You are on page 1of 34

pubs.acs.

org/jmc Perspective

Chemical Biology of Sortase A Inhibition: A Gateway to Anti-


infective Therapeutic Agents
Rachit Sapra, Amit K. Rajora, Pushpendra Kumar, Govind P. Maurya, Nalin Pant,* and V. Haridas*
Cite This: J. Med. Chem. 2021, 64, 13097−13130 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Staphylococcus aureus is the leading cause of


hospital-acquired infections. The enzyme sortase A, present on
the cell surface of S. aureus, plays a key role in bacterial virulence
without affecting the bacterial viability. Inhibition of sortase A
Downloaded via UNIV PARIS-SUD on August 27, 2023 at 02:41:39 (UTC).

activity offers a powerful but clinically less explored therapeutic


strategy, as it offers the possibility of not inducing any selective
pressure on the bacteria to evolve drug-resistant strains. In this
Perspective, we offer a chemical space narrative for the design of
sortase A inhibitors, as delineated into three broad domains:
peptidomimetics, natural products, and synthetic small molecules.
This provides immense opportunities for medicinal chemists to
alleviate the ever-growing crisis of antibiotic resistance.

1. INTRODUCTION nosocomial infections and pose a serious threat worldwide.


Staphylococcus aureus is a highly prevalent Gram-positive Across the globe, around 700 000 patients die every year due to
bacterial species and is also known as the “all-time champion” infections caused by drug-resistant strains.24 The death toll due
of microbial human pathogens.1 It is an opportunistic pathogen to antimicrobial resistance is estimated to rise to 10 million by
that causes a wide number of diseases such as minor skin 2050. According to recent statistical data from the U.S. Centers
infections, soft tissue infections, pneumonia, endocarditis, for Disease Control and Prevention (CDC), in the United States
osteomyelitis, and blood infections.2−4 Infections caused by S. alone every year more than 2.8 million people suffer from
aureus are typically acute and aggressive by nature and also top infections caused by drug-resistant bacteria, and 35 000 of them
the list of both hospital and community-associated infections. die.25 As per the CDC data, out of 323 700 estimated cases of
The β-lactam antibiotics are the most successful clinically infections caused by methicillin-resistant S. aureus (MRSA) in
effective antibiotics used to treat bacterial infections in humans.5 hospitalized patients, an estimated annual death rate of 10 600
Their wide therapeutic success is largely due to their broad- has been reported in 2017. Thus, there is an urgent need
spectrum activity, oral availability, excellent pharmacokinetics, worldwide to develop new and more effective anti-infective
safety, and bactericidal action. They have been the drugs of agents against S. aureus.26
choice for the treatment of S. aureus-induced infections. The development of new antibiotics is at present the most
However, these antibiotics are often seen as ineffective due to promising option available for treating bacterial infections.
the development of antibiotic-resistant strains.6−12 They However, due to the emergence of antibiotic resistance, anti-
prevent bacterial infections by targeting bacterial factors which virulence agents have gained widespread interest as a promising
either inhibit bacterial growth or result in bacterial killing.13,14 method for combating bacterial infections (Figure 1).27−30 This
These bacteriostatic and bactericidal properties of antibiotics shift to anti-virulence agents is due to an increasing prevalence of
along with their extensive use and abuse have led to the rapid drug-resistant strains and a shortage of new antibiotics for
emergence of drug-resistant strains, causing global concern. The clinical use.
key progress toward the development of antibiotics and the Virulence factors are the secretory, membrane-associated, or
emergence of antibiotic-resistant strains is outlined in Figure cytosolic molecules that bacteria use to infect hosts. They play a
1.15−20
S. aureus has been included as one of the six high-priority
threatening pathogens collectively referred to as the “ESKAPE” Received: March 3, 2021
pathogens (Enterococcus faecium, Staphylococcus aureus, Klebsiel- Published: September 13, 2021
la pneumoniae, Acinetobacter baumanni, Pseudomonas aeruginosa,
and Enterobacter).15,21−23 These ESKAPE pathogens can evade
the action of antibiotics, resulting in morbidity and mortality.
They are among the most common opportunistic pathogens in

© 2021 American Chemical Society https://doi.org/10.1021/acs.jmedchem.1c00386


13097 J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

Figure 1. Timeline showing milestones in the development of drugs against S. aureus and emergence of drug-resistant strains.

key role in the establishment of bacterial infections leading to 2. S. AUREUS SORTASE A: AN IDEAL ANTI-VIRULENCE
bacterial colonization, invasion of the host, adhesion, biofilm TARGET
formation, and evasion or suppression of the host’s immune Host cell adhesion is a crucial step of virulence in both Gram-
system. The anti-infective agents targeting the virulence factors positive and Gram-negative bacteria.45−48 S. aureus and other
affect the bacterial cell’s ability for host recognition, thereby Gram-positive bacteria produce a wide array of surface proteins
preventing the infection. In contrast to the drugs which inhibit that mediate the attachment of bacterial cells to host
cell growth or directly kill the bacteria, anti-virulence agents extracellular matrices such as fibrinogen, fibronectin, and
disarm the pathogen rather than destroying it.27,31 Therefore, it collagen, leading to adhesion.46,49,50 These surface proteins are
was believed that bacteria would generate weakened resistance referred to as microbial surface components recognizing
to such molecules compared to traditional antibiotics. However, adhesive matrix molecules (MSCRAMMs). These
there are reports on the development of resistance to anti- MSCRAMMs play a critical role in bacterial colonization, host
virulence drugs.32−34 invasion, and evasion of host’s immune system. These
The bacterial surface proteins involved in the recognition of extracellular proteins are synthesized in the cytoplasm and
hosts’ extracellular matrix proteins are anchored to the S. aureus displayed on the bacterial surface. Owing to the crucial role
cell wall via a transpeptidation reaction catalyzed by sortase A played by these surface proteins in the infectious process, anti-
infective therapies that impair bacterial adhesion have emerged
(SrtA).35−44 SrtA plays a critical role in the bacterial patho-
as ideal strategies for the treatment of S. aureus-induced
genesis; however, it is not vital for the growth or survival of S.
infections.46
aureus.38−40 SrtA-targeting anti-infective agents interfere with or The covalent anchoring of surface proteins on the cell wall of
destroy the cell wall anchoring pathway and restrain the S. aureus is catalyzed by a membrane-bound transpeptidase
virulence of pathogenic bacterial strains without affecting enzyme, namely sortase.36,37,51 The activity of the sortase
bacterial cell viability. This would lead to a reduced possibility enzyme was discovered by Schneewind and co-workers in the
of development of drug resistance; hence, SrtA is considered as a early 1990s (Figure 1).52−55 The purification and character-
potential anti-virulence therapeutic target to combat S. aureus- ization of SrtA from S. aureus by Schneewind in 1999 marked a
caused bacterial infections. Moreover, the enzyme’s extracellular new direction in chemical and biological research.56 Since then,
location and the absence of SrtA homologs in humans render the many other sortase isoforms including sortase B, C, D, E, and F
drugs with anti-SrtA activity specific to bacterial cells with low have been discovered from several other Gram-positive
off-target effects. bacteria.57−61
An excellent perspective article on this topic by Cascioferro Among the various isoforms, SrtA is the most studied sortase,
and co-workers summarized important developments in SrtA as it plays a key role in the pathogenesis of Gram-positive
inhibitors up to 2015.42 A recent review by Paek and co-workers bacteria. It is referred to as “housekeeping” sortase as it is present
describes small synthetic molecules with SrtA inhibition ability, in the majority of Gram-positive bacteria and is involved in the
chemical ligation of a variety of cell-surface proteins. SrtA
primarily focusing on their structural and synthetic aspects.44
recognizes and covalently anchors surface proteins containing
This Perspective provides a comprehensive treatment of S. an N-terminal LPXTG motif (X = any amino acid) to the
aureus SrtA inhibitors, with emphasis on recent developments. bacterial cell wall. It is known to anchor about 20 S. aureus
In the first part, the potential of SrtA as an anti-virulence target, surface proteins, including fibronectin-binding proteins (FnbA
its mechanism of action, and its structure are briefly discussed, and FnbB), clumping factors (ClfA and ClfB), serine-aspartate
followed by a comprehensive discussion on natural and synthetic repeat proteins (SdrC, SdrD, and SdrE), Staphylococcal protein
inhibitors of SrtA. In addition to these, a section dealing with A (Spa), and iron-regulated surface determinant A (IsdA), onto
commonly used assays to determine SrtA inhibition is also the bacterial surface.61−64 Their cell wall anchoring is essential
included. for the bacteria to exert pathogenic effect. S. aureus mutants
13098 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

Figure 2. SrtA-mediated cell wall anchoring pathway of surface proteins. This involves a surface protein recognition and thioesterification (a), further
nucleophilic attack by pentaglycine of lipid II to generate a lipid II-linked surface protein (b). The surface protein is incorporated into cell wall by
transglycosylation (c(i)) and transpeptidation (c(ii)). Abbreviations: Sec, secretory pathway; LPXTG, leucine, proline, any amino acid, threonine, and
glycine; GN, N-acetylglucosamine; MN, N-acetylmuramic acid; PP, undecaprenyl pyrophosphate; (Gly)5, pentaglycine; iGln, isoglutamine.

lacking SrtA cause virulence attenuation by the defective display Similarly, loss in bacterial adhesion was observed upon
of surface proteins.65,66 inactivation of SrtA for other Gram-positive bacteria, including
SrtA plays an important role in the formation of biofilm; Listeria monocytogenes, Streptococcus pneumoniae, Streptococcus
hence, SrtA inhibitors reduce or destroy the biofilm formation mutans, and Streptococcus suis.75−79 Additionally, SrtA is not
ability of S. aureus. Biofilms are organized microbial assemblies required for bacterial viability, which would exert less pressure
encased in an extracellular matrix that are used by bacteria to on bacteria for developing drug resistance. The extracellular
adhere to biotic or abiotic surfaces.67,68 They confer a favorable location and absence of any eukaryotic homologs make SrtA
growth environment to the bacteria and shield it from targeting easier and selective. Based on these advantages, SrtA is
antibiotics, disinfectants, preservatives, or host immune systems. considered an ideal target for developing anti-virulence drugs.41
These also contribute to the evolution of bacteria with drug- The mechanism of SrtA-mediated surface protein anchoring
resistant strains by promoting the transfer of genes associated in S. aureus is well established (Figure 2).54,61,80 The bacterial
with resistance.69−71 surface proteins are synthesized in the cytoplasm and consist of
Initial attachment, cell aggregation, biofilm maturation, and an N-terminal globular domain and a C-terminal cell wall sorting
signal. The cell wall sorting signal comprises a conserved
dispersal are the steps involved in S. aureus biofilm formation.72
LPXTG motif, a hydrophobic domain, and a tail of positively
S. aureus surface proteins, including fibronectin-binding proteins
charged residues.52 The LPXTG sorting sequence is secreted
(Fnbps), mediate bacterial attachment to fibrinogen, fibronec-
from the cytoplasm, along with the N-terminal globular domain,
tin, and other hosts’ extracellular matrix proteins. These proteins through the secretory (Sec) pathway. The hydrophobic domain
are anchored to the S. aureus cell wall through a SrtA-catalyzed remains embedded in the bacterial cell membrane, with the tail
transpeptidation reaction. Thus, anti-virulence agents that target retained in the cytoplasm. The LPXTG motif of the protein is
SrtA are likely to suppress the biofilm formation ability of S. recognized by SrtA, followed by nucleophilic attack of the
aureus. Studies have revealed that overexpression of SrtA in enzyme’s Cys184 on the carbonyl of Thr-Gly peptide bond.56
bacteria leads to an increase in the level of biofilm formation, This results in the cleavage of the Thr-Gly peptide bond,
whereas the SrtA knockout S. aureus strains are devoid of biofilm generating a thioester acyl-enzyme intermediate (Figure 2). The
formation ability.73,74 Reduction in the biofilm formation ability thioester bond between SrtA and surface proteins is
of various MRSA strains was observed as a result of mutations in subsequently attacked by the free amine group of the
Fnbps. Therefore, SrtA inhibitors should reduce the binding of pentaglycine unit within the lipid II precursor (undecaprenyl-
the bacteria to the host cell-matrix proteins, reducing infection. pyrophosphoryl-N-acetylmuramoyl-(L-Ala-D-iGln-Nε(NH2-
13099 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

Scheme 1. Schematic Representation of the Reverse Protonation Model for SrtA Catalysis

Gly5)-L-Lys-D-Ala-D-Ala)-β(1 → 4)-N-acetylglucosamine),
regenerating the active-site Cys184.61,81,82 Finally, the surface
proteins are incorporated into the peptidoglycan via trans-
glycosylation and transpeptidation reactions involved in
bacterial cell wall synthesis.83 This leads to covalent attachment
of the proteins to the cell wall peptidoglycan, to become true
surface proteins.
SrtA contains His120, Cys184, and Arg197 residues that are
found to be essential for the enzyme’s bioactivity.56,84−89
McCafferty and co-workers studied the catalytic mechanism of
SrtA based on mutational analysis, protein stability, isotope
effects, pH rate profile, and kinetic parameters.87 The Figure 3. (a) NMR structure of SrtAΔN59 from S. aureus. The β-strands
mechanism, namely, the reverse protonation model (Scheme and helices are colored blue and green, respectively. Adapted with
1), proposes that the active form of SrtA exists in a reverse permission from ref 84. Copyright 2001 National Academy of Sciences,
protonated state and represents a small fraction (0.06%) of the USA. (b) NMR structure of covalent SrtAΔN59-LPAT complex. Image
from the RCSB PDB ID: 2KID.85
total enzyme.87 These studies are in agreement with a Cys184-
His120 catalytic dyad and with Arg197 stabilizing the
tetrahedral intermediate. separated by another hairpin turn, followed by a long loop for
The crystal and NMR structures of SrtA in free and bound antiparallel pairing. The sulfhydryl group-containing Cys184 is
states have been successfully elaborated in the litera- located toward the end of the β7 strand, followed by another
ture.84−86,90S. aureus SrtA is a 206-amino acid long protein, loop leading to the β8 strand. The β7 and β8 strands form the
comprising unstructured and structurally ordered regions. The base of a hydrophobic pocket, whereas the amino acid residues
unordered domain, containing nonpolar amino acid residues, is connecting β3-β4, β2-β3, β6-β7, and β7-β8 strands constitute its
responsible for embedding the enzyme in the bacterial cell wall.
membrane, whereas the ordered domain is responsible for the The X-ray crystal structure of SrtAΔN59 was reported by
SrtA-catalyzed transpeptidation. Clubb and co-workers first Narayana and co-workers.90 The solid state structure revealed
reported the structure of SrtAΔN59 using NMR spectroscopy.84 that the unit cell comprises three molecules with different
SrtAΔN59 is a SrtA mutant without 59 residues from the enzyme’s conformations of amino acid residues in the connecting loop of
N-terminal. The recombinant SrtAΔN59 is fully capable of β6 and β7 strands. In contrast to the NMR structure, the crystal
catalyzing in vitro cleavage of peptides bearing the LPXTG motif structure revealed that one α and two 310 helices connect the
and the transpeptidation reaction. The NMR studies revealed eight β strands, giving rise to the β-barrel structure. In addition
that SrtA primarily possesses a unique 8-stranded β-barrel fold to this, conformational differences in the amino acid side chains
(β1−β8) including two short helices and several loops (Figure were also observed.
3a). The structure indicated that the β1 strand lies antiparallel to Experimental determination of the underlying molecular
the β2 strand, separated by a short hairpin structure. A 310 helix interactions between LPXTG-containing substrates and SrtA is
(H1, Figure 3a) tethers the β3 strand to the β2 strand in a difficult due to the transient existence of the Thr-Gly bond.
parallel orientation. The β3, β4, and β5 strands are positioned Clubb and co-workers synthesized an N-terminal carbobenzy-
antiparallel to each other, with β4 and β5 strands connected by a loxy-protected tetrapeptide, LPAT, with the carbonyl group of
long loop and α-helix (H2, Figure 3a). The β6 and β7 strands are Thr replaced by −CH2SH.85 The LPAT peptide covalently
13100 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

