fused 5-
The combination of structural features from the tricyclic ring membered ring system exhibited an optimal balance between po-
system carbamoylpyridone and naphthyridinone generated the tency and PK. Also, an extensive structural optimization efforts
hydroxyquinoline tetracyclic ring derivatives 12 (Fig. 10) [76]. All generated a novel bicycle[3.1.0]hexane ring system (14, Fig. 11) with
the designed compounds exhibited excellent inhibitory activities a good resistance profile against RAL-resistant mutants. Further
against IN with EC50 values in the range of 0.6e18 nM. Most of them optimization was directed towards improving the serum-shifted
illustrated a slight loss of potency against the IN signature resis- potency by increasing the polar surface area (PSA), which pro-
tance mutations Q148K and G140S/Q148H. The best compound 12a vided compound 15 containing a chiral secondary alcohol group on

Fig. 10. The design of hydroxyquinoline tetracyclic ring derivatives 12.

6
Y. Wang, S.-X. Gu, Q. He et al. European Journal of Medicinal Chemistry 225 (2021) 113787

Fig. 11. Lead optimization starting from MK-0536.

Table 6 to MK-0636 and the first-generation INSTIs, and resemble the in-
Fold change (FC) of MK-0536 compared with that of RAL. teractions observed with DTG, which may be responsible for the
No. WTa EC50 (nM)b excellent potency of 15 and 16 against a panel of RAL-resistant
E92Q Y143R Q148R Q148K N155H
mutations (Fig. 13).

RAL 9/53 3 13  16  22  8
MK-0536 6/53 1 2 1 1 3 4. Conclusion and perspectives
15 2/67 1.5  1 1.5  1 1.5 
16 3/32 1 1 1 1 1 Since the approval of RAL as the first INSTI in 2007, five drugs
a
Cells were infected with viral vectors carrying IN mutations and the indicated targeting IN ST step have been approved by FDA. Moreover, a
values correspond to the fold change (FC) in EC50 relative to WT as measured in a number of structurally diverse INSTIs which are able to retain the
viral kinetic assay. intrinsic potency against the mutant virus have been identified. In
b
Measured in the presence of 0/50% NHS.
all the cases, there is a halogenated benzyl ring that interacts with
the penultimate cytosine base near the 30 -end of the viral DNA, a
the alkyl chain of the amine nitrogen as well as compound 16 bidentate metal-binding pharmacophore whose function is to
bearing a hydroxymethyl substituent at the optimal position on the chelate the two Mg2þ held in place by the DDE motif, and a linker.
bicycle[3.1.0]hexane ring (Fig. 11). Both compounds exhibited The first-generation INSTIs RAL and EVG suffer from a relatively low
excellent potency against WT virus and extremely low fold-change barrier to resistance development. The second-generation INSTIs
in potency against RAL-resistant mutations (Table 6). Additionally, show improved efficacies against RAL and EVG-resistant strains of
compounds 15 and 16 also displayed favorable PK profiles, which HIV. It is apparent that the second-generation INSTIs have relatively
anticipates a once-daily human dose without the need for a PK longer linkers compared to the first-generation INSTIs and are
booster. However, they suffered from less than dose proportionality better able to adapt to changes in the active-site geometry caused
during oral exposure due to their poor physicochemical properties. by resistance mutations [70]. Clinically significant resistance to the
To address this limitation, a prodrug strategy was undertaken to second-generation INSTIs is rare. The success of the second-
enhance the solubility and permeability, which led to carbonate generation INSTIs and recent structural studies using both HIV-1
acetal/phosphate prodrug 15a and carbonate acetal prodrug 16a and SIVrcm provide implications for the design of next-
with high exposure (Fig. 12). generation INSTIs [36,37].
Both the cocrystal structures of PFV intasome with 15 and 16 The main challenges in designing next-generation INSTIs are
revealed that the bicycle[3.1.0]hexane ring system of 15 and 16 reducing dosing frequency to once daily without a PK booster and
render an additional contact with G187 in the b4-a2 loop compared improving the genetic barrier to resistance, especially for G140S/
Q148H. The INSTIs share a similar mode of action; they chelate two

Fig. 12. A prodrug strategy to advance 15a and 16a.

7
Y. Wang, S.-X. Gu, Q. He et al. European Journal of Medicinal Chemistry 225 (2021) 113787

Fig. 13. Binding modes of compounds 15 (a, from PDB 4ZTF) and 16 (b, from PDB 4ZTJ) in the active site of PFV intasome.

Mg2þ in the IN catalytic site, creating the potential to develop cross- Abbreviations
resistance. To successfully obtain the next-generation INSTIs,
extending the inhibitors toward the IN backbone to fill these spaces HIV-1 human immunodeficiency virus type 1
more fully should be concerned in the design process. The proposed IN integrase
approaches involve: 1) introducing a proper lengthy linker to INSTIs integrase strand transfer inhibitors
connect a bidentate metal-binding pharmacophore and a haloge- AIDS Acquired immunodeficiency syndrome
nated benzyl ring. The nature of this linker can help adapt to WHO The World Health Organization
changes in the active-site geometry caused by resistance muta- RT reverse transcriptase
tions, 2) filling the substrate envelope. The rationale is that the PR protease
inhibitor which interacts more completely with the highly RAL raltegravir
conserved bases will be less prone to inducing resistance, and 3) EVG elvitegravir
making extensive interactions with the b4-a2 loop. DTG dolutegravir
Recently, attempts to identify inhibitors that block the integra- BIC bictegravir
tion process through a distinct mechanism of action have provided CAB cabotegravir
an alternative approach to overcome cross-resistance, leading to FDA Food and Drug Administration
the development of allosteric IN inhibitors (ALLINIs). ALLINIs, NTD N-terminal domain
alternatively referred to as lens epithelium-derived growth factor CCD catalytic core domain
(LEDGF/p75) integrase inhibitors (LEDGINs), noncatalytic site IN CTD C-terminal domain
inhibitors (NCINIs), or IN-LEDGF/p75 allosteric inhibitors (INLAIs), vDNA viral DNA
bind to the dimer interface of IN CCD at the main LEDGF/p75 LTR long terminal repeats
binding site [80e84]. Although originally designed to block IN- HAART highly active antiretroviral therapy
LEDGF/p75 interactions during viral integration, these inhibitors PFV prototype foamy virus
could also induce aberrant multimerization during virion matura- PK pharmacokinetic
tion [80,82,85]. In addition, ALLINIs retarget residual integrants FC fold change
away from transcription units and toward a more repressive ALLINIs allosteric IN inhibitors
chromatin environment, providing a potential “block-and-lock” LEDGF Lens Epithelium Derived Growth Factor
functional cure strategy [86,87]. To date, ALLINIs haven't yet LEDGINs LEDGF/p75 integrase inhibitors
approved for use in humans, but there has been growing interest in NCINIs noncatalytic site integrase inhibitors
the development of ALLINIs due to their immense potential to INLAIs integrase-LEDGF/p75 allosteric inhibitors
overcome HIV-1 drug resistance.
References

[1] HIV and AIDS: medicines to help you. https://www.fda.gov/consumers/free-


Declaration of competing interest publications-women/hiv-and-aids-medicines-help-you, 2020. (Accessed 1
July 2020) accessed.
The authors declare that they have no known competing [2] World Health Organization, UNAIDS Report on the Global AIDS Epidemic,
2020.
financial interests or personal relationships that could have
[3] S.D. Young, Inhibition of HIV-1 integrase by small molecules: the potential for
appeared to influence the work reported in this paper. a new class of AIDS chemotherapeutics, Curr. Opin. Drug Discov. Dev 4 (2001)
402e410.
[4] D. Esposito, R. Craigie, HIV integrase structure and function, Adv. Virus Res. 52
(1999) 319e333.
[5] E. Asante-Appiah, A.M. Skalka, HIV-1 integrase: structural organization,
Acknowledgement conformational changes, and catalysis, Adv. Virus Res. 52 (1999) 351e369.
[6] V. Summa, A. Petrocchi, F. Bonelli, B. Crescenzi, M. Donghi, M. Ferrara, F. Fiore,
This work was supported by the National Natural Science C. Gardelli, O.G. Paz, D.J. Hazuda, P. Jones, O. Kinzel, R. Laufer, E. Monteagudo,
E. Muraglia, E. Nizi, F. Orvieto, P. Pace, G. Pescatore, R. Scarpelli, K. Stillmock,
Foundation of China (21971043, 81861138046, 21877087), the Sci-
M.V. Witmer, M. Rowley, Discovery of raltegravir, a potent, selective orally
ence and Technology Commission of Shanghai Municipality bioavailable HIV-integrase inhibitor for the treatment of HIV-AIDS infection,
(17XD1404400, 18XD1400800, 19ZR1403400), Key Laboratory for J. Med. Chem. 51 (2008) 5843e5855.
Green Chemical Process of Ministry of Education (GCP20200201), [7] M. Anker, R.B. Corales, Raltegravir (MK-0518): a novel integrase inhibitor for
the treatment of HIV infection, Expet Opin. Invest. Drugs 17 (2008) 97e103.
and Hubei Key Laboratory of Novel Reactor and Green Chemical [8] K. Shimura, E.N. Kodama, Elvitegravir: a new HIV integrase inhibitor, Antiviral
Technology (40201002). Chem. Chemother. 20 (2009) 79e85.