binds to the Cys184 of SrtA via a disulfide bond. This could biochemical assays are then conducted on hit candidates
structurally mimic the thio-acyl intermediate, formed by the obtained from screening. These early-stage measurements,
covalent attachment of LPXTG-containing surface protein to including the determination of half-maximal inhibitory concen-
SrtA. The results gathered from the NMR screening agree with tration (IC50) and minimum inhibitory concentration (MIC),
mutagenesis studies. The studies indicated that the recognition help in analyzing the compound’s bioactivity.
of surface protein involves accommodation of the LPXTG motif The nature of data on the biological activity of a compound is
in a large groove that leads into the enzyme’s active site. The dependent on the type of the biological assay performed. Many
base of the binding groove is constituted by β4 and β7 strands, conventional assays such as fluorescence resonance energy
while the amino acids comprising loops connecting β6/β7, β7/ transfer (FRET),37,56,62,85,102−107 high-performance liquid
β8, β3/β4, and β2/H1 form its walls (Figure 3b). chromatography (HPLC),92,108,109 cell adhesion,110,111 clump-
Particularly, the amino acids constituting the loop connecting ing,66 sodium dodecyl sulfate−polyacrylamide gel electro-
β6 and β7 strands play a key role in catalysis, which is phoresis (SDS-PAGE),112,113 and fluorescein incorporation
demonstrated by a significant loss in SrtA bioactivity upon assay,114 have been employed and reported in the literature to
mutations of these residues.85,89,91 Upon substrate binding, the assess the SrtA inhibitory activity of inhibitors. It is useful to
loop undergoes extensive structural changes and adopts a closed briefly survey these different techniques since it benefits the
conformation containing a 310 helix. The loop closure upon medicinal chemistry community.
substrate binding facilitates hydrophobic contacts of SrtA 3.1. Fluorescence Method. The anti-SrtA activity of the
residues with the leucine residue in the LPAT peptide. The test compound is most commonly determined using fluorogenic
proline residue is in contact with the nonpolar side chains of peptide cleavage assay. Kinetic parameters like catalytic rate
Ala92, Ala104, Ala118, Leu169, and Ile182 and buries the constants, catalytic efficiency, apparent rate constant, and IC50
peptide in the active-site binding groove. The covalently bound values are evaluated by monitoring the fluorescence enhance-
LPAT peptide adopts an L-shaped kinky structure with the Ala- ment of self-quenched fluorogenic substrate.56,103 The substrate
Pro peptide bond in the trans conformation. Mutations of amino consists of a fluorophore−quencher FRET pair, EDANS (5-[(2-
acids including Leu97, Ala104, Ala118, Val168, and Leu169 aminoethyl)amino]naphthalene-1-sulfonic acid) and Dabcyl (4-
resulted in significant losses of SrtA transpeptidation activ- ([4′-(dimethylamino)phenyl]azo)benzoic acid), suitably placed
ity.85,89,91 at the ends of a peptide sequence containing the SrtA-sensitive
The NMR studies on SrtA and SrtA−substrate complex LPXTG motif.37,56,62,105 Peptide substrate Abz-LPETG-Dap-
suggest that SrtA binds to a single calcium ion through side (DNP)-NH2 is also used in FRET assay to evaluate SrtA activity
chains of Glu and Asp residues within β3-β4 and β6-β7 with 2-aminobenzoic acid (Abz) as the fluorophore.85,103 The
loops.84,85 It is further indicated that the electrostatic interaction quencher, 2,4-dinitrophenylaniline (DNP), is linked to the
between Glu171 and calcium ion upon substrate-binding leads peptide sequence using 2,3-diaminopropionic acid (Dap). SrtA
to structural and mobility changes in the β6-β7 loop. The inhibition by fluorescence measurements can also be deter-
binding to calcium ion near the active site stimulates the catalytic mined by using commercially available assay kits.104,105 In the
activity of SrtA by 8-fold.84 absence of SrtA, the fluorophore and quencher are held closely,
Molecular dynamics (MD) simulations revealed multiple leading to fluorescence quenching. Upon incubation of a peptide
conformations for SrtA and LPATG sorting signals.92,93 Because with SrtA, Thr-Gly in the peptide sequence undergoes
of the highly dynamic nature of SrtA, it is also suspected that the enzymatic cleavage (Figure 4a). The peptide segment attached
catalytic site is not necessarily the binding site for at least some of to SrtA via a thioester bond can undergo hydrolysis (Figure
the inhibitors. Therefore, given the dynamic nature of this non- 4b).56 This results in the formation of two peptides, with one
covalent association, it is not surprising that very few
bearing the fluorophore and the other with the quencher. The
compounds with strong SrtA inhibition have been reported.
separation of fluorophore and quencher restores the fluores-
The catalytic mechanism and structural information gathered
on SrtA and its complex have been successfully utilized for
protein engineering and synthesizing designer macromole-
cules.94−101 This has also served as a basis for designing SrtA-
targeting anti-infective agents that reduce bacterial virulence.

3. EXPERIMENTAL METHODS FOR EVALUATING


SORTASE A INHIBITORS
For more than a decade now, research efforts have focused on
strategies to discover drugs that target bacterial virulence factors
that are not essential for cellular viability/survival. Since then, a
large number of natural or synthetic SrtA inhibitors have been
identified using diverse strategies. Several literature reports on
SrtA inhibitors are based on the identification of hit candidates
through high-throughput screening (HTS) of large compound
libraries. HTS allows the identification of molecules with drug- Figure 4. (a) Schematic representation of FRET-based enzymatic assay
using Dabcyl-EDANS appended peptide substrate. (b) Mechanism of
like properties among several thousands of compounds SrtA-catalyzed reactions: (i) initial cleavage of LPXTG substrate by
screened. SrtA, generating a thioester acyl−enzyme intermediate, and nucleo-
In combination with HTS, virtual screening of chemical philic attack on intermediate by (ii) water and (iii) pentaglycine unit as
libraries against the structure of SrtA is another popularly nucleophilic substrate. Abbreviations: Dabcyl, 4-([4′-
followed method for the identification of suitable drug (dimethylamino)phenyl]azo)benzoic acid; EDANS, 5-[(2-
candidates. Preliminary experimental studies using various aminoethyl)amino]naphthalene-1-sulfonic acid.

13101 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

cence of the donor molecule. The increase in fluorescence intermediates and products. In particular, the kinetic parameters
intensity will be proportional to the amount of peptide which has for the transpeptidation step can be accurately quantified.
undergone cleavage. Treatment of SrtA with a test compound Additionally, the identification of individual intermediates and
(inhibitor) would reduce the enzymatic cleavage of peptides, products might also be useful to decipher the mechanism of SrtA
consequently resulting in lower fluorescence intensity. The inhibition.109 Although the HPLC-based assay is more time-
experimental protocol can be validated by performing blank consuming, usage of fast-flow HPLC columns might enable
measurements including positive and negative controls. high-throughput applications of this assay.
Inhibitors including phenyl vinyl sulfone, 1-(3,4-dichlorophen- 3.3. Cell Adhesion and Clumping Methods. In addition
yl)-3-(dimethylamino)propan-1-one, and p-hydroxymercuri- to FRET and HPLC assays already described, the SrtA
benzoic acid with known SrtA-inhibition ability are used as inhibition potency of test compounds can also be evaluated
positive controls, whereas the bacterial solutions in the absence using cell adhesion assay.110,111 Bacterial attachment to host
of inhibitor or with SrtA knockout strains of S. aureus are used as tissues occurs via cell wall anchored surface proteins like
negative controls.56 fibronectin- and fibrinogen-binding proteins.72 These proteins
The FRET-based enzymatic assay is a quick and effective are displayed on the bacterial surface via the transpeptidation
method for coarse/gross analysis of the anti-SrtA activity of test reaction catalyzed by SrtA.37,38 Thus, SrtA inhibition reduces
compounds. It is the commonly used technique, since it is well the surface display of the fibronectin- and fibrinogen-binding
suited for the HTS approach. However, it can only provide proteins on bacterial cell walls, which in turn would hinder S.
information on the SrtA-catalyzed initial cleavage of LPXTG- aureus attachment to host tissues. Therefore, the cell adhesion
containing peptides. The kinetic parameters associated with the assay can also be used to evaluate SrtA inhibition activity. This
transpeptidation step (Figure 4b) cannot be quantified, since the assay involves monitoring the binding of bacterial cells to
fluorophore has already regained its fluorescence in the first step. fibronectin- or fibrinogen-coated surfaces. As the first step,
Thus, no additional nucleophilic substrate is generally used to overnight-cultured bacterial cells are inoculated and incubated
resolve the thioester acyl−enzyme intermediate generated in the with different concentrations of inhibitors under appropriate
first step. However, the intermediate can undergo hydrolysis due conditions for cellular growth. The cellular growth in the
to the presence of water molecules.56 Since the assay relies on incubation medium is monitored by determining its optical
fluorescence measurements, it is likely that fluorescence density (OD), generally measured at 600 nm. After attaining a
quenching interferes with the results due to the inner filter effect. suitable population of bacteria, the cells are pelleted by
Recently, more sophisticated fluorescence-based assays have centrifugation, washed, and resuspended in phosphate-buffered
been developed to measure the SrtA inhibition potency of a test saline (PBS). The suspension is added and incubated in a well
compound. The assay designed by Zhang and co-workers plate coated with bovine fibrinogen or human plasma
exploited the use of FRET between cyanine dye−peptide fibronectin. The well plate after washing and cell fixation using
conjugates, Cy3-CLPETGG and GGGG-Cy5.106 In contrast to glutaraldehyde is stained with crystal violet. After washing and
Dabcyl-EDANS and Abz-DNP-based peptide cleavage assays, subsequent drying, the absorbance of the well plates at 570 nm is
this assay relies on SrtA-catalyzed chemical ligation of peptides recorded. In the absence of test compound (inhibitor), more
bringing donor (Cy3) and acceptor (Cy5) in proximity to each immobilization of crystal-violet-stained S. aureus on the well
other. This results in enhanced fluorescence emission of Cy5 plate is expected, resulting in high absorbance. However, after
which can be quantified by fluorescence imaging. The incubating bacterial cells with SrtA inhibitor, decreased enzyme
measurements in this assay are less likely to be affected by the activity would lead to low absorbance, as the tendency of S.
inner filter effect over other fluorescence-based assays. A aureus to adhere to protein-coated well plates is reduced. The
fluorescence-based assay to determine SrtA inhibition efficacy, test results are validated by performing blank and control
based on the increase in fluorescence intensity of an externally experiments under identical conditions.
added biarsenical dye, has also been developed.107 The assay is Cell clumping assay is another method to determine the
relatively simple and 24.7-fold more sensitive than peptide potency of a compound to inhibit SrtA activity.66 Similar to the
cleavage assays. It can be successfully applied for identifying new cell adhesion assay, culturing the bacterial cells and subsequent
candidates with potent SrtA deactivation ability. incubation with test compound, followed by centrifugation are
3.2. HPLC Method. To address discrepancies associated the steps involved in the cell clumping assay. The pelleted cells
with the fluorescence intensity attenuation method, an HPLC- are suspended and incubated in a fibrinogen solution. Unlike the
based assay has been devised to analyze SrtA-inhibition cell adhesion assay, immobilization of bacterial cells to
activity.108,109 This protocol utilizes the same peptide substrate, fibrinogen is analyzed by measuring the OD of the suspension
Abz-LPETG-Dap(DNP)-NH2, as in the FRET assay, and a at 600 nm. In the absence of an inhibitor, SrtA-anchored surface
nucleophilic substrate, like pentaglycine or triglycine for the proteins bind to fibrinogen in solution, forming visible
transpeptidation reaction.92,108,109 The Abz and DNP motifs aggregates which settle down with time, resulting in a lower
facilitate UV/fluorescence detection of the unreacted peptide OD value. In the presence of SrtA inhibitors, cells do not
substrate, or acylation and transpeptidation products. After aggregate upon incubation with fibrinogen and remain
incubation of SrtA with the two substrates in the presence or suspended in the solution.
absence of the inhibitor, the reaction mixture is purified by 3.4. SDS-PAGE Assay. The inhibitory activities of SrtA
centrifugation to remove the enzyme. The mixture is subjected inhibitors can also be determined using SDS-PAGE assay.112,113
to HPLC analysis, and the area under the peaks corresponding The method involves the separation of proteins from complex
to starting materials, intermediates, and product formed is used protein mixtures and their complexes, based on molecular
to quantify the percentage conversion.108,109 Additionally, the weight, under the influence of an applied electric field. The
HPLC data can be verified by MALDI-TOF mass spectrometric analysis of SrtA inhibition by this method involves the
analysis. Unlike FRET-based enzymatic detection, the HPLC incubation of the candidate molecule with SrtA, S. aureus
assay enables monitoring the concentrations of reaction surface protein (IsdA), and nucleophilic substrate (triglycine),
13102 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

followed by quenching of the reaction. The reaction mixture is In 2002, two separate research teams reported inhibition of S.
then subjected to SDS-PAGE. After electrophoresis, the protein aureus SrtA based on synthetic and natural sources.116,117 A
components in the reaction mixture are separated, and the multitude of publications have been published since then
intensities of the bands are analyzed to quantify the amount of regarding the SrtA inhibitors. Although SrtA inhibition studies
starting materials and transpeptidation products. The experi- seem to indicate that all three aforementioned classes of
ments are validated by performing blanks in the absence of test molecules have great potential, many of them have only
compound, SrtA, or S. aureus protein and by control experiments demonstrated in vitro anti-SrtA activity, with very few of them
using known SrtA inhibitors. exhibiting strong inhibitory properties. Consequently, to date,
3.5. Fluorescein Incorporation Assay. Another fluores- there remains a paucity of clinically relevant candidates.
cence-based method used for quantification of SrtA inhibition A critical look at the inhibitors reveals that only a handful of
activity of an unknown compound involves the incorporation of them belong to peptidomimetics, though enormous design
fluorescently labeled LPXTG-containing peptides into the S. possibility exists. The same is true for natural product-based
aureus cell wall.114 The peptide is ligated to the bacterial cell wall sortase inhibitors. Screening of a limited number of marine and
by a SrtA-mediated transpeptidation reaction. This assay plant-based natural products identified potent SrtA inhibitors,
requires culturing S. aureus in a suitable growth medium with while activities of plenty of natural products are unknown.
different concentrations of a test compound in the presence of Synthetic small molecules are a treasure trove of inhibitors. The
peptide bearing the fluorophore (5(6)-carboxyfluorescein). enormous chemical space available for design and the power of
After a suitable incubation time, cells are pelleted and washed combinatorial and diversity-oriented approaches are not fully
with cold PBS. The cells are subsequently washed with 5% exploited. The involvement of talented and creative medicinal
sodium dodecyl sulfate to remove non-covalently bound chemists is urgently needed.
molecules. The cells are washed twice with cold PBS, Although HTS is used for the efficient discovery of synthetic
resuspended, and subjected to fluorescence measurements. small-molecule SrtA inhibitors, leading to the identification of
Analysis of the fluorescence intensity in the absence of the test many hits, some of them may be falsely positive.118,119
compound or using a S. aureus SrtA knockout strain serves as a Compounds that are reactive to proteins, aggregates, or that
control to validate the experimental results. In the absence of the interfere with the assay are likely to show false-positive results.
test compound, the extent of inclusion of the fluorescently Compounds such as rhodanines, hydroxyphenylhydrazones,
labeled peptide into the S. aureus cell wall is indicated by strong catechols, benzofurazans, quinones, pyrroles, and thiophenes are
emission from the fluorophore. Incubation of cells in the some of the Pan Assay INterference compoundS (PAINS).120
presence of a SrtA inhibitor is expected to show low fluorescence Such problematic classes of compounds have also been reported
intensity due to inhibition of SrtA. as sortase inhibitors. The removal of such compounds using a
A rotor fluorogenic D-amino acid (RfDAA)-based inclusion PAINS filter is necessary before further optimization. It is further
assay recently developed by Hsu et al. has been successfully possible that some of the real inhibitors might also be
applied to track the transpeptidation reaction involved in inappropriately classified as nuisance compounds.121 Therefore,
peptidoglycan biosynthesis.115 RfDAAs are terminal D-lysine- an electronic filter to remove PAINS, and then a detailed
linked fluorescent probes which show fluorescence turn-on experimental follow-up is necessary.
response upon integration into the bacterial cell wall. The assay Although the exact mechanism of SrtA inhibition remains
has also been demonstrated to track SrtA-catalyzed trans- uncertain for many compounds, hypothetical mechanisms based
peptidation reactions and could facilitate HTS of SrtA-targeting on docking and experimental data have been proposed. Broadly,
inhibitors. the inhibitors are found to inactivate SrtA through either
covalent attachment or non-covalent binding to the enzyme’s
4. INHIBITORS OF SORTASE A active site (Figure 5). The non-covalent inhibitors can bind the
Infectious diseases are re-emerging at an alarming rate globally enzyme’s active site through hydrogen bonding, van der Waals,
due to bacterial resistance against conventional antibiotics. To electrostatic, and π−π interactions, while covalent inhibitors
win the battle against infectious diseases, our strategy for react with the sulfhydryl group of Cys184 present at the active
combating bacteria needs to change. In this context, attacking the site of the enzyme. Covalent inhibitors can thus irreversibly bind
bacterial virulence target in the host cell is perhaps the most the active site of the enzyme, thereby blocking the LPXTG-
effective option.31 We have a reasonably good structural and containing bacterial proteins from binding SrtA. This con-
biological understanding of the SrtA enzyme which is involved sequently leads to the failure of bacteria to present surface
in the virulence of S. aureus. Molecules that target bacterial proteins, leading to virulence attenuation. The following
virulence would avoid the emergence of drug-resistant strains. sections outline various literature reports focused on the
The emergence of multidrug-resistant organisms is having a identification and development of SrtA inhibitors.
huge impact on the human health sector and is the driving force 4.1. Peptidomimetics as Sortase A Inhibitors. A large
for the development of anti-virulence agents as therapeutics. number of bioactive peptides have been discovered and show
In this Perspective article, we have analyzed all the sortase great promise as drug candidates. While peptide-based
inhibitors reported in the literature and classified them into molecular structures have several advantages from the point of
three general classes: peptidomimetics, natural product-based, view of “molecular recognition”, they also have some inherent
and synthetic small molecules. Further, we have selected those limitations. These limitations are primarily in terms of their
inhibitors that showed better inhibitory activities, and they are stability toward proteolysis, poor selectivity due to inherent
presented in tabular forms under each class. This will provide flexibility being a short peptide, and rapid clearance from the
both an extensive and a quick overview of each class, allowing body, hindering pharmaceutical applications. Therefore, pepti-
medicinal chemists to further fine-tune their structural intuition domimetic compounds have been developed to address these
to generate more potent inhibitors. shortcomings.
13103 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

order rate constant of inhibition (k2nd, ki/Ki) of (5.3 ± 0.6) × 104


M−1 min−1.
Similarly, an N-protected LPAT-containing peptidomimetic
has been designed with vinyl sulfone at the C-terminus (1c,
Supporting Information (SI) Table S1).124 Compound 1c
exhibited irreversible inhibition of SrtA by covalent attachment
through the vinyl sulfone group of 1c to the SrtA active-site thiol.
Its enzyme inhibition efficiency was quantified by evaluating the
kinetic constants Ki (9 μM), ki (4 × 10−4 min−1), and k2nd (44.4
M−1 min−1). The lower ki and k2nd of 1c over those of 1a and 1b
indicate slow binding to SrtA, which is in agreement with the low
electrophilicity of the vinyl sulfone moiety.
4.1.2. Phosphinic Peptides as SrtA Inhibitors. To get insight
into the mechanism by which the substrate-derived inhibitors
function, McCafferty and co-workers designed compound 2 (SI
Table S1), with the scissile Thr-Gly bond replaced by a non-
hydrolyzable phosphorus−carbon bond.125 Preliminary inhib-
itory analysis of 2 indicated a reversible and competitive SrtA
inhibition, with IC50 = 10 mM. The poor SrtA inhibition efficacy
of 2 indicated the need for an accurate design for the
development of more potent inhibitors.
4.1.3. Leu-Pro-Ala Tripeptide Derivative. Jung et al. reported
an N-protected Leu-Pro-Ala tripeptide linked to 3-amino-4-
mercapto-2-butanol via an amide linkage exhibiting SrtA
inhibition.126 Incubation of SrtA in pH 8 buffer with 5 M of
the tripeptide followed by reverse-phase HPLC analysis
indicated covalent modification of SrtA via the formation of a
disulfide bond involving the thiol group of the tripeptide
inhibitor and active-site Cys184.
4.1.4. Peptide Macrocycles. Driven by the high selectivity,
binding affinity, and enhanced molecular interactions of bicyclic
peptides, Heinis and co-workers screened bicyclic peptides for
Figure 5. General schematic representation of SrtA-catalyzed trans- SrtA inhibitory activity.114 These bicyclic peptides showed
peptidation reaction and SrtA inhibition. substantially better SrtA inhibition than the many potent small-
molecule inhibitors. Phage display and HTS of peptides against
Peptidomimetics could overcome the limitations associated SrtA revealed that bicyclic peptide(s) containing the Leu-Pro-
with bioactive peptides but at the same time mimic the Pro motif, similar to the LPXTG motif, were more potent
conformation and other structural features of the native inhibitors of SrtA. In vitro inhibition studies performed on 13
peptides.122,123 Peptidomimetic inhibitors are developed based shortlisted bicyclic peptides showed 3a (ACPQLPPCRVSCG)
on the structures and conformations of the native peptide and 3b (ACTQRCPQLPPCG) (Table 1) as the potent
ligands; hence, they could bind the biological targets with high inhibitors, with inhibitory constants of 4.3 ± 0.3 and 2.4 ± 0.3
specificity. The strategy adopted to develop peptidomimetic μM, respectively. Substitution of amino acids in the second ring
inhibitors for SrtA involves designing structural mimics of the of 3a by alanine shows no reduction in the inhibitory activity,
LPXTG motif, keeping the necessary parts of the ligand for indicating that the second ring is not interacting with SrtA.
molecular recognition. The Thr-Gly peptide bond of the However, the substitution of Arg and Gln by Ala in 3b reduced
LPXTG motif is replaced by a stable and cleavage-resistant the activity, indicating that the amino acids of the first ring are
bond to block transpeptidation. Peptidomimetics are therefore involved in the interaction with SrtA. Bicyclic peptide 3b was
considered an excellent choice as SrtA inhibitors. shown to exhibit a high binding selectivity for SrtA over other
4.1.1. LPAT-Coupled Diazoketone, Chloromethyl Ketone, proteins possessing surface-accessible cysteine. The increased
and Vinyl Sulfone. Walker and co-workers synthesized activity of inhibition shown by bicyclic peptides
peptidomimetic inhibitors by replacing the scissile Thr-Gly ACTSRCPQLPPCG and ACHSRCPQLPPCG over 3b in-
bond of the natural substrate (LPXTG) of SrtA with dicates that inhibitors can be made more potent by optimizing
diazomethyl ketone (1a) and chloromethyl ketone (1b) groups the amino acids in the first ring. The treatment of S. aureus with
(Table 1).116 Both compounds showed an irreversible time- the bicyclic peptide ACHSRCPQLPPCG reduced the incorpo-
dependent inhibition, as a result of alkylation of active site ration of the fluorescein-labeled LPETG substrate into the
Cys184. Compounds 1a and 1b exhibited inhibitory constants bacterial cell wall, thus supporting SrtA’s inhibition activity.
(Ki) around 0.2 μM. Their identical values of Ki indicated that 4.1.5. Leu-Pro-Arg-Asp-Ala Peptide. Virtual screening of
both inhibitors have an equal affinity of interaction toward SrtA. various oligopeptides revealed that pentapeptide Leu-Pro-Arg-
However, it was observed that 1b exhibited ∼2-fold higher first- Asp-Ala (4, Table 1) exhibited SrtA inhibition with IC50 = 10.6
order rate constant of inactivation (ki) (1.1 ± 0.1 × 10−2 min−1) μM.127 Moreover, treatment of S. aureus with peptide 4 hindered
than 1a, which was attributed to the higher electrophilicity of the bacterial adhesion, biofilm formation, and bacterial invasion. In
chloromethyl group over the diazomethyl group. Thus, 1b vivo studies on Bagg Albino C (BALB/c) mice indicated that S.
showed greater effectiveness as a SrtA inhibitor, with a second- aureus-induced mastitis was effectively treated with 4. This
13104 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

Table 1. Chemical Structures of Selected Peptidomimetic SrtA Inhibitors and Their Binding Parameters

a
Second-order rate constant of inhibition/inactivation potency (M−1 min−1). bApparent dissociation constant (μM). cIC50(μM).

highlighted the potential of 4 as an effective anti-virulence agent. for biological activity; hence, many potential drugs are possible
To determine the structure of the enzyme−inhibitor complex, in the future. It is noteworthy that many natural products are still
molecular modeling studies were performed. The studies unidentified, and enormous opportunities for drug discovery are
revealed that 4 binds to the SrtA active site and interacts with still open.
Asn114, Ser116, Gln172, Trp194, and Arg197 residues (Figure Natural products are widely distributed in chemical space and
6). display enormous structural and chemical diversity; hence, they
Future Prospects and Outlook. Only a limited number of are a treasure chest for potential SrtA inhibitors. However, only a
peptides have been screened for SrtA inhibitory activity. Within few natural products have been investigated, many of which
this limited number of compounds screened, peptide 4 showed show good inhibitory effects. In the following section, the
promising activity; hence, efforts are needed in the design of natural products that have been investigated for SrtA inhibition
peptide-based SrtA inhibitors. Peptidomimetics are potential are discussed.
drug candidates; hence, more efforts in this direction are 4.2.1. Marine Natural Products. Oceans cover about 70% of
urgently needed for the discovery of potent inhibitors. The the Earth’s surface and provide diverse conditions such as low
progress toward the discovery of SrtA inhibitors depends on the oxygen and lack of light compared to the terrestrial environment.
combined efforts of organic and medicinal chemists. Well- Consequently, the physiological adaptation of organisms to
developed synthetic methodologies available today along with these conditions increases the likelihood of finding structurally
rational de novo design of inhibitors will provide hope for the unique natural products.129,130 In the past decade, a plethora of
discovery of potent SrtA inhibitors. marine natural products with new structures and diverse
4.2. Natural Products as Sortase A Inhibitors. Natural bioactivities have been discovered.131,132
products have been used as medicines throughout the human 4.2.1.1. Indole Alkaloids. Alkaloids containing indole units
history. This knowledge of natural products as medicines came are commonly found in marine invertebrates. The bis(indole)
as a result of hundreds of years of trial and error experiments. alkaloids 5a−5i (Table 2 and SI Table S2) were isolated from
Many natural products are current drug candidates.128 the bright yellow marine sponge Spongosorites sp., consisting of
Interestingly, less than 10% of natural products are evaluated two indole moieties connected via heterocyclic units.133 Such
13105 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

Figure 6. (a) Structure of SrtA−4 complex determined using docking studies. (b) Modeled structure showing interactions between 4 and SrtA. From
ref 127, CC BY 4.0.

heterocyclic compounds are abundant in marine sponges. The SrtA.138 Two sesterterpenes, 9a and 9b (Table 2), showed IC50
compounds 5a−5d belong to the topsentin class, while 5e−5i = 36 and 98 μM, respectively.
belong to the hamacanthin class. Among these compounds, 5c 4.2.1.6. β-Carboline Alkaloids. Six β-carboline alkaloids
showed the highest anti-SrtA activity, having IC50 = 39.6 ± 1.3 10a−10f (SI Table S2) isolated from marine Korean ascidian
μM. Synoicum sp. showed inhibitory activity toward SrtA.139 Several
4.2.1.2. Psammaplin. A previously characterized natural derivatives of these compounds also showed weak to moderate
compound, psammaplin A (6, Table 2), isolated from the SrtA inhibition.
sponge Aplysinella rhax showed SrtA inhibition with IC50 = 47.4 4.2.1.7. Cadiolides and Isocadiolides. Five tris-aromatic
μM and MIC > 200 μg/mL.134 Treatment of S. aureus with 6 furanones and four related bis-aromatic diesters were isolated
reduced the fibronectin-binding ability of bacterial cells. The from the ascidian Synoicum sp. collected from off the coast of
compound also exhibited moderate anti-SrtB activity. Chuja-do, South Korea.140 The structures of these compounds
4.2.1.3. Aaptamines. Four aaptamines isolated from the were determined by spectroscopic analysis. Many of these
marine sponge Aaptos aaptos showed strong inhibitory activity isolated compounds exhibited SrtA inhibition. Among them,
against S. aureus SrtA.135 Among them, isoaaptamine 7 (Table cadiolide 11 (Table 2) was found to be the most potent SrtA
inhibitor, with an IC50 value of 78.8 μM. However, it also
2) was found to be the most potent SrtA inhibitor, with IC50 =
displayed high antibacterial activity against S. aureus, with MIC =
16.2 ± 0.9 μM. Structure−activity relationship (SAR) studies
3.9 μM, which is not a desirable property of an anti-virulence
highlighted the crucial role of the methyl group at the N-1
agent.
position of 7 for SrtA inhibition. The MD-simulated structure of Recently, polybrominated aromatic compounds extracted
the SrtA-7 complex revealed hydrogen-bonding interactions of from Korean Synoicum sp. exhibited moderate SrtA inhibition
the hydroxyl and nitrogen of 7 with Glu105 and Lys196 of activity.141 Among them, isocadiolide A 12a and isocadiolide H
SrtA.136 Strong hydrophobic interactions of the methoxy and N- 12b (Table 2) moderately inhibited SrtA activity, with IC50 = 67
methylamine of 7 with the enzyme’s Pro91, Ala104, and Gly192 and 70 μM, respectively.
were also evident from the MD-simulated structure. 4.2.1.8. Naphthoquinone. Herqueidiketal 13 (Table 2) was
4.2.1.4. Discorhabdins. Twelve compounds, including isolated from marine-derived fungus Penicillium sp. The
(−)-3-dihydrodiscorhabdin 8a and (−)-discorhabdin 8b compound showed inhibitory activity against SrtA, with IC50 =
(Table 2), were isolated from the dark green sponge Sceptrella 23.6 μM and with moderate cytotoxicity.142
sp.137 The pyrroloiminoquinone alkaloids 8a and 8b showed 4.2.1.9. Lumichrome. Marine-derived Streptomyces sp. is a
SrtA inhibition with IC50 = 75.7 and 6.5 μM, respectively. storehouse of novel antibiotics. Lumichrome 14 (SI Table S2)
4.2.1.5. Halisulfate. Bioactive metabolites from marine was extracted from the Streptomyces sp. MBTH32 strain, which
sponge Coscinoderma sp. showed inhibitory activity against was isolated from marine sediment of Shinjin Island, Republic of
13106 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

Table 2. Chemical Structures of Selected Marine Natural Products as SrtA Inhibitors with Their IC50 Values

13107 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

Table 3. Chemical Structures of Selected Plant-Based Natural Products as SrtA Inhibitors with Their IC50 Values

13108 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

Table 3. continued

a
IC50 in μg/mL. bSecond-order rate constant of inactivation (M−1 min−1).