8
Y. Wang, S.-X. Gu, Q. He et al. European Journal of Medicinal Chemistry 225 (2021) 113787

[9] L. Vandekerckhove, GSK-1349572, a novel integrase inhibitor for the treat- integration, Science 355 (2017) 93e95.
ment of HIV infection, Curr. Opin. Invest. Drugs 11 (2010) 203e212. [36] D.O. Passos, M. Li, I.K. Jozwik, X.Z. Zhao, D. Santos-Martins, R. Yang, S.J. Smith,
[10] C.E. Kandel, S.L. Walmsley, Dolutegravir e a review of the pharmacology, ef- Y. Jeon, S. Forli, S.H. Hughes, T.R. Burke Jr., R. Craigie, D. Lyumkis, Structural
ficacy, and safety in the treatment of HIV, Drug Des. Dev. Ther. 9 (2015) basis for strand-transfer inhibitor binding to HIV intasomes, Science 367
3547e3555. (2020) 810e814.
[11] E.D. Deeks, Bictegravir/emtricitabine/tenofovir alafenamide: a review in HIV-1 [37] N.J. Cook, W. Li, D. Berta, M. Badaoui, A. Ballandras-Colas, A. Nans, A. Kotecha,
infection, Drugs 78 (2018) 1817e1828. E. Rosta, A.N. Engelman, P. Cherepanov, Structural basis of second-generation
[12] Novel drug approvals for 2021. https://www.fda.gov/drugs/new-drugs-fda- HIV integrase inhibitor action and viral resistance, Science 367 (2020)
cders-new-molecular-entities-and-new-therapeutic-biological-products/ 806e810.
novel-drug-approvals-2021, 2021. (Accessed 3 August 2021). [38] K.K. Pandey, S. Bera, K. Shi, M.J. Rau, A.V. Oleru, J.A.J. Fitzpatrick,
[13] P.J. Lodi, J.A. Ernst, J. Kuszewski, A.B. Hickman, A. Engelman, R. Craigie, A.N. Engelman, H. Aihara, D.P. Grandgentt, Cryo-EM structure of the Rous
G.M. Clore, A.M. Gronenborn, Solution structure of the DNA binding domain of sarcoma virus octameric cleaved synaptic complex intasome, Commun. Biol. 4
HIV-1 integrase, Biochemistry 34 (1995) 9826e9833. (2021) 330.
[14] M. Cai, Y. Huang, M. Caffrey, R. Zheng, R. Craigie, G.M. Clore, A.M. Gronenborn, [39] R.G. Maroun, S. Gayet, M.S. Benleulmi, H. Porumb, L. Zargarian, H. Merad,
Solution structure of the His 12/Cys mutant of the N-terminal zinc binding H. Leh, J.-F. Moucadet, F. Troalen, S. Fermandjian, Peptide inhibitors of HIV-1
domain of HIV-1 integrase complexed to cadmium, Protein Sci. 7 (1998) integrase dissociate the enzyme oligomers, Biochemistry 40 (2001)
2669e2674. 13840e13848.
[15] Y. Galdgur, R. Craigie, G.H. Cohen, T. Fujiwara, T. Yoshinaga, T. Fujishita, [40] M.R. Fesen, K.W. Kohn, F. Leteurtre, Y. Pommier, Inhibitors of human immu-
H. Sugimoto, T. Endo, H. Murai, D.R. Davies, Structure of the HIV-1 integrase nodeficiency virus integrase, Proc. Natl. Acad. Sci. U.S.A. 90 (1993)
catalytic domain complexed with an inhibitor: a platform for antiviral drug 2399e2403.
design, Proc. Natl. Acad. Sci. U.S.A. 96 (1999) 13040e13043. [41] A. Mazumder, A. Gazit, A. Levitzki, M. Nicklaus, J. Yung, G. Kohlhagen,
[16] J.-Y. Wang, H. Ling, W. Yang, R. Craigie, Structure of a two-domain fragment of Y. Pommier, Effects of tyrphostins, protein kinase inhibitors, on human im-
HIV-1 integrase: implications for domain organization in the intact protein, munodeficiency virus type 1 integrase, Biochemistry 34 (1995) 15111e15122.
EMBO J. 20 (2001) 7333e7343. [42] Z. Lin, N. Neamati, H. Zhao, Y. Kiryu, J.A. Turpin, C. Aberham, K. Strebel,
[17] J.C.-H. Chen, J. Krucinski, L.J. Miercke, J.S. Finer-Moore, A.H. Tang, A.D. Leavitt, K. Kohn, M. Witvrouw, C. Pannecouque, Z. Debyser, E. De Clercq, W.G. Rice,
R.M. Stroud, Crystal structure of the HIV-1 integrase catalytic core and C- Y. Pommier, T.R. Burke Jr., Chicoric acid analogues as HIV-1 integrase in-
terminal domains: a model for viral DNA binding, Proc. Natl. Acad. Sci. U.S.A. hibitors, J. Med. Chem. 42 (1999) 1401e1414.
97 (2000) 8233e8238. [43] N. Jing, X. Xu, Rational drug design of DNA oligonucleotides as HIV inhibitors,
[18] A. Engelman, F.D. Bushman, R. Craigie, Identification of discrete functional Curr. Drug Targets - Infect. Disord. 1 (2001) 79e90.
domains of HIV-1 integrase and their organization within an active multi- [44] D.J. Hazuda, P. Felock, M. Witmer, A. Wolfe, K. Stillmock, J.A. Grobler,
meric complex, EMBO J. 12 (1993) 3269e3275. A. Espeseth, L. Gabryelski, W. Schleif, C. Blau, M.D. Miller, Inhibitors of strand
[19] P. Cherepanov, G. Maertens, P. Proost, B. Devreese, J. Van Beeumen, transfer that prevent integration and inhibit HIV-1 replication in cells, Science
Y. Engelborghs, E. De Clercq, Z. Debyser, HIV-1 integrase forms stable tetra- 287 (2000) 646e650.
mers and associates with LEDGF/p75 protein in human cells, J. Biol. Chem. 278 [45] M.L. Barreca, S. Ferro, A. Rao, L. De Luca, M. Zappala , A.-M. Monforte,
(2003) 372e381. Z. Debyser, M. Witvrouw, A. Chimirri, Pharmacophore-based design of HIV-1
[20] A. Faure, C. Calmels, C. Desjobert, M. Castroviejo, A. Caumont-Sarcos, integrase strand-transfer inhibitors, J. Med. Chem. 48 (2005) 7084e7088.
L. Tarrago-Litvak, S. Litvak, V. Parissi, HIV-1 integrase crosslinked oligomers [46] R.J. Herr, 5-Substituted-1H-tetrazoles as carboxylic acid isosteres: medicinal
are active in vitro, Nucleic Acids Res. 33 (2005) 977e986. chemistry and synthetic methods, Bioorg. Med. Chem. 10 (2002) 3379e3393.
[21] C.J. McKee, J.J. Kessl, N. Shkriabai, M.J. Dar, A. Engelman, M. Kvaratskhelia, [47] D.J. Hazuda, S.D. Young, J.P. Guare, N.J. Anthony, R.P. Gomez, J.S. Wai,
Dynamic modulation of HIV-1 integrase structure and function by cellular J.P. Vacca, L. Handt, S.L. Motzel, H.J. Klein, G. Dornadula, R.M. Danovich,
lens epithelium-derived growth factor (LEDGF) protein, J. Biol. Chem. 283 M.V. Witmer, K.A.A. Wilson, L. Tussey, W.A. Schleif, L.S. Gabryelski, L. Jin,
(2008) 31802e31812. M.D. Miller, D.R. Casimiro, E.A. Emini, J.W. Shiver, Integrase inhibitors and
[22] S. Hare, F. Di Nunzio, A. Labeja, J. Wang, A. Engelman, P. Cherepanov, Struc- cellular immunity suppress retroviral replication in rhesus macaques, Science
tural basis for functional tetramerization of lentiviral integrase, PLoS Pathog. 5 305 (2004) 528e532.
(2009), e1000515. [48] L. Zhuang, J.S. Wai, M.W. Embrey, T.E. Fisher, M.S. Egbertson, L.S. Payne,
[23] R. Zheng, T.M. Jenkins, R. Craigie, Zinc folds the N-terminal domain of HIV-1 J.P. Guare, J.P. Vacca, D.J. Hazuda, P.J. Felock, A.L. Wolfe, K.A. Stillmock,
integrase, promotes multimerization, and enhances catalytic activity, Proc. M.V. Witmer, G. Moyer, W.A. Schleif, L.J. Gabryelski, Y.M. Leonard, J.J. Lynch,
Natl. Acad. Sci. U.S.A. 93 (1996) 13659e13664. S.R. Michelson, S.D. Yong, Design and synthesis of 8-hydroxy-[1,6]naphthyr-
[24] M. Cai, R. Zheng, M. Cafrey, R. Craigie, G.M. Clore, A.M. Gronenborn, Solution idines as novel inhibitors of HIV-1 integrase in vitro and in infected cells,
structure of the N-terminal zinc binding of HIV-1 integrase, Nat. Struct. Biol. 4 J. Med. Chem. 46 (2003) 453e456.
(1997) 567e577. [49] M. Sato, T. Motomura, H. Aramaki, T. Matsuda, M. Yamashita, Y. Ito,
[25] Z. Zhao, C.J. McKee, J.J. Kessl, W.L. Santos, J.E. Daigle, A. Engelman, G. Verdine, H. Kawakami, Y. Matsuzaki, W. Watanabe, K. Yamataka, S. Ikeda, E. Kodama,
M. Kvaratskhelia, Subunit-specific protein footprinting reveals signifcant M. Matsuoka, H. Shinkai, Novel HIV-1 integrase inhibitors derived from qui-
structural rear-rangements and a role for N-terminal Lys-14 of HIV-1 inte- nolone antibiotics, J. Med. Chem. 49 (2006) 1506e1508.
grase during viral DNA binding, J. Biol. Chem. 283 (2008) 5632e5641. [50] A. Billich, S-1360 Shionogi-GlaxoSmithKline, Curr. Opin. Invest. Drugs 4
[26] A.P.A.M. Eijkelenboom, R. Sprangers, K. Hård, R.A. Puras Lutzke, (2003) 206e209.
R.H.A. Plasterk, R. Boelens, R. Kaptein, Refined solution structure of the C- [51] J.D. Croxtall, K.A. Lyseng-Williamson, C.M. Perry, Raltegravir, Drugs 68 (2008)