Korea.143 The compound exhibited SrtA inhibitory activity, IC50 proportion of plant species have been evaluated for the discovery
= 198.20 ± 0.94 μM, without affecting the viability of S. aureus. of bioactive compounds. Thus, the future exploration of plants
4.2.1.10. Polyketide. Recently, a series of alkaloids and one as a source of novel SrtA inhibitors may lead to more potent
polyketide were isolated from Aspergillus sp. F452. Among them, inhibitors or serve as pointers for developing highly active SrtA
the polyketide Aspermytin A 15 (SI Table S2) showed higher inhibitors.
SrtA inhibition with IC50 = 146 μM.144 The SrtA inhibition by 4.2.2.1. Isoquinoline Alkaloid. The powdered rhizome of
Aspermytin A resulted in the reduction of bacterial adhesion to Coptis chinensis was extracted, and it showed inhibitory activity
the fibronectin-coated surfaces. toward SrtA.147 Chloride and sulfate salts of berberine (16a and
4.2.2. Plant-Based Natural Products. Plant-based natural 16b, Table 3), palmatine chloride 17, and β-hydrastine 18 were
products continue to play an important role in the discovery of found to be the components that showed SrtA inhibition.
compounds with medicinal properties, therefore a cornerstone 4.2.2.2. Curcumin. Methanol extract from Curcuma longa L.
of drug discovery and development. Plants are found in every rhizome showed inhibitory activity toward SrtA. Chromato-
habitable environment and are the storehouse of a wide variety graphic separation of the components revealed the presence of
of natural products with diverse structures and high medicinal curcumin 19a, demethoxycurcumin 19b, and bisdemethoxy-
properties. A large proportion of the world’s population relies on curcumin 19c (Table 3 and SI Table S3).148 Curcumin showed
plant-based natural products for medicines.145,146 A very small higher potency for SrtA inhibition with IC50 = 37.5 ± 1.9 μM,
13109 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

Figure 7. MD-simulated structure of the 20b-Glc−SrtA complex. From ref 161, CC BY 4.0.

followed by 19b (IC50 = 70.3 ± 1.8 μM) and 19c (IC50 = 103.5 inhibition with IC50 = 71.8 ± 11.9 μM, with no influence on
± 3.9 μM). All these compounds showed MIC > 543.3 μM, bacterial cell viability. S. aureus Newman strain treated with 20b-
which indicates no effect on the growth of S. aureus. Molecular Glc revealed that it blocked adhesion of bacterial cells to
docking studies showed that these compounds bind to the fibronectin-coated plates, supporting the inhibitory activity. MD
catalytic region of SrtA, making contacts with Pro163, Gln172, simulations indicated that 20b-Glc binds directly to the active
Ile182, and Ile199.149 site of SrtA through multiple hydrogen bonding and hydro-
S. mutans is a Gram-positive bacterium associated with dental phobic interactions (Figure 7). The 4H-chromen-4-one moiety
cavities in humans.150 It is regarded as the major contributor of of 20b-Glc showed strong interactions with Ala104 and Ser106.
tooth decay and is known for its ability to form biofilm on dental In addition to these interactions, van der Waals interactions of
surfaces, referred to as dental plaque.151,152 Biofilm formation 20b-Glc with Val166, Gly167, Val168, and Val193 stabilized the
starts with the initial attachment of S. mutans to the tooth SrtA−inhibitor complex.
surface. The adhesion can be mediated by surface proteins, Based on the beneficial pharmacological properties of
including protein antigen c (PAc).153,154 The PAc protein is myricetin and its SrtA inhibition efficacy, Macedo and co-
covalently anchored to the bacterial cell wall by SrtA-mediated workers investigated the anti-biofilm formation potency of 20c
transpeptidation. Thus, targeting SrtA is a powerful approach for and its 3-O-glycoside derivative 20c-Glc (SI Table S3).162 The
the prevention of dental cavities.155 This is supported by the studies revealed that the inhibitor 20c-Glc has no influence on
decreased biofilm formation ability of SrtA knockout strains of S. bacterial cell growth but strongly inhibited its adhesion and
mutans consequently reducing caries formation.156 Curcumin biofilm formation. MD simulations suggested that the inhibitor
was found to inhibit SrtA with IC50 = 10.2 ± 0.2 μM, leading to 20c binds to SrtA as its natural substrate. The SrtA inhibition
reduced biofilm formation by S. mutans.77,157 ability of myricetin was also evaluated using a FRET-based
4.2.2.3. Flavonoids. Flavonoids constitute an important class enzymatic assay by Nitulescu and co-workers.104 The bioactivity
of natural products widely found in fruits, vegetables, and studies indicated strong SrtA-inhibition by 20c. Docking studies
beverages. These natural compounds having polyphenolic revealed that myricetin non-covalently binds to the enzyme’s
structures can be classified into flavones, flavonols, flavanones, active site and thereby hinders the interaction between SrtA and
isoflavones, and anthocyanidins. Flavonoids have been success- LPXTG-containing natural substrate (Figure 8).104 The studies
fully demonstrated to possess many biological and pharmaco- indicated the involvement of SrtA residues, including Ala104,
logical activities, including anti-cancer, anti-Alzheimer’s, anti- Glu105, Asn114, Ser116, Gly167, Val168, Thr180, Ile182,
inflammatory, and anti-oxidative effects.158 Several flavonoids Arg197, and Ile199, in stabilizing the SrtA−20c complex.
exhibiting SrtA inhibitory activity have also been reported. The search for more potent SrtA inhibitors led to the
Among the series of flavonoid-based inhibitors, a few promising identification of another series of flavonoids isolated from the
compounds are discussed below. roots of Sophora f lavescens. Among the tested compounds,
Oh and co-workers analyzed the SrtA inhibitory activity of a kurarinol 21 (SI Table S3) exhibited SrtA inhibitory activity,
series of naturally occurring flavonols and identified that morin with IC50 = 107.7 ± 6.6 μM.163 It was demonstrated to possess
20a, quercetin 20b, and myricetin 20c are active against weak antibacterial activity against S. aureus, with MIC = 219 μM.
recombinant SrtA with IC50 = 37.39 ± 3.14, 52.70 ± 1.56, and Among 50 natural compounds screened using FRET assay,
44.03 ± 3.52 μM, respectively (Table 3).159 These compounds acacetin 22 (SI Table S3) was demonstrated to significantly
exerted no growth inhibition effect on the S. aureus Newman inhibit S. aureus SrtA activity with IC50 = 128.4 ± 16.5 μM and
strain. Among these, compound 20a was found to be a more MIC > 3.60 mM.164 Further studies revealed that acacetin
potent SrtA inhibitor than 20b and 20c. Additionally, it also reduced the surface display of bacterial proteins such as Fnbps
inhibited SrtA from S. mutans with IC50 = 27.2 ± 2.6 μM and by inhibiting SrtA-catalyzed transpeptidation.
reduced the biofilm formation of S. mutans without affecting the The MD-simulated structure of the SrtA−22 complex (Figure
bacterial viability.160 9) suggested the placement of acacetin’s phenyl ring in the
Compound 20b-Glc (Table 3), a 3-O-glycoside derivative of hydrophobic pocket of SrtA, surrounded with Ala34, Pro36,
20b, was also found to inhibit the activity of SrtA.161 In vitro Leu39, Ala60, Gly61, and Trp136. Other interactions
studies to determine SrtA inhibition efficacy revealed SrtA responsible for SrtA binding include CH−π interactions
13110 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

between the phenyl ring of 22 and Trp136 and hydrogen


bonding of the inhibitor’s carbonyl oxygen with Arg139 and
Lys140. Furthermore, treatment with 22 protected BALB/c
mice from S. aureus-induced renal abscess and reduced the
mortality rate.
A naturally occurring flavonoid named Dryocrassin ABBA 23
(Table 3), previously extracted from Dryopteridis crassirhizoma-
tis Rhizoma, was found to exhibit SrtA inhibitory activity with
IC50 = 24.17 μM.165 MD simulations showed that 23 binds to
the SrtA active site due to its strong interactions with Ala104,
Glu105, Val166, Gly167, Val168, Ile182, Val193, and Trp194
(Figure 10).
Flavonoids 24 and 25 (Table 3), isolated from Spatholobus
suberectus Dunn., possessed SrtA inhibitory activity.166 Fla-
vonoid 24 showed IC50 = 96.1 μM and MIC > 4.7 mM, while
IC50 and MIC for compound 25 were found to be 74.9 μM and
>477.5 μM, respectively. In another report, naturally occurring
Figure 8. The binding of 20c to the active site of SrtA highlighting the flavonoid, isovitexin 26 (apigenin-6-C-D-glucopyranoside, Table
molecular interactions stabilizing the 20c−SrtA complex. The structure 3), was shown to inhibit SrtA with IC50 = 67.02 μM.167
is based on molecular docking studies. From ref 104, CC BY 4.0. Thus, many flavonoids can be used to treat S. aureus infections
by reducing SrtA-mediated adhesion of bacterial cells to host
tissues. However, these molecules bind to many unrelated
enzymes, and hence the results need to be evaluated with care.
The promiscuous binding behavior and the aggregation-prone
nature of hydrophobic flavonoids could interfere with the
assay.168
4.2.2.4. Sterolglycosides. Liliaceae is a pharmacologically
important plant family containing many steroidal alkaloids. The
genus Fritillaria belongs to the family of Lilliaceae which have
been used as expectorant and anti-hypertensive drugs and are
also prescribed to treat cough, bronchitis, and tumors. The
medically relevant compound, namely β-sitosterol-3-O-gluco-
pyranoside 27 (Table 3) isolated from Fritillaria verticillata
showed SrtA inhibitory activity with IC50 = 31.8 μM.134,169
4.2.2.5. Phenylpropanoids. Traditional Chinese medicines
consist of a large number of biologically active natural products.
Lonicerae f los has been used for the treatment of fever, sores, and
arthritis for many years in China. Chlorogenic acid 28 (Table 3)
Figure 9. Structure of the SrtA−22 complex determined by molecular is a major component of this medicinal herb. Compound 28
modeling. From ref 164, CC BY 4.0. inhibited S. aureus SrtA with IC50 = 95.6 ± 15.7 μM. A high MIC
value of 2.9 mM indicates its negligible effect on bacterial
viability.170 Upon treatment of S. aureus with 28, the cells’ ability

Figure 10. Structure of the SrtA−23 complex based on MD simulations. From ref 165, CC BY 4.0.

13111 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

Figure 11. Structure of SrtA−38 complex determined by MD simulations. From ref 177, CC BY 4.0.

to adhere to fibrinogen-coated plates was quantified using cell interactions with 32 include Gln113, Val166, Val168, Leu169,
adhesion assay. The studies revealed that 28 inhibits the Lys173, Gln178, Thr180, Arg197, and Val201.
adhesion of S. aureus to fibrinogen, hence reducing bacterial 4.2.2.6. Quinone Derivatives. Using HTS, three naturally
virulence. Docking studies and MD simulations were performed occurring compounds 33a, 33b, and 34 (Table 3) possessing a
to determine the structure of the 28−SrtA complex. The studies quinone skeleton were identified as efficient SrtA inhibitors.174
indicated that the phenyl unit of 28 binds to the substrate The naphthoquinones, namely shikonin 33a and alkannin 33b,
recognition site of SrtA through hydrophobic interactions with previously isolated from the roots of Lithospermum erythro-
Leu168 and Val169. In addition to this, hydrophobic rhizon, strongly inhibited S. aureus SrtA with IC50 in the
interactions between Ile182 and Cys184 of SrtA and the central nanomolar range (Table 3). However, both compounds were
region of 28 stabilize the SrtA−28 complex. The studies also found to be toxic to the bacterial cell, which is not desirable.
indicated hydrogen bonding interactions of the guanidine Although 34 exhibited weaker SrtA inhibition activity (IC50 =
moiety of Arg197 with the hydroxyl groups on the cyclohexane 6.2 ± 1.1 μM) than 33a and 33b, it was found to have little
unit of 28. influence on the viability of S. aureus even at high
In another study, five natural products structurally similar to concentrations. Compound 34 irreversibly inhibits SrtA via a
28 were examined for their SrtA inhibitory activity.171 The Michael addition reaction involving the attack of the enzyme’s
bioactivity studies revealed that none of these compounds, reactive thiol group on the carbon−carbon double bond of
including caffeic acid 29a (SI Table S3; IC50 = 860.4 μM) and 34.174 The values of ki, Ki, and k2nd were quantified to be 3.88 ×
ferulic acid 29b (SI Table S3; IC50 = 1323.5 μM), showed 10−2 min−1, 3.57 μM, and 1.1 × 104 M−1 min−1 (Table 3),
improved SrtA inhibition over 28. High MIC values (>5.75
respectively. Reduction in biofilm formation upon treatment
mM) for compounds 29a and 29b indicated their minimal effect
with 34 was also observed. In another study, naturally occurring
on S. aureus viability. However, the cyclized derivative of caffeic
naphthoquinones juglone 35a (Table 3) and plumbagin 35b
acid 30 (Table 3) showed higher S. aureus SrtA inhibition, with
with strong SrtA inhibition potency were discovered.105
IC50 = 36.16 μM.104
Organic extract from dried roots of Pulsatilla koreana, based Naphthoquinones 35a and 35b covalently attach to SrtA, like
on the Korean folk medicine, exhibited anti-SrtA activity against 34.
S. mutans.172 Detailed chromatographic analysis of the extract 4.2.2.7. Bisbenzyl Derivatives. A naturally occurring neutral
showed several lignans and phenylpropanoids with SrtA bisbenzyl compound, erianin 36 (Table 3), could inhibit SrtA
inhibitory activity. Among the isolated 12 natural products, with IC50 = 65.7 ± 7.2 μM and MIC 1.6 mM.175 In silico analysis
compounds 29a and 31 (SI Table S3) exhibited the highest revealed that 36 binds to the active site of SrtA, involving
inhibition against S. mutans SrtA, with IC50 = 20.2 and 20.1 μM, interactions such as electrostatic interactions with Arg197,
respectively. hydrogen bonding interactions with Val193, and hydrophobic
Recently another phenylpropanoid, 32 (SI Table S3), isolated interactions with Ala92, Ala104, Asn112, Ser116, Ile182,
from the pine leaves of Cedrus deodara, inhibited S. aureus Cys184, Trp194, and Ile199.
biofilm formation without affecting bacterial growth.173 The in 4.2.2.8. Chalcone. Chalcone 37 (Table 3) was demonstrated
vitro experiments indicated that 32 inhibited SrtA trans- to exhibit inhibitory activity against SrtA, with an IC50 = 53.15
peptidation activity with IC50 = 161.02 μM. SrtA inactivation μM.176 The compound 37 also reduced the surface display of the
by 32 reduced the display of surface-anchored proteins that are Spa protein in S. aureus and altered the bacteria’s fibronectin-
required for biofilm development. The MD-simulated structure binding ability in a concentration-dependent manner. In
of the SrtA−32 complex revealed that 32 non-covalently binds addition to this, 37 reduced the biofilm formation and the
to the active site of SrtA, involving hydrogen bonding and host-invading ability of S. aureus. In vivo studies showed a lower
hydrophobic interactions. The SrtA residues involved in mortality rate in S. aureus-infected mice upon treatment with 37.
13112 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

Table 4. Chemical Structures of Selected Synthetic Small Molecule SrtA Inhibitors and Their Binding Parameters

13113 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

Table 4. continued

13114 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

Table 4. continued

13115 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

Table 4. continued

13116 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

Table 4. continued

a
Apparent dissociation constant (μM). bSecond-order rate constant of inhibition (M−1 min−1). *Compound is a disulfide dimer.