terminal DNA-binding domain of human immunovirus-1 integrase, Proteins 131e138.
36 (1999) 556e564. [52] J.S. Cavalcanti, A.M. Lança, J.L. de Paula Ferreira, M. da Eira, D.S. de Souza
[27] A. Engelman, R. Craigie, Identification of conserved amino acid residues crit- Dantas, L.F. de Macedo Brígido, In-vivo selection of the mutation F121Y in a
ical for human immunodeficiency virus type 1 integrase function in vitro, patient failing raltegravir containing salvage regimen, Antivir. Res. 95 (2012)
J. Virol. 66 (1992) 6361e6369. 9e11.
[28] F. Dyda, A.B. Hickman, T.M. Jenkins, A. Engelman, R. Craigie, D.R. Davies, [53] K.O. Delelis, I. Malet, L. Na, L. Tchertanov, V. Calvez, A.-G. Marcelin, F. Subra,
Crystal structure of the catalytic domain of HIV-1 integrase: similarity to other E. Deprez, J.-F. Mouscadet, The G140S mutation in HIV integrases from
polynucleotidyl transferases, Science 266 (1994) 1981e1986. raltegravir-resistant patients rescues catalytic defect due to the resistance
[29] O. Delelis, K. Carayon, A. Saïb, E. Deprez, J.-F. Mouscadet, Integrase and inte- Q148H mutation, Nucleic Acids Res. 37 (2009) 1193e1201.
gration: biochemical activities of HIV-1 integrase, Retrovirology 5 (2008) 114. [54] S. Ramanathan, A.A. Mathias, P. German, B.P. Kearney, Clinical pharmacoki-
[30] S. Hare, S.S. Gupta, E. Valkov, A. Engelman, P. Cherepanov, Retroviral intasome netic and pharmacodynamic profile of the HIV integrase inhibitor elvitegravir,
assembly and inhibition of DNA strand transfer, Nature 464 (2010) 232e236. Clin. Pharmacokinet. 50 (2011) 229e244.
[31] S. Hare, A.M. Vos, R.F. Clayton, J.W. Thuring, M.D. Cummings, P. Cherepanov, [55] K. Shimura, E. Kodama, Y. Sakagami, Y. Matsuzaki, W. Watanabe, K. Yamataka,
Molecular mechanisms of retroviral integrase inhibition and the evolution of Y. Watanabe, Y. Ohata, S. Doi, M. Sato, M. Kano, S. Ikeda, M. Matsuoka, Broad
viral resistance, Proc. Natl. Acad. Sci. U.S.A. 107 (2010) 20057e20062. antiretroviral activity and resistance profile of the novel human immunode-
[32] Z. Yin, K. Shi, S. Banerjee, K.K. Pandey, S. Bera, D.P. Grandgentt, H. Aihara, ficiency virus integrase inhibitor elvitegravir (JTK-303/GS-9137), J. Virol. 82
Crystal structure of the Rous sarcoma virus intasome, Nature 530 (2016) (2008) 764e774.
362e366. [56] M. Kobayashi, T. Yoshinaga, T. Seki, C. Wakasa-Morimoto, K.W. Brown,
[33] G. Eilers, K. Gupta, A. Allen, J. Zhou, Y. Hwang, M.B. Cory, F.D. Bushman, G. Van R. Ferris, S.A. Foster, R.J. Hazen, S. Miki, A. Suyama-Kagitani, S. Kawauchi-Miki,
Duyne, Influence of the amino-terminal sequence on the structure and T. Taishi, T. Kawasuji, B.A. Johns, M.R. Underwood, E.P. Garvey, A. Sato,
function of HIV integrase, Retrovirology 17 (2020) 28. T. Fujiwara, In vitro antiretroviral properties of S/GSK1349572, a next-
[34] A. Ballandras-Colas, M. Brown, N.J. Cook, T.G. Dewdney, B. Demeler, generation HIV integrase inhibitor, Antimicrob. Agents Chemother. 55
P. Cherepanov, D. Lyumkis, A.N. Engelman, Cryo-EM reveals a novel octameric (2011) 813e821.
integrase structure for betaretroviral intasome function, Nature 530 (2016) [57] S. Hare, S.J. Smith, M. Me tifiot, A. Jaxa-Chamiec, Y. Pommier, S.H. Hughes,
358e361. P. Cherepanov, Structural and functional analyses of the second-generation
[35] A. Ballandras-Colas, D.P. Maskell, E. Serrao, J. Locke, P. Swuec, S.R. Jo nsson, integrase strand transfer inhibitor dolutegravir (S/GSK1349572), Mol. Phar-
A. Kotecha, N.J. Cook, V.E. Pye, I.A. Taylor, V. Andre sdottir, A.N. Engelman, macol. 80 (2011) 565e572.
A. Costa, P. Cherepanov, A supermolecular assembly mediates lentiviral DNA [58] H.-J. Stellbrink, J. Reynes, A. Lazzarin, E. Voronin, F. Pulido, F. Felizarta,