4.2.2.9. Salvianolic Acid A. Salvianolic acid 38 (Table 3), hydrophobicity seems to be a dominant component of the
isolated from dried roots and rhizomes of Salvia miltiorrhiza molecular recognition. It is capable of attenuating SrtA binding
Bunge, was tested to determine its SrtA inhibition efficacy using across the chemical space domains sampled by these diverse
the FRET-based enzymatic assay.177 The bioactivity studies structural motifs.
revealed that 38 possesses potent anti-SrtA activity with IC50 = 4.3. Synthetic Small-Molecule Sortase A Inhibitors.
11.6 μM. The compound was found to suppress the surface HTS and virtual screening of small-molecule chemical libraries
expression of the Spa protein and reduced S. aureus adhesion to have been used to discover novel SrtA inhibitors. This has led to
fibrinogen and biofilm formation. The MD-simulated structure the evolution of more potent synthetic inhibitors with high
of the SrtA−38 complex (Figure 11) and mutational analysis on selectivity, fewer off-target effects, and higher anti-virulence
SrtA revealed that Ala92, Ala104, Ile182, and Arg197 play activities. As indicated by the structure of SrtA, residues Ala92,
indispensable roles in the binding of 38 to SrtA. Ala104, Val168, Leu169, Ile182, Val193, and Trp194 with
In vivo studies on S. aureus-infected C57BL/6J mice indicated lipophilic side chains constitute a hydrophobic binding pocket in
an 80% enhancement in survival rate upon treatment with 38. the proximity of the enzyme’s active site. Nonpolar groups of the
Moreover, a 100% survival rate in infected mice was observed synthetic inhibitors including aromatic rings and alkyl chains can
upon co-administration of the antibiotic latamoxef along with interact with the SrtA residues in the hydrophobic pocket.
38. Inhibitors can interact via π−π interactions with aromatic side
4.2.2.10. Coumarins. A series of coumarins, 39a−39e (Table chains and the polar functionalities of SrtA and can also form
3), isolated from Poncirus trifoliata showed SrtA inhibitory hydrogen-bonding and electrostatic interactions. Inhibitors can
activity in the range of 19.1−34.5 μM and had no influence on S. reversibly or irreversibly inactivate SrtA by non-covalent and/or
aureus growth.178 Further studies revealed that treatment of S. covalent interactions.
aureus with the coumarin-based inhibitors 39a−39c diminished 4.3.1. Vinyl Sulfones. Based on the inactivation of cysteine
their adhesion ability toward fibronectin- and fibrinogen-coated proteinases by vinyl sulfones,179 McCafferty and co-workers
surfaces. This supported the anti-SrtA activity of coumarins, due examined the SrtA inhibition activity of a series of vinyl
to a reduction in the surface display of fibronectin- and sulfones.180 Among them, vinyl sulfones 40a and 40b (SI Table
fibrinogen-binding proteins. S4) were identified as attractive candidates as SrtA inhibitors
Future Prospects and Outlook. Although more screening of with IC50 = 736 and 190 μM and k2nd = 20.1 and 90 M−1 min−1,
natural products has been performed compared to that of respectively. Mass spectrometric analysis revealed that 40a
peptide-based molecules, in comparison to the available natural reacted with Cys184 present at the enzyme’s active site. The
product resources, this number is very low. The screening of covalent attachment of the inhibitors 40a and 40b through vinyl
marine and plant-based natural products provided promising sulfone groups to SrtA leads to the formation of a thioether
candidates such as 8b, 34, 35, and 38; however, a large chunk of adduct. Compound 40b was more active than 40a due to the
natural products still available provides high hope for discovery. presence of electron-withdrawing −CF3 group. The influence on
Testing many more natural products for inhibitory activity is the adhesion of the S. aureus Newman strain treated with 40a to
necessary and will provide more potent candidates. From our fibronectin-coated plates was also quantified. The studies
survey of the available data, it merits to mention that revealed that 40a reduced bacterial adherence to fibronectin-
13117 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

coated surfaces, indicating the potential of compound 40a as an conducted on BALB/c mice showed that intravenous
effective SrtA inhibitor. administration of 41j to S. aureus-infected mice reduced kidney
4.3.2. Diarylacrylonitriles. Based on the random screening of and joint infections and increased mice survival rate.182
a small-molecules library of 1000 compounds and further in vitro 4.3.3. Prolinated Vinyl Sulfones. Based on SrtA inhibition by
activity, Kim and co-workers reported diarylacrylonitrile 41a, vinyl sulfones180 and the importance of the nitrile group in
with IC50 = 231.09 ± 6.53 μM (SI Table S4), as a hit enzyme binding,181 Kudryavtsev et al. synthesized a series of
compound.181 SAR studies on 41a revealed that replacement of compounds with vinyl sulfone or nitrile groups functionalized on
the ester group by carboxylic acid (41b), amide (41c), or a pyrrolidine scaffold.183 Inhibitors appended with a nitrile
hydrogen (41d) led to a reduction in SrtA inhibitory activity. group showed partial SrtA inhibition even at high concentration.
However, substitution of the ester by nitrile (41e) showed Among the inhibitors evaluated, compounds 42a−42c (SI Table
enhancement in SrtA inhibition with IC50 = 187.40 ± 3.96 μM. S4) exhibited SrtA inhibition, with IC50 ranging from 0.85 to 1.3
The loss in inhibition tendency of 41f, a saturated analog of 41e, mM and k2nd around 104 M−1 min−1. Kinetic and mechanistic
outlined the role played by a double bond for effective studies revealed irreversible time-dependent enzyme inhibition
inhibition. The (Z)-isomer 41g (Table 4) exhibited ∼7-fold by covalent attachment of the inhibitors through the vinyl
more anti-SrtA potency than the (E)-isomer 41e, thus
sulfone group to the active-site Cys184. The high potency of
suggesting that the presence of a nitrile group and suitable
prolinated vinyl sulfones over simple alkyl-vinyl sulfones180
orientation of phenyl rings are important for diarylacrylonitriles
outlined the importance of additional non-covalent hydro-
to interact with SrtA. To further improve SrtA inhibition
potencies, compounds 41h−41l (Table 4) were synthesized. phobic interactions in stabilizing the enzyme−inhibitor
Among them, compound 41j showed the highest activity against complex.
SrtA (IC50 = 9.24 ± 0.47 μM). 4.3.4. Aryl(β-amino)ethyl Ketones. Schneewind and co-
Molecular modeling studies revealed non-covalent interac- workers screened a library of 135 625 small molecules and
tions between SrtA and compound 41j (Figure 12). The phenyl shortlisted 6154 compounds displaying >20% SrtA inhib-
ition.184 Subsequent shortlisting based on their reactivity,
toxicity, and drug-like properties gave 407 compounds. The
promising hits were further analyzed, which led to the
identification of aryl(β-amino)ethyl ketones (AAEKs) 43a and
43b (Table 4) exhibiting Bacillus anthracis and S. aureus SrtA
inhibitory activities. The IC50 values of AAEKs 43a and 43b for
S. aureus SrtA were found to be 47 and 15 μM, respectively.
AAEK 43a was shown to irreversibly inhibit SrtA by covalent
attachment to the enzyme’s active site. This involves
deprotonation of AAEK, followed by β-elimination and attack
by active-site Cys184 on the generated electrophilic olefin
intermediate (Scheme 2).
Figure 12. Structure of the SrtA−41j complex highlighting the 4.3.5. Rhodanines, Pyrazolethiones, and Pyridazinones.
underlying molecular interactions. Adapted from ref 181. Copyright HTS of a small-molecule library containing ∼30 000 com-
2004 American Chemical Society. pounds and sequential elimination of compounds based on the
percentage of SrtA activity inhibition gave 44 compounds as
rings of 41j interacted with side chains of Ala92, Ala104, Val168, potential inhibitors.185 Among these, only 10 compounds were
Leu169, Ile182, Val193, and Trp194 present in the enzyme’s subjected to further studies based on their inhibitory activity and
hydrophobic cavity. In addition to this, hydrogen bonding similarity in physicochemical properties to known drugs. The
interaction between the nitrile group of 41j and guanidine unit IC50 evaluation of these compounds yielded three lead
of Arg197 facilitated enzyme binding. compounds belonging to families of rhodanines, pyrazole-
In another study, inhibitor 41j was demonstrated to prevent thiones, and pyridazinones. These lead compounds were
bacterial infections by inhibiting SrtA and thus reducing demonstrated to reversibly bind the enzyme’s active site, as
fibronectin-binding activity.134 Cytotoxicity analysis on the S. indicated by kinetic studies and mass spectrometric analysis.
aureus Newman strain revealed no influence on cell viability SAR studies on lead inhibitors were performed to discover more
upon treatment with 41j. Further, the in vivo investigations potent SrtA inhibitors.

Scheme 2. Mechanism of SrtA Inhibition by AAEKs 43a and 43b

a
Deprotonation. bβ-Elimination. cMichael-type conjugate addition.

13118 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

The IC50 values of 13 analogs of hit compound 44 (Table 4) Docking studies revealed that the phenyl ring of 47a is
revealed that none of the structural modifications resulted in suitably placed in the enzyme’s hydrophobic pocket. The
higher SrtA inhibition.185 Induced-Fit Docking studies of 44 carbonyl oxygen in 47a is positioned toward Arg197, whereas
into the S. aureus SrtA active site indicated hydrophobic the thiol group points toward His120 at the SrtA active site. The
interactions of the two methyl groups of 44 with Val166 to observed decrease in inhibition ability of 47k over 47e indicates
Val168 and Ile199. The enzyme−inhibitor complex was the importance of the suitable placement of the thiol group.
stabilized by hydrogen-bonding between the nitro and the Among the compounds tested, 47c and 47f do not exert
hydroxyl groups of 44 with SrtA His120, Tyr187, and Trp194. In bactericidal activity and hence are the most promising
the docked molecular structure of the complex, the carbonyl pyridazinone inhibitors.
oxygen of 44 was oriented toward Arg197 and the sulfide toward Representative compounds belonging to the rhodanine,
His120. Rhodanine 44 influenced the bacterial cell viability and pyrazolethione, and pyridazinone series were subjected to
thus is not a promising candidate for the development of anti- combined pharmacophore and 3D quantitative SAR studies to
infective agents against S. aureus infections. identify pharmacophoric features.186 In later studies, to outline
Based on the SAR analysis of 45a, compound 45b (Table 4) the mechanism of SrtA inhibition by 47e and to gain insight into
was synthesized and found to exhibit higher SrtA inhibition the SrtA−47e complex, the water-soluble sodium thiolate
efficacy. Molecular docking studies revealed hydrophobic analog 47l (Table 4) was synthesized.187 Mass spectrometric
interactions of the inhibitor’s phenyl rings with the side chains analysis indicated the formation of a disulfide bond between
of Pro94, Val166, Val168, and Trp194 residues in the enzyme’s SrtA active-site Cys184 and the thiolate of 47l. The NMR
hydrophobic pockets (Figure 13). Significant enhancement in solution structure of the SrtA−47l complex revealed hydro-
phobic interactions between phenyl ring of 47l and Leu97,
Val166, Val168, Leu169, Ile182, and Val193 of SrtA (Figure 14).

Figure 13. Modeled structures of (a) 45a and (b) 45b in the active site
of S. aureus SrtA, generated by Induced-fit docking. Adapted with Figure 14. Structure of the SrtA−47l complex (a, b) showing
permission from ref 185. Copyright 2009 Elsevier. interactions of 47l with residues in the hydrophobic pocket of SrtA.
Adapted with permission from ref 187. Copyright 2017 John Wiley &
Sons.
SrtA inhibition potency was observed upon replacement of the
nitrophenyl group in 45a with a more lipophilic tribromophenyl Stronger inhibition was observed upon substituting R4 with a
substituent. The methyl and thione groups of the pyrazole fluoro (47m) or methoxy (47n) group or positioning an
nucleus were found to be in contact with Ala92 and Arg197 in amidine substituent at the R3 position (47o, Table 4). The rate
the 45b−enzyme complex (Figure 13). Replacement of thione constant, k2nd, for 47o was determined to be 3.6 × 104 M−1
in 45a by ketone yielded 45c, with significantly reduced potency min−1. The improved inhibitory activity is due to additional
of inhibition, whereas replacing phenyl group by an electron- interactions such as multipolar contacts (47m), hydrophobic
withdrawing pyridyl group (45d) increased the SrtA inhibition interactions (47m and 47n), and electrostatic interactions
efficacy. The compound 46 (Table 4) was found to exhibit lower (47o) with SrtA. The length of the hydroxyalkoxy moiety (47p
anti-SrtA activity over 45a. Compounds 45a, 45b, and 45e vs 47q) at the solvent-exposed R2 position influenced binding
exhibited high SrtA inhibitory activity, and no detrimental effect toward SrtA. The compounds with hydroxyalkoxy moiety at R2
on bacterial growth was observed, thus indicating the potential and fluorine at R4 position (47r and 47s) showed strong SrtA
of pyrazolethione compounds as SrtA-targeting anti-virulence inhibition with apparent dissociation constants in the nano-
agents. molar range (Table 4). The high value of k2nd (1.8 × 105 M−1
Among the pyridazinone inhibitors, lead compound 47a min−1) for 47s indicated high SrtA inhibition potency. The
(Table 4) and its derivatives, 47b−47g, were identified as terminal hydroxyl of the 3-hydroxypropoxy group of 47s acted as
potentially active SrtA inhibitors. The SAR studies performed on a hydrogen bond donor to the carbonyl oxygen of Gly192 and as
hit compound 47a highlighted the significant influence of a hydrogen bond acceptor from the hydroxyl of Tyr187 (Figure
substitution pattern on both phenyl and pyridazinone rings. 15). However, due to the shorter length of the alkyl chain at the
Replacement of the methoxy group by an ethoxy along with the R2 position of 47r, the inhibitor did not form favorable
removal of chlorine positioned on the phenyl ring of 47a gave interactions with Tyr187 and Gly192; instead, it formed a single
inhibitor 47e. The ethoxy-thiol pyridazinone inhibitor 47e was hydrogen bond with Ala92, resulting in lower SrtA-binding
found to be one of the efficient SrtA inhibitors, with IC50 = 13 ± ability.
1 μM. 4.3.6. Morpholinobenzoate Derivatives. A commercial
Replacement of thiol by hydroxyl and removal of the chlorine small-molecule library comprising 150 000 compounds was
unit in 47a gave 47h, which significantly reduced the inhibitory subjected to in silico virtual screening by Velu and co-
activity. The symmetrical (47g) and asymmetrical (47i and 47j) workers.188,189 Hit compounds were identified upon applying
disulfide derivatives of 47e were also found to be potent SrtA limiting factors, such as molecular weight, hydrophobicity, and
inhibitors. distance from the SrtA active site. The 108 hit compounds
13119 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

4.3.8. Benzo[d]isothiazol-3(2H)-one-adamantane Amine


Derivatives. HTS of a small-molecule library containing 50 240
compounds against a catalytically active recombinant SrtAΔN59-
6His (SrtAΔN59 with six His units at the C-terminal) gave 41
active compounds with SrtA inhibition >65%.192 Subsequently,
to evaluate binding reversibility and to exclude false positives,
NMR studies on hit compounds were performed. The studies
confirmed that 28 compounds inhibited SrtA activity by
covalently modifying the enzyme’s active site. Among them,
compounds 51a and 52a (Table 4) were found to be promising
leads for further development of anti-infective agents.192,193
Compound 51a binds to the enzyme’s active site with high
affinity and inhibits SrtA with IC50 = 6.11 ± 0.34 μM. However,
it displayed significant bactericidal activity against S. aureus and
was found to be cytotoxic toward NIH 3T3 cells. To improve
SrtA inhibition efficacy and to reduce toxicity toward S. aureus
Figure 15. 3D structures of SrtA−47r (a, c) and SrtA−47s (b, d) and NIH 3T3 cells, compound 51a was subjected to structural
complexes generated by docking the inhibitors to the NMR structure of optimization. This led to the identification of 51b and 51c
SrtA. Adapted with permission from ref 187. Copyright 2017 John (Table 4) as active candidates. NMR spectroscopic studies
Wiley & Sons. indicated the formation of the enzyme−inhibitor complex by
covalent attachment of 51a to the enzyme’s active-site thiol. The
formation of a disulfide bond by benzo[d]isothiazol-3(2H)-one
identified were further screened for inhibition against SrtAΔN59. ring-opening of 51a by SrtA Cys184 was primarily responsible
This led to the discovery of 48a (Table 4) as a lead candidate for for the observed inhibition. A complete loss in SrtA inhibition
the development of SrtA inhibitors. was observed after the removal of the heterocyclic benzisothia-
The FlexX docking model revealed hydrogen bonding zole nucleus or sulfur (51d and 51e, SI Table S4). The NMR-
interactions of the morpholine ring oxygen, amide carbonyl based structure of the SrtA−51a complex revealed the
oxygen, and NH of 48a with Trp194, Ser116, and Glu105, interaction of the phenyl ring with Leu97, Ala118, and
respectively.188 A salt bridge between the carboxylic acid of 48a Trp194, while the adamantyl unit is suitably located in the
and guanidine side chain of Arg197 was also observed. The SrtA hydrophobic pocket, interacting with Val166, Val168,
orientation of thiophene and phenyl unit of 48a suggested that Leu169, Thr180, Ile182, and Ile199. The linker connecting the
appropriate substitutions on these rings could facilitate stronger adamantyl and heterocyclic units plays a crucial role in SrtA
SrtA binding. The ester derivative 48b was found to exhibit inhibition. It performs the dual function of strengthening SrtA
improved SrtA inhibition, with IC50 = 71 ± 2.5 μM. binding by forming a hydrogen bond with Arg197 and suitably
Replacement of the thiophene ring in 48b by furan (48c) orienting the adamantyl group and the heterocyclic ring in a
showed enhancement in enzyme inactivation, whereas hydro- conformation that maximizes SrtA binding.
genation of the double bond of 48b was found to destroy Compounds with extended linkers (51a vs 51b and 51c)
inhibitory activity. Replacement of the carboxylic acid or methyl showed slightly higher inhibition activity. In comparison to 51a,
ester by −CH2OH (48d), −CHO (48e), or −CONH2 resulted inhibitors 51b and 51c reduced the unfavorable influence on
in lower binding to the enzyme. The corresponding acetylenic bacterial cell growth and showed ∼10-fold decrease in
derivatives of 48a and 48b were found to possess less SrtA cytotoxicity against NIH 3T3 cells.
inhibitory activity. In addition to the above structural In another study, the authors demonstrated the potential of
modifications, double bond configuration, amide hydrogen, compound 52a (Table 4) as a SrtA inhibitor due to its fast
and morpholine ring oxygen, and carbonyl group influenced enzyme binding kinetics and low cytotoxicity.193 Results based
SrtA inhibition efficacy. Among the series of morpholino- on NMR spectroscopic and mass spectrometric analyses
benzoate derivatives, compound 48c turned out to be a revealed that the inhibition mechanism involved nucleophilic
promising candidate for further development. attack of Cys184 thiol on the chiral carbon of 52a. The reaction
4.3.7. Dihydro-β-carbolines. Guided by the inhibitory proceeds via the formation of an olefin intermediate generated
activities of indole-containing compounds 49a and 50a (SI after β-elimination of the 2-aminopyridine unit. The NMR
Table S4) against isocitrate lyase from Candida albicans,190 studies also indicated hydrophobic interactions of the thiophene
structurally analogous compounds were synthesized and unit of 52a with Leu97 and Ala118 and π−π stacking
evaluated for their SrtA inhibitory activity.191 SAR studies on interactions with His120 and Trp194 (Figure 16). Replacement
49a revealed that retention of one of the two indole units and of the thiophene by a substituted phenyl ring (52b−52e) led to
both of the α-oxocarbonyl groups are critical to enzyme small changes in their inhibitory activity with little influence on
inhibition. Among the series of compounds synthesized, 49b cytotoxicity.
(Table 4) and 49c showed the highest SrtA inhibition potency 4.3.9. 2-Phenyl-2,3-dihydro-1H-perimidine Derivatives.
with IC50 = 61 and 69 μM, respectively. Virtual screening of a library containing 55 789 compounds by
Structural optimization on compound 50a (SI Table S4) molecular docking led to the identification of 53a (Table 4) as a
highlighted that introduction of phenyl 50b (Table 4) and SrtA inhibitor with IC50 = 47.2 ± 5.9 μM.194 The mechanism of
methoxy group 50c led to dramatic increase in SrtA inhibition, SrtA binding by 53a was investigated using MD simulations and
as indicated by their IC50 values. However, their corresponding SAR analysis. The molecule 53a was shown to function by
tetrahydro-β-carboline derivatives showed a loss of inhibitory blocking the SrtA active site through non-covalent interactions.
activity. MD simulations revealed hydrogen bonds between the carboxyl
13120 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

54c with higher SrtA inhibition potency. Due to the high in vitro
SrtA inhibitory activity and water solubility, inhibitor 54b was
subjected to surface plasmon resonance studies to unravel the
SrtA inhibition mechanism. The studies indicated that 54b
binds to the active site of SrtA, resulting in reversible inhibition.
Treatment of S. aureus with 54b was found to have no influence
on bacterial growth, thus highlighting the efficacy of 54b as an
anti-infective agent. To evaluate its in vivo SrtA inhibition
activity, BALB/c mice infected with S. aureus Newman strain
were injected with 54b. The studies demonstrated that 54b
effectively prevented bacteremia caused by S. aureus.
4.3.11. 2-Phenylhydrazinylidene Derivatives. Maggio and
co-workers screened 200 randomly selected compounds for
SrtA inhibition and identified 55a (SI Table S4) as a lead
compound.195 The SAR studies performed on 55a indicated
that the presence of electron-withdrawing groups on the phenyl
ring increased the anti-SrtA activity of 2-phenylhydrazinylidene
derivatives, whereas the activity decreased with electron-
Figure 16. Structure of SrtA−52a complex showing hydrogen bonding donating substituents. Among the derivatives, compounds
and π−π stacking interactions. The structure was determined by
with chlorine (55b, Table 4) and nitro (55c) substituents
molecular docking studies. Adapted from ref 193. Copyright 2019
American Chemical Society. emerged as strong SrtA inhibitors, with IC50 = 50 and 57 μM,
respectively. The anti-biofilm formation ability of synthesized
group of 53a and His120 and Arg197 (Figure 17). They also derivatives was examined against S. aureus ATCC 25923, ATCC
indicated the existence of a hydrogen bond between the 29213, and ATCC 6538 strains. Among these, compound 55d
was found to reduce the biofilm formation ability of S. aureus by
60% for all three bacterial strains. Taking the studies further,
Cascioferro and co-workers synthesized another series of 2-
phenylhydrazinylidene derivatives, including 55e−55i, and
evaluated them for anti-biofilm and anti-SrtA activity.196 The
compounds 55e and 55f (Table 4) were found to be active
against SrtA, with IC50 = 75 and 76.5 μM, respectively. However,
compounds 55g−55i (SI Table S4) showed weak SrtA
inhibition, with IC50 > 200 μM. The ability of these compounds
to suppress biofilm formation by S. aureus ATCC 25923, ATCC
29213, and ATCC 6538 strains was also examined. The studies
indicated that 55e significantly reduced the biofilm formation
ability of S. aureus ATCC 6538 (IC50 = 0.9 μM). Compounds
55f, 55g, and 55h inhibited biofilm formation by the S. aureus
ATCC 29213 strain, with IC50 = 29, 0.8, and 1.7 μM,
respectively. In addition to this, compounds 55h and 55i
decreased the biofilm formation ability of all three S. aureus
strains. The comparison of bioactivities revealed no correlation
Figure 17. Structure of the SrtA−53a complex generated by molecular between SrtA inhibition and anti-biofilm formation activity for
docking. From ref 194, CC BY-NC-ND 4.0.
the phenylhydrazinylidene derivatives, which can be attributed
to the completely different assays employed for the evaluation.
dihydroperimidine moiety and SrtA Pro163, providing addi- 4.3.12. 2-Phenyl-benzo[d]oxazole-7-carboxamide Deriva-
tional stability to the SrtA−inhibitor complex. The naphthyl ring tives. A series of 2-phenyl-benzo[d]oxazole-7-carboxamide
of 53a interacts with amino acids in the enzyme’s hydrophobic derivatives were synthesized and evaluated for their activity as
pocket that led to further enhancement in SrtA binding affinity. SrtA inhibitors.197 Compounds 56a−56e (Table 4) showed
The suitably oriented polar substituents on the phenyl ring SrtA enzyme inhibition, with the lowest IC50 values among the
facilitated hydrogen-bonding with Gly119, Thr183, and Cys184. series. The replacement of −OH with −OCH3 (56c) had
The in vitro inhibition potency of various 53a derivatives 53b− minimal influence on their SrtA inhibition potency. However,
53e (Table 4 and SI Table S4) supported the binding replacement of −OH by −NO2 in 56a and 56f (Table 4) gave
mechanism. 56g and 56h (SI Table S4) with lower SrtA inhibitory activity.
4.3.10. 3,6-Disubstituted Triazolothiadiazoles. Virtual The studies also indicated that the isobutyl group is
screening of a small-molecule library containing about 300 000 indispensable for good inhibition (56b vs 56i−56k, SI Table
drug-like molecules revealed 105 compounds exhibiting SrtA S4). The docking studies revealed that 56d interacts with
binding.112 Further, evaluation of the inhibition activity of the residues Ala118, Val166, Val168, Val169, and Ile182 through
shortlisted compounds against SrtAΔN24 (SrtA with 24 residues hydrophobic interactions, with Cys184 and Arg197 through
truncated from N-terminus) led to the identification of the 3,6- hydrogen-bonding interactions, and with Tyr194 through π−π
disubstituted triazolothiadiazole derivative (54a, Table 4) as a stacking interactions.
SrtA inhibitor. Structure optimization of the enzyme−substrate 4.3.13. 2-Phenyl-benzofuran-3-carboxamide Derivatives.
complex with hit compound 54a afforded compounds 54b and Based on the discovery of 2-phenyl-benzo[d]oxazole-7-carbox-
13121 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

amide as potent SrtA inhibitors,197 Fu and co-workers designed


a series of 2-phenyl-benzofuran-3-carboxamides derivatives
57a−57n as SrtA inhibitors.198 A total of 37 compounds with
L-shape similar to the LPXTG motif were synthesized. The
higher IC50 value of compound 57a (SI Table S4) over 57b
(Table 4) indicates that the isobutyl group plays a crucial role in
binding with SrtA. SAR studies indicate a correlation between
the number of hydroxyl groups and the potency of inhibition.
Compounds 57c, 57d, and 57j bearing two hydroxyl groups are
more active than 57a, 57b, 57e, 57f, and 57i. The nature of the
substituents at the terminal phenyl ring was found to
significantly influence the inhibitory activity. The higher IC50
of 57g and 57h (SI Table S4) over 57f and 57j (Table 4)
indicated that replacement of the −OH group(s) by −CF3 or
−NO2 reduced the SrtA inhibition activity. In addition to this, Figure 19. MD-simulated structure of the covalent adduct formed by
substitution of −OH by −SH, −NH2 (57b vs 57k and 57l, Table 58c and Cys184 of SrtA. Adapted with permission from ref 199.
4), or −F (57b vs 57m, SI Table S4) also reduced the efficacy of Copyright 2019 Elsevier.
inhibition. Replacement of the ester group as the linker (57e) by
an ether (57n) leads to a decrease in inhibitory activity. docking studies revealed Arg197 and His120 make non-covalent
Compounds 57c, 57d, and 57j are the most potent inhibitors interactions with the p-nitrophenyl and the pyridine rings of 58c.
among the series of compounds evaluated. Molecular docking The liquid chromatography−mass spectrometry (LC-MS)
simulations revealed that the isobutyl group of 57d interacts analysis showed the formation of covalent disulfide adduct,
with Thr164, Val166, Val168, Leu169, and Ile182 SrtA residues. supporting the irreversible inhibition.
A hydrogen bonding interaction of the ester carbonyl of 57d 4.3.15. Thiadiazolidinone Derivatives. Thidiazolidinone
with Cys184 and Arg197 is also indicated (Figure 18). scaffold is a well-known pharmacophore in drug discovery. It
is present in tideglusib (59a, Table 4), a candidate under clinical
trial for myotonic dystrophy. Tideglusib is an irreversible
inhibitor of glycogen synthase kinase 3β and is also under clinical
investigation for the treatment of Alzheimer’s disease.200 HTS of
2400 drug candidates using FRET assay led to the identification
of tideglusib with strong in vitro SrtA inhibition.113 SAR studies
gave compounds 59b−59i, showing inhibition potency
comparable to that of 59a, with minimal influence on S. aureus
viability.
The SrtA inhibitory activity of the derivatives was determined
using the FRET and PAGE-based assays. The orthogonal
PAGE-based assay could rule out any false positives from the
FRET assay. Further studies revealed that 59a showed reversible
inhibition by non-covalently binding to the enzyme’s substrate-
binding pocket (Figure 20). The docked structure of the SrtA−

Figure 18. Structure of the SrtA−57d complex determined using


molecular modeling. Adapted with permission from ref 198. Copyright
2017 Elsevier.