9
Y. Wang, S.-X. Gu, Q. He et al. European Journal of Medicinal Chemistry 225 (2021) 113787

S. Almond, M. St Clair, N. Flack, S. Min, Dolutegravir in antiretroviral-naive evaluation of BMS-707035, a potent HIV-1 integrase strand transfer inhibitor,
adults with HIV-1: 96-week results from a randomized dose-ranging study, Bioorg. Med. Chem. Lett 28 (2018) 2124e2130.
AIDS 27 (2013) 1771e1778. [74] K.M. Peese, B.N. Naidu, M. Patel, C. Li, D.R. Langley, B. Terry, T. Protack, V. Gali,
[59] J.M. Llibre, F. Pulido, F. García, M. García Deltoro, J.L. Blanco, R. Delgado, Ge- Z. Lin, H.K. Samanta, M. Zheng, S. Jenkins, I.B. Dicker, M.R. Krystal,
netic barrier to resistance for dolutegravir, AIDS Rev. 17 (2015) 59e68. N.A. Meanwell, M.A. Walker, Heterocycle amide isosteres: an approach to
[60] K.E. Hightower, R. Wang, F. DeAnda, B.A. Johns, K. Weaver, Y. Shen, overcoming resistance for HIV-1 integrase strand transfer inhibitors, Bioorg.
G.H. Tomberlin, H.L. Carter 3rd, T. Broderick, S. Sigethy, T. Seki, M. Kobayashi, Med. Chem. Lett 30 (2020) 126784.
M.R. Underwood, Dolutegravir (S/GSK1349572) exhibits significantly slower [75] M. Patel, B.N. Naidu, I. Dicker, H. Higley, Z. Lin, B. Terry, T. Protack, M. Krystal,
dissociation than raltegravir and elvitegravir from wild-type and integrase S. Jenkins, D. Parker, C. Panja, R. Rampulla, A. Mathur, N.A. Meanwell,
inhibitor-resistant HIV-1 integrase-DNA complexes, Antimicrob. Agents Che- M.A. Walker, Design, synthesis and SAR study of bridged tricyclic pyr-
mother. 55 (2011) 4552e4559. imidinone carboxamides as HIV-1 integrase inhibitors, Bioorg. Med. Chem. 28
[61] World Health Organization. Update of Recommendations on First- and (2020) 115541.
Second-Line Antiretroviral Regimens. [76] E.J. Velthuisen, B.A. Johns, D.P. Temelkoff, K.W. Brown, S.C. Danehower, The
[62] S.E. Lazerwith, R. Cai, X. Chen, G. Chin, M.C. Desai, S. Eng, R. Jacques, M. Ji, design of 8-hydroxyquinoline tetracyclic lactams as HIV-1 integrase strand
G. Jones, H. Martin, C. McMahon, M. Mish, P. Morganelli, J. Mwang, H.-J. Pyun, transfer inhibitors, Eur. J. Med. Chem. 117 (2016) 99e112.
U. Schmitz, G. Stepan, J. Szwarcberg, J. Tang, M. Tsiang, J. Wang, K. Wang, [77] M. Me tifiot, B. Johnson, S. Smith, X.Z. Zhao, C. Marchand, T. Burke, S. Hughes,
K. White, L. Wiser, J. Zack, H. Jin, Discovery of Bictegravir (GS-9883), a Novel, Y. Pommier, MK-0536 inhibits HIV-1 integrases resistant to raltegravir,
Unboosted, Once-Daily HIV-1 Integrase Strand Transfer Inhibitor (INSTI) with Antimicrob. Agents Chemother. 55 (2011) 5127e5133.
Improved Pharmacokinetics and in Vitro Resistance Profile, Proceedings of the [78] M.S. Egbertson, J.S. Wai, M. Cameron, R.S. Hoerrner, Discovery of MK-0536: a
ASM Microbe, Boston, MA, USA, 2016. June 16e20. potential second-generation HIV-1 integrase strand transfer inhibitor with a