4.3.14. Thiadiazole Derivatives. HTS of two commercially


available small-molecule libraries containing ∼28 500 com- Figure 20. Modeled structure of the SrtA−inhibitor complex showing
pounds gave 110 hits. Further concentration-dependent studies potential binding modes between tideglusib (59a) and SrtAΔN59.
and applying the PAINS filter led to the identification of Adapted from ref 113. Copyright 2020 American Chemical Society.
thiadiazole 58a (Table 4) as an inhibitor of SrtA.199 NMR
experiments revealed that 58a binds to SrtA, and so it was 59a complex indicated π−π stacking of the naphthyl ring and
selected as the starting point for further structural optimization. hydrogen bonding interactions of the sulfur atom of 59a with the
A series of structural modifications on 58a, such as removal and active-site His120 and Arg197. The in vivo experiments revealed
addition of different groups, were done to optimize the structure that 59a is a potential candidate for developing anti-virulence
for obtaining good inhibitors. Structure optimization of 58a led agents.
to the identification of inhibitors 58b and 58c. Among them, 58c 4.3.16. Alkaloid Derivatives. Diana and co-workers synthe-
with IC50 = 3.8 ± 0.3 μM was found to possess the strongest SrtA sized a series of topsentin analogs133 and evaluated their biofilm
inhibition activity without influencing S. aureus viability. inhibition efficacy.201 Many of the synthesized 1,2,4-oxadiazole
Molecular dockings using standard and induced fit approaches topsentin analogs inhibited biofilm formation in Gram-positive
showed that 58c occupies the active site (Figure 19). The S. aureus ATC25923 and Gram-negative P. aeruginosa
13122 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

ATC15442. Among them, compounds 60a−60c (Table 4) lacrylonitrile derivative 41j showed decreased mortality in S.
displayed higher biofilm inhibition in S. aureus ATCC25923 aureus-infected mice.182 The triazolothiadiazole derivative 54b
with BIC50 (concentration that showed 50% inhibition of showed good SrtA inhibition and reduced mortality in S. aureus-
biofilm formation) of 9.7, 0.7, and 2.2 μM, respectively. These infected mice. The compound 54b is under preclinical
compounds were also demonstrated as strong SrtA inhibitors evaluation.112,203 Diana and co-workers reported 1,2,4-oxadia-
with IC50 = 2.2 μM (60a), 10.4 μM (60b), and 2.2 μM (60c), zole derivatives 60a−60c that showed very promising activity
with minimal influence on bacterial growth. against biofilm formation, and their non-toxicity against human
4.3.17. Disulfanylbenzamides. Based on the previously cells is worth mentioning.201
reported benzisothiazolinones inhibitor 51a,192 Schirmeister
and co-workers designed a series of disulfanylbenzamide 5. CONCLUSIONS AND FUTURE PERSPECTIVES
derivatives 61a−61h (Table 4) containing disulfide warhead Antibiotics that are available today act due to their bacteriostatic
chemotype and evaluated their SrtA inhibition efficiencies.202 In or bactericidal properties. The indiscriminate use of antibiotics
vitro and in silico studies revealed that the majority of resulted in the emergence of drug resistance that has wreaked
disulfanylbenzamide derivatives showed time-dependent and havoc worldwide. It is an imminent threat to public health, and
irreversible inhibition of SrtA. The irreversible inhibition is therefore new strategies for combating antibiotic resistance are
confirmed by LC-MS analysis, which revealed the disulfide urgently needed.
adduct. The presence of reducing agents like 1,4-dithiothreitol The widespread and dangerous S. aureus infections demand
(DTT) or tris(2-carboxyethyl)phosphine (TCEP) during effective anti-infective therapy. SrtA as a biological anti-virulence
inhibition was also investigated to find the mode of inhibition. target for therapeutic intervention is promising, since the
Disulfide reducing agents DTT and TCEP could typically regain inhibition against SrtA may not put selective pressure on
the enzymatic activity in the presence of covalent inhibitors. bacteria resulting in the emergence of drug-resistant strains. In
Molecular modeling studies using the NMR structure of SrtA addition, the extracellular location of SrtA and the absence of a
(Figure 21) indicated that the side-chain carbonyl oxygen of 61a eukaryotic homolog make SrtA an attractive target. Therefore,
SrtA inhibition appears to be an excellent anti-infective
therapeutic strategy.41
Although a sufficient number of SrtA inhibitors are known
today, many inhibitors are not further pursued. The majority of
the SrtA inhibitors were evaluated in vitro, and the interactions
are rationalized based on in silico studies. Only a limited number
of SrtA inhibitors, for example, 41j and 54b, have been studied in
vivo,182,112 and none of the reported SrtA inhibitors are currently
under clinical evaluation. This suggests that there is an urgent
need to follow up on the potent inhibitors for translation into
drugs.
A large number of the SrtA inhibitors have limited potency
and poor selectivity. Many of the inhibitors may be falsely
positive or promiscuous. Introduction of the PAINS filter might
improve the situation by removing bad ones and selecting the
true hit compounds.121 Optimization of the true SrtA inhibitors
from the present list of inhibitors will be a starting point for drug
development.118
Peptidomimetics are an important class of compounds in
medicinal chemistry since they are closer to the native ligands,
but with enhanced proteolytic stability, improved bioavailability,
Figure 21. The structure of SrtA−61a complex indicating the and pharmacokinetics.122,123 Only a very limited number of
molecular interactions. The structure was generated by docking 61a peptidomimetics SrtA inhibitors have been developed. Given
in the NMR structure of SrtA. From ref 202, CC BY-NC-ND 4.0. their central importance in drug discovery, more investigations
in this direction are needed. A diversity-oriented synthesis that
is hydrogen bonded to Arg197. The hydrophobic interactions enables combinatorial variation of amino acid residues at various
and π−π stacking interactions assisted the formation of positions of the native ligand will be a powerful approach for
enzyme−inhibitor complex. The biochemical studies indicated finding peptidomimetic SrtA inhibitors.
that 61h showed significant SrtA inhibition, efficient reduction Natural products or their derivatives constitute almost half of
in fibrinogen binding, weak bactericidal activity, and low the best-selling drugs.204 Many natural products have been
cytotoxicity toward mammalian cells. In addition to these, tested and showed promising SrtA inhibitory properties; hence,
compound 61h showed selectivity toward S. aureus SrtA over more extensive screening is a way forward. Natural products
other structurally related proteases. represent an untapped reservoir of drugs and hence are good
Future Prospects and Outlook. Various research groups have starting points for designing ligands for SrtA. Natural products
attempted to find small-molecule SrtA inhibitors. However, only are biologically validated starting points in structural space for
a few promising candidates have been identified that could serve library design.205 Focused libraries based on natural products
as a starting point for further development of SrtA inhibitors are, therefore, another way to obtain highly active inhibitors.
with therapeutic potential. Pyridazinone-based small-molecule Synthetic molecules provide a rich alternative to natural
covalent inhibitor 47s is a highly promising candidate with high products and leverage the power of organic chemistry to invent
SrtA inactivation efficiency.187 The in vivo studies on diary- required molecular shape for a given chemical space. Natural-
13123 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

product-inspired library design is a promising approach for the Authors


discovery of SrtA inhibitors. The available structural space for Rachit Sapra − Department of Chemistry, Indian Institute of
synthetic molecules is high and unending. Though considerable Technology Delhi, New Delhi 110016, India
efforts have been put into the discovery of small-molecule SrtA Amit K. Rajora − Department of Chemistry, Indian Institute of
inhibitors, many potent structural units are not yet discovered. Technology Delhi, New Delhi 110016, India; Present
Fine-tuning of the promising scaffolds and new entities still Address: Pharmaceutical Sciences Laboratory, Faculty of
awaits discovery. Science and Engineering, Åbo Akademi University, 20520
The anti-virulence drug discovery is in its infancy. Though Turku, Finland
anti-virulence drugs are urgently needed for combating Pushpendra Kumar − Department of Chemistry, Indian
antibiotic-resistant bacteria, there are many challenges ahead. Institute of Technology Delhi, New Delhi 110016, India
The broad chemical diversity of SrtA inhibitors provides Govind P. Maurya − Department of Chemistry, Indian Institute
important cues for the future design of effective SrtA inhibitors, of Technology Delhi, New Delhi 110016, India
but of course, there are many more hurdles ahead.
Complete contact information is available at:
Anti-virulence drugs act by attenuation of virulence; hence, it
https://pubs.acs.org/10.1021/acs.jmedchem.1c00386
is believed that no resistant strains will emerge. However, recent
studies have shown that resistance does exist, challenging the
Notes
evolution-proof hypothesis.32−34 Therefore, more effective
therapies are needed. In a recent report, salvianolic acid A, a The authors declare no competing financial interest.
natural product that inhibits SrtA (IC50 = 11.6 μM), along with Biographies
the antibiotic latamoxef, protected mice against lethal pneumo- Rachit Sapra completed his Bachelor’s degree in 2014 from St.
nia caused by MRSA.177 This study indicates that the use of an Stephen’s College, University of Delhi, India. He completed his
anti-virulence drug along with a known antibiotic compound is a Master’s degree from the same college in 2016. Currently, he is
promising approach for combating antibiotic resistance.30,206
pursuing his Ph.D. from the Department of Chemistry, Indian Institute
The good functioning of the host immune system is a
of Technology Delhi (IITD), under the supervision of Prof. V. Haridas.
requirement for the effective functioning of the SrtA inhibitors
as drugs. This will be a problem for immunocompromised His research interests include the design and synthesis of peptide-based
patients; hence, the use of immune-stimulating agents along molecular sensors and nanoparticles for biomedical applications.
with SrtA inhibitors is a viable option. Amit Kumar Rajora has completed his Master of Science in Organic
Finally, while the discovery of anti-virulence agents is still in Chemistry from Jamia Millia Islamia University, New Delhi, India, and
its infancy, the detailing of the many promising SrtA inhibitors Ph.D. in Pharmacy (pharmaceutics) from the University of Reading,
elaborated in this Perspective should galvanize the research United Kingdom. He has also worked as Research Associate at the
community to explore the possibilities on offer. Indian Institute of Technology, New Delhi, India, and as a National
Concerted and focused efforts from synthetic organic
Postdoctoral Fellow at the Regional Centre for Biotechnology,
chemists and biologists are urgently needed to discover
Faridabad, Haryana. Currently, he is working as a Postdoctoral Fellow
molecules with high SrtA inhibition and their eventual
transformation to useful drugs. As this Perspective has shown, at Åbo Akademi University, Turku, Finland.
based on a very limited screening and rather elementary small- Pushpendra Kumar received his Bachelor’s (2017) and Masters’s
molecule design efforts, potent SrtA inhibitors have already been degrees in Chemistry (2019) from Dayalbagh Educational Institute,
discovered. In the future, the identification of peptide-based Agra, India. He is interested in the design and synthesis of self-
molecules, natural products, and designed small molecules as assembling peptide molecules for biomedical and material applications.
potent SrtA inhibitors will create a paradigm shift in the
treatment of bacterial diseases. Govind P. Maurya received his M.Sc. degree in Synthetic Organic


Chemistry from M. L. K. PG College, Balrampur (affiliated with RMLA
ASSOCIATED CONTENT University, Faizabad, Uttar Pradesh, India), in 2014. He is currently a
Ph.D. student in the research group of Prof. V. Haridas in the
*
sı Supporting Information
Department of Chemistry, IITD. His research interest is in the design
The Supporting Information is available free of charge at and synthesis of self-assembling peptides for biological applications.
https://pubs.acs.org/doi/10.1021/acs.jmedchem.1c00386.
Nalin Pant is a Professor in the Department of Chemistry at IITD,
Chemical structures of peptidomimetic (Table S1), India. His research group is involved in various aspects of molecular
marine (Table S2), and plant-based natural products structures, bioorganic chemistry, and constrained molecular systems.
(Table S3), and synthetic small-molecule (Table S4) SrtA
inhibitors with IC50 > 100 μM (PDF) V. Haridas is a Professor in the Department of Chemistry at IITD,


India. His research group is involved in the design and synthesis of a
variety of peptidic and pseudopeptidic molecules for biomedicinal
AUTHOR INFORMATION applications. In addition, the group is also working on various aspects of
Corresponding Authors synthetic and bioorganic chemistry.


V. Haridas − Department of Chemistry, Indian Institute of
Technology Delhi, New Delhi 110016, India; orcid.org/ ACKNOWLEDGMENTS
0000-0002-2931-0585; Phone: +011-26591380;
Email: haridasv@chemistry.iitd.ac.in We thank DST-SERB (Project: EMR/2017/003192/OC) for
Nalin Pant − Department of Chemistry, Indian Institute of funding. R.S. and P.K. thank CSIR for the doctoral fellowships.
Technology Delhi, New Delhi 110016, India; Phone: +011- G.P.M. thanks University Grants Commission (UGC), New
26591525; Email: nalinp@chemistry.iitd.ac.in Delhi, for the doctoral fellowship.
13124 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry


pubs.acs.org/jmc Perspective

ABBREVIATIONS USED (13) Patrick, G. L. An Introduction to Medicinal Chemistry, 5th ed.;


Oxford University Press: Oxford, 2013.
AAEK, aryl(β-amino)ethyl ketone; Abz, 2-aminobenzoyl; CDC, (14) Parrino, B.; Carbone, D.; Cirrincione, G.; Diana, P.; Cascioferro,
U.S. Centers for Disease Control and Prevention; Dabcyl, 4- S. Inhibitors of Antibiotic Resistance Mechanisms: Clinical Applica-
([4′-(dimethylamino)phenyl]azo)benzoic acid; Dap, 2,3-dia- tions and Future Perspectives. Future Med. Chem. 2020, 12, 357−359.
minopropionic acid; DNP, 2,4-dinitrophenyl; Fnbp’s, fibronec- (15) Rice, L. B. Federal Funding for the Study of Antimicrobial
tin-binding proteins; DTT, 1,4-dithiothreitol; EDANS, 5-[(2- Resistance in Nosocomial Pathogens: No ESKAPE. J. Infect. Dis. 2008,
aminoethyl)amino]naphthalene-1-sulfonic acid; FRET, fluores- 197, 1079−1081.
cence resonance energy transfer; GN, N-acetyl glycosamine; (16) Taubes, G. The Bacteria Fight Back. Science 2008, 321, 356−361.
HPLC, high-performance liquid chromatography; HTS, high- (17) Hu, Q.; Peng, H.; Rao, X. Molecular Events for Promotion of
throughput screening; IC50, half-maximal inhibitory concen- Vancomycin Resistance in Vancomycin Intermediate Staphylococcus
aureus. Front. Microbiol. 2016, 7, 1601.
tration; IsdA, iron-regulated surface determinant A; k2nd, (18) Monserrat-Martinez, A.; Gambin, Y.; Sierecki, E. Thinking
second-order rate constant of inhibition; ki, first-order rate Outside the Bug: Molecular Targets and Strategies to Overcome
constant of inactivation; Ki, inhibitory constant; LC-MS, liquid Antibiotic Resistance. Int. J. Mol. Sci. 2019, 20, 1255.
chromatography−mass spectrometry; MD, molecular dynam- (19) Fala, L. Sivextro (Tedizolid Phosphate) Approved for the
ics; MIC, minimum inhibitory concentration; MN, N- Treatment of Adults with Acute Bacterial Skin and Skin-Structure
acetylmuramic acid; MRSA, methicillin-resistant S. aureus; Infections. Am. Health Drug Benefits 2015, 8, 111−115.
MSCRAMMs, microbial surface components recognizing (20) 2019 Antibacterial Agents in Clinical Development: An Analysis of
adhesive matrix molecules; OD, optical density; PAc, protein the Antibacterial Clinical Development Pipeline; WHO, 2020.
antigen c; PAINS, pan assay interference compounds; PBS, (21) Santajit, S.; Indrawattana, N. Mechanisms of Antimicrobial
phosphate-buffered saline; SAR, structure−activity relationship; Resistance in ESKAPE Pathogens. BioMed Res. Int. 2016, 2016, 1−8.
SDS-PAGE, sodium docdecyl sulfate−polyacrylamide gel (22) De Oliveira, D. M. P.; Forde, B. M.; Kidd, T. J.; Harris, P. N. A.;
Schembri, M. A.; Beatson, S. A.; Paterson, D. L.; Walker, M. J.
electrophoresis; Spa, Staphylococcal protein A; SrtA, sortase Antimicrobial Resistance in ESKAPE Pathogens. Clin. Microbiol. Rev.
A; S. aureus, Staphylococcus aureus; S. mutans, Streptococcus 2020, 33, No. e00181-19.
mutans; Sec, secretory; TCEP, tris(2-carboxyethyl)phosphine (23) Ma, Y.-X.; Wang, C.-Y.; Li, Y.-Y.; Li, J.; Wan, Q.-Q.; Chen, J.-H.;