[63] M. Tsiang, G.S. Jones, J. Goldsmith, A. Mulato, D. Hansen, E. Kan, L. Tsai, high genetic barrier to mutation, in: W.M. Kazmierski (Ed.), Antiviral Drugs:
R.A. Bam, G. Stepan, K.M. Stray, A. Niedziela-Majka, S.R. Yant, H. Yu, G. Kukolj, from Basic Discovery through Clinical Trials, John Wiley & Sons, Inc., New
T. Cihlar, S.E. Lazerwith, K.L. White, H. Jin, Antiviral activity of Bictegravir (GS- York, 2011, pp. 163e180.
9883), a novel potent HIV-1 integrase strand transfer inhibitor with an [79] I.T. Raheem, A.M. Walji, D. Klein, J.M. Sanders, D.A. Powell, P. Abeywickrema,
improved resistance profile, Antimicrob. Agents Chemother. 60 (2016) G. Barbe, A. Bennet, S.-D. Clas, D. Dubost, M. Embrey, J. Grobler, M.J. Hafey,
7086e7097. T.J. Hartingh, D.J. Hazuda, M.D. Miller, K.P. Moore, N. Pajkovic, S. Patel, V. Rada,
[64] K.L. White, N. Osman, E. Cuadra-Foy, B.G. Brenner, D. Shivakumar, P. Rearden, J.D. Schreier, J. Sisko, T.G. Steele, J.-F. Truchon, J. Wai, M. Xu,
F. Campigotto, M. Tsiang, P.A. Morganelli, N. Novikov, S.E. Lazerwith, H. Jin, P.J. Coleman, Discovery of 2-pyridinone aminals: a prodrug strategy to
A. Niedziela-Majka, Long dissociation of bictegravir from HIV-1 integrase- advance a second generation of HIV-1 integrase strand transfer inhibitors,
DNA complexes, Antimicrob. Agents Chemother. 65 (2021) e02406e20. J. Med. Chem. 58 (2015) 8154e8165.
[65] D. Hodge, D.J. Back, S. Gibbons, S.H. Khoo, C. Marzolini, Pharmacokinetics and [80] J.J. Kessl, N. Jena, Y. Koh, H. Taskent-Sezgin, A. Slaughter, L. Feng, S. de Silva,
drug-drug interactions of long-acting intramuscular cabotegravir and rilpi- L. Wu, S.F.J. Le Grice, A. Engelman, J.R. Fuchs, M. Kvaratskhelia, Multimode,
virine, Clin. Pharmacokinet. 60 (2021) 835e853. cooperative mechanism of action of allosteric HIV-1 integrase inhibitors,
[66] FDA approves first extended-release, injectable drug regimen for adults living J. Biol. Chem. 287 (2012) 16801e16811.
with HIV. https://www.fda.gov/news-events/press-announcements/fda- [81] J. Demeulemeester, P. Chaltin, A. Marchand, M. De Maeyer, Z. Debyser,
approves-first-extended-release-injectable-drug-regimen-adults-living-hiv, F. Christ, LEDGINs, non-catalytic site inhibitors of HIV-1 integrase: a patent
2021. (Accessed 21 January 2021). review (2006-2014), Expert Opin. Ther. Pat. 24 (2014) 609e632.
[67] S.J. Smith, X.Z. Zhao, T.R. Burke Jr., S.H. Hughes, Efficacies of cabotegravir and [82] F. Christ, S. Shaw, J. Demeulemeester, B.A. Desimmie, A. Marchand, S. Butler,
bictegravir against drug-resistant HIV-1 integrase mutants, Retrovirology 15 W. Smets, P. Chaltin, M. Westby, Z. Debyser, C. Pickford, Small-molecule in-
(2018) 37. hibitors of the LEDGF/p75 binding site of integrase block HIV replication and
[68] M.A. Wainberg, T. Mesplede, P.K. Quashie, The development of novel HIV modulate integrase multimerization, Antimicrob. Agents Chemother. 56
integrase inhibitors and the problem of drug resistance, Curr. Opin. Virol. 2 (2012) 4365e4374.
(2012) 656e662. [83] M. Balakrishnan, S.R. Yant, L. Tsai, C. O'Sullivan, R.A. Bam, A. Tsai, A. Niedziela-
[69] X.Z. Zhao, S.J. Smith, M. Me tifiot, B.C. Johnson, C. Marchand, Y. Pommier, Majka, K.M. Stray, R. Sakowicz, T. Cihlar, Non-catalytic site HIV-1 integrase
S.H. Hughes, T.R. Burke Jr., Bicyclic 1-hydroxy-2-oxo-1,2-dihydropyridine-3- inhibitors disrupt core maturation and induce a reverse transcription block in
carboxamide-containing HIV-1 integrase inhibitors having high antiviral po- target cells, PloS One 8 (2013), e74163.
tency against cells harboring raltegravir-resistant integrase mutants, J. Med. [84] S. Sugiyama, T. Akiyama, Y. Taoda, T. Iwaki, E. Matsuoka, E. Akihisa, T. Seki,
Chem. 57 (2014) 1573e1582. T. Yoshinaga, T. Kawasuji, Discovery of novel HIV-1 integrase-LEDGF/p75
[70] X.Z. Zhao, S.J. Smith, M. Me tifiot, C. Marchand, P.L. Boyer, Y. Pommier, allosteric inhibitors based on a pyridine scaffold forming an intramolecular
S.H. Hughes, T.R. Burke Jr., 4-Amino-1-hydroxy-2-oxo-1,8-naphthyridine- hydrogen bond, Bioorg, Med. Chem. Lett. 33 (2021) 127742.
containing compounds having high potency against raltegravir-resistant [85] M. Tsiang, G.S. Jones, A. Niedziela-Majka, E. Kan, E.B. Lansdon, W. Huang,
integrase mutants of HIV-1, J. Med. Chem. 57 (2014) 5190e5202. M. Hung, D. Samuel, N. Novikov, Y. Xu, M. Mitchell, H. Guo, K. Babaoglu, X. Liu,
[71] X.Z. Zhao, S.J. Smith, D.P. Maskell, M. Me tifiot, V.E. Pye, K. Fesen, C. Marchand, R. Geleziunas, R. Sakowicz, New class of HIV-1 integrase (IN) inhibitors with a
Y. Pommier, P. Cherepanov, S.H. Hughes, T.R. Burke Jr., Structure-guided dual mode of action, J. Biol. Chem. 287 (2012) 21189e21203.
optimization of HIV integrase strand transfer inhibitors, J. Med. Chem. 60 [86] Z. Debyser, G. Vansant, A. Bruggemans, J. Janssens, F. Christ, Insight in HIV
(2017) 7315e7332. integration site selection provides a block-and-lock strategy for a functional
[72] S.J. Smith, X.Z. Zhao, T.R. Burke Jr., S.H. Hughes, HIV-1 integrase inhibitors that cure of HIV infection, Viruses 11 (2019) 12.
are broadly effective against drug-resistant mutants, Antimicrob, Agents [87] A. Bruggemans, G. Vansant, M. Balakrishnan, M.L. Mitchell, R. Cai, F. Christ,
Chemother 62 (2018) e01035e18. Z. Debyser, GS-9822, a preclinical LEDGIN candidate, displays a block-and-
[73] B.N. Naidu, M.A. Walker, M.E. Sorenson, Y. Ueda, J.D. Matiskella, T.P. Connolly, lock phenotype in cell culture, Antimicrob, Agents Chemother 65 (2021)
I.B. Dicker, Z. Lin, S. Bollini, B.J. Terry, H. Higley, M. Zheng, D.D. Parker, D. Wu, e02328-20.
S. Adams, M.R. Krystal, N.A. Meanwell, The discovery and preclinical

10

You might also like