■ REFERENCES
(1) Sherris Medical Microbiology, 6th ed.; Ryan, K. J., Ray, C. G., Eds.;
Tay, F. R.; Niu, L.-N. Considerations and Caveats in Combating
ESKAPE Pathogens against Nosocomial Infections. Adv. Sci. 2020, 7,
1901872.
(24) O’Neill, J. Tackling Drug-Resistant Infections Globally: Final Report
McGraw-Hill Education Medical: New York, 2014. and Recommendations. The Review on Antimicrobial Resistance, May
(2) Tong, S. Y. C.; Davis, J. S.; Eichenberger, E.; Holland, T. L.; 2016.
Fowler, V. G. Staphylococcus aureus Infections: Epidemiology, (25) Antibiotic Resistance Threats in the United States; U.S. Department
Pathophysiology, Clinical Manifestations, and Management. Clin. of Health and Human Services, 2019.
Microbiol. Rev. 2015, 28, 603−661. (26) Cascioferro, S.; Carbone, D.; Parrino, B.; Pecoraro, C.;
(3) Lowy, F. D. Staphylococcus aureus Infections. N. Engl. J. Med. 1998, Giovannetti, E.; Cirrincione, G.; Diana, P. Therapeutic Strategies To
339, 520−532. Counteract Antibiotic Resistance in MRSA Biofilm-Associated
(4) Diekema, D. J.; Pfaller, M. A.; Schmitz, F. J.; Smayevsky, J.; Bell, J.; Infections. ChemMedChem 2021, 16, 65−80.
Jones, R. N.; Beach, M. Survey of Infections Due to Staphylococcus (27) Rasko, D. A.; Sperandio, V. Anti-Virulence Strategies to Combat
Species: Frequency of Occurrence and Antimicrobial Susceptibility of Bacteria-Mediated Disease. Nat. Rev. Drug Discovery 2010, 9, 117−128.
Isolates Collected in the United States, Canada, Latin America, Europe, (28) Mühlen, S.; Dersch, P. Anti-Virulence Strategies to Target
and the Western Pacific Region for the SENTRY Antimicrobial Bacterial Infections. Curr. Top. Microbiol. Immunol. 2015, 398, 147−
Surveillance Program, 1997−1999. Clin. Infect. Dis. 2001, 32, S114− 183.
S132. (29) Parrino, B.; Diana, P.; Cirrincione, G.; Cascioferro, S. Bacterial
(5) Fernandes, R.; Amador, P.; Prudêncio, C. β-Lactams: Chemical Biofilm Inhibition in the Development of Effective Anti-Virulence
Structure, Mode of Action and Mechanisms of Resistance. Rev. Med. Strategy. Open Med. Chem. J. 2018, 12, 84−87.
Microbiol. 2013, 24, 7−17. (30) Watkins, K.; Unnikrishnan, M. New Strategies and Targets for
(6) Bal, A. M.; Gould, I. M. Antibiotic Resistance in Staphylococcus Antibacterial Discovery. Drug Discovery Targeting Drug-Resistant
aureus and Its Relevance in Therapy. Expert Opin. Pharmacother. 2005, Bacteria; Elsevier. 2020, 249−272.
6, 2257−2269. (31) Cegelski, L.; Marshall, G. R.; Eldridge, G. R.; Hultgren, S. J. The
(7) Foster, T. J. Can β-Lactam Antibiotics be Resurrected to Combat Biology and Future Prospects of Antivirulence Therapies. Nat. Rev.
MRSA? Trends Microbiol. 2019, 27, 26−38. Microbiol. 2008, 6, 17−27.
(8) Llarrull, L. I.; Testero, S. A.; Fisher, J. F.; Mobashery, S. The (32) Imperi, F.; Fiscarelli, E. V.; Visaggio, D.; Leoni, L.; Visca, P.
Future of the β-Lactams. Curr. Opin. Microbiol. 2010, 13, 551−557. Activity and Impact on Resistance Development of Two Antivirulence
(9) Maple, P. A. C.; Hamilton-Miller, J. M. T.; Brumfitt, W. World- Fluoropyrimidine Drugs in Pseudomonas aeruginosa. Front. Cell. Infect.
Wide Antibiotic Resistance in Methicillin-Resistant Staphylococcus Microbiol. 2019, 9, 49.
aureus. Lancet 1989, 333, 537−540. (33) Allen, R. C.; Popat, R.; Diggle, S. P.; Brown, S. P. Targeting
(10) Cosgrove, S. E.; Carroll, K. C.; Perl, T. M. Staphylococcus aureus Virulence: Can We Make Evolution-Proof Drugs? Nat. Rev. Microbiol.
with Reduced Susceptibility to Vancomycin. Clin. Infect. Dis. 2004, 39, 2014, 12, 300−308.
539−545. (34) Totsika, M. Benefits and Challenges of Antivirulence
(11) Marty, F. M.; Yeh, W. W.; Wennersten, C. B.; Venkataraman, L.; Antimicrobials at the Dawn of the Post-Antibiotic Era. Drug Delivery
Albano, E.; Alyea, E. P.; Gold, H. S.; Baden, L. R.; Pillai, S. K. Lett. 2016, 6, 30−37.
Emergence of a Clinical Daptomycin-Resistant Staphylococcus aureus (35) Suree, N.; Jung, M. E.; Clubb, R. T. Recent Advances Towards
Isolate During Treatment of Methicillin-Resistant Staphylococcus aureus New Anti-Infective Agents That Inhibit Cell Surface Protein Anchoring
Bacteremia and Osteomyelitis. J. Clin. Microbiol. 2006, 44, 595−597. in Staphylococcus aureus and Other Gram-Positive Pathogens. Mini-Rev.
(12) Foster, T. J. Antibiotic Resistance in Staphylococcus aureus. Med. Chem. 2007, 7, 991−1000.
Current Status and Future Prospects. FEMS Microbiol. Rev. 2017, 41, (36) Mazmanian, S. K. Staphylococcus aureus Sortase, an Enzyme That
430−449. Anchors Surface Proteins to the Cell Wall. Science 1999, 285, 760−763.

13125 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

(37) Ton-That, H.; Mazmanian, S. K.; Faull, K. F.; Schneewind, O. (60) Spirig, T.; Weiner, E. M.; Clubb, R. T. Sortase Enzymes in Gram-
Anchoring of Surface Proteins to the Cell Wall of Staphylococcus aureus: Positive Bacteria. Mol. Microbiol. 2011, 82, 1044−1059.
Sortase Catalyzed In Vitro Transpeptidation Reaction Using LPXTG (61) Marraffini, L. A.; DeDent, A. C.; Schneewind, O. Sortases and the
Peptide and NH2-Gly3 Substrates. J. Biol. Chem. 2000, 275, 9876−9881. Art of Anchoring Proteins to the Envelopes of Gram-Positive Bacteria.
(38) Clancy, K. W.; Melvin, J. A.; McCafferty, D. G. Sortase Microbiol. Mol. Biol. Rev. 2006, 70, 192−221.
Transpeptidases: Insights into Mechanism, Substrate Specificity, and (62) Mazmanian, S. K.; Ton-That, H.; Su, K.; Schneewind, O. An
Inhibition. Biopolymers 2010, 94, 385−396. Iron-Regulated Sortase Anchors a Class of Surface Protein during
(39) Si, L.; Li, P.; Liu, X.; Luo, L. Chinese Herb Medicine against Staphylococcus aureus Pathogenesis. Proc. Natl. Acad. Sci. U. S. A. 2002,
Sortase A Catalyzed Transformations, a Key Role in Gram-Positive 99, 2293−2298.
Bacterial Infection Progress. J. Enzyme Inhib. Med. Chem. 2016, 31, (63) Zecconi, A.; Scali, F. Staphylococcus aureus Virulence Factors in
184−196. Evasion from Innate Immune Defenses in Human and Animal Diseases.
(40) Maresso, A. W.; Schneewind, O. Sortase as a Target of Anti- Immunol. Lett. 2013, 150, 12−22.
Infective Therapy. Pharmacol. Rev. 2008, 60, 128−141. (64) Mazmanian, S. K.; Ton-That, H.; Schneewind, O. Sortase-
(41) Cascioferro, S.; Totsika, M.; Schillaci, D. Sortase A: An Ideal Catalysed Anchoring of Surface Proteins to the Cell Wall of
Target for Anti-Virulence Drug Development. Microb. Pathog. 2014, 77, Staphylococcus aureus. Mol. Microbiol. 2001, 40, 1049−1057.
105−112. (65) Mazmanian, S. K.; Liu, G.; Jensen, E. R.; Lenoy, E.; Schneewind,
(42) Cascioferro, S.; Raffa, D.; Maggio, B.; Raimondi, M. V.; Schillaci, O. Staphylococcus aureus Sortase Mutants Defective in the Display of
D.; Daidone, G. Sortase A Inhibitors: Recent Advances and Future Surface Proteins and in the Pathogenesis of Animal Infections. Proc.
Perspectives. J. Med. Chem. 2015, 58, 9108−9123. Natl. Acad. Sci. U. S. A. 2000, 97, 5510−5515.
(43) Guo, Y.; Cai, S.; Gu, G.; Guo, Z.; Long, Z. Recent Progress in the (66) Weiss, W. J.; Lenoy, E.; Murphy, T.; Tardio, L.; Burgio, P.;
Development of Sortase A Inhibitors as Novel Anti-Bacterial Virulence Projan, S. J.; Schneewind, O.; Alksne, L. Effect of SrtA and SrtB Gene
Agents. RSC Adv. 2015, 5, 49880−49889. Expression on the Virulence of Staphylococcus aureus in Animal Models
(44) Ha, M. W.; Yi, S. W.; Paek, S.-M. Design and Synthesis of Small of Infection. J. Antimicrob. Chemother. 2004, 53, 480−486.
Molecules as Potent Staphylococcus aureus Sortase A Inhibitors. (67) Flemming, H.-C.; Wingender, J.; Szewzyk, U.; Steinberg, P.; Rice,
Antibiotics 2020, 9, 706. S. A.; Kjelleberg, S. Biofilms: An Emergent Form of Bacterial Life. Nat.
(45) Pizarro-Cerdá, J.; Cossart, P. Bacterial Adhesion and Entry into Rev. Microbiol. 2016, 14, 563−575.
Host Cells. Cell 2006, 124, 715−727. (68) Suresh, M. K.; Biswas, R.; Biswas, L. An Update on Recent
(46) Cascioferro, S.; Cusimano, M. G.; Schillaci, D. Antiadhesion Developments in the Prevention and Treatment of Staphylococcus
Agents against Gram-Positive Pathogens. Future Microbiol. 2014, 9, aureus Biofilms. Int. J. Med. Microbiol. 2019, 309, 1−12.
1209−1220. (69) Figueiredo, A. M. S.; Ferreira, F. A.; Beltrame, C. O.; Côrtes, M.
(47) Joh, D.; Wann, E. R.; Kreikemeyer, B.; Speziale, P.; Höök, M. F. The Role of Biofilms in Persistent Infections and Factors Involved in
Role of Fibronectin-Binding MSCRAMMs in Bacterial Adherence and Ica-Independent Biofilm Development and Gene Regulation in
Entry into Mammalian Cells. Matrix Biol. 1999, 18, 211−223. Staphylococcus aureus. Crit. Rev. Microbiol. 2017, 43, 602−620.
(48) Scott, J. R.; Barnett, T. C. Surface Proteins of Gram-Positive (70) Savage, V. J.; Chopra, I.; O’Neill, A. J. Staphylococcus aureus
Bacteria and How They Get There. Annu. Rev. Microbiol. 2006, 60, Biofilms Promote Horizontal Transfer of Antibiotic Resistance.
397−423. Antimicrob. Agents Chemother. 2013, 57, 1968−1970.
(49) Vengadesan, K.; Narayana, S. V. L. Structural Biology of Gram- (71) Parrino, B.; Schillaci, D.; Carnevale, I.; Giovannetti, E.; Diana, P.;
Positive Bacterial Adhesins. Protein Sci. 2011, 20, 759−772. Cirrincione, G.; Cascioferro, S. Synthetic Small Molecules as Anti-
(50) Foster, T. J.; Geoghegan, J. A.; Ganesh, V. K.; Höök, M. Biofilm Agents in the Struggle against Antibiotic Resistance. Eur. J. Med.
Adhesion, Invasion and Evasion: The Many Functions of the Surface Chem. 2019, 161, 154−178.
Proteins of Staphylococcus aureus. Nat. Rev. Microbiol. 2014, 12, 49−62. (72) Moormeier, D. E.; Bayles, K. W. Staphylococcus aureus Biofilm: A
(51) Mandlik, A.; Swierczynski, A.; Das, A.; Ton-That, H. Pili in Complex Developmental Organism: Molecular Mechanisms of S. Aureus
Gram-Positive Bacteria: Assembly, Involvement in Colonization and Biof ilm Development. Mol. Microbiol. 2017, 104, 365−376.
Biofilm Development. Trends Microbiol. 2008, 16, 33−40. (73) Sibbald, M. J. J. B.; Yang, X.-M.; Tsompanidou, E.; Qu, D.;
(52) Schneewind, O.; Model, P.; Fischetti, V. A. Sorting of Protein a to Hecker, M.; Becher, D.; Buist, G.; van Dijl, J. M. Partially Overlapping
the Staphylococcal Cell Wall. Cell 1992, 70, 267−281. Substrate Specificities of Staphylococcal Group A Sortases. Proteomics
(53) Schneewind, O.; Mihaylova-Petkov, D.; Model, P. Cell Wall 2012, 12, 3049−3062.
Sorting Signals in Surface Proteins of Gram-Positive Bacteria. EMBO J. (74) O’Neill, E.; Pozzi, C.; Houston, P.; Humphreys, H.; Robinson, D.
1993, 12, 4803−4811. A.; Loughman, A.; Foster, T. J.; O’Gara, J. P. A Novel Staphylococcus
(54) Navarre, W. W.; Schneewind, O. Proteolytic Cleavage and Cell aureus Biofilm Phenotype Mediated by the Fibronectin-Binding
Wall Anchoring at the LPXTG Motif of Surface Proteins in Gram- Proteins, FnBPA and FnBPB. J. Bacteriol. 2008, 190, 3835−3850.
Positive Bacteria. Mol. Microbiol. 1994, 14, 115−121. (75) Bierne, H.; Mazmanian, S. K.; Trost, M.; Pucciarelli, M. G.; Liu,
(55) Schneewind, O.; Fowler, A.; Faull, K. F. Structure of the Cell Wall G.; Dehoux, P.; Jänsch, L.; Garcia-del Portillo, F.; Schneewind, O.;
Anchor of Surface Proteins in Staphylococcus aureus. Science 1995, 268, Cossart, P. Inactivation of the SrtA Gene in Listeria monocytogenes
103−106. Inhibits Anchoring of Surface Proteins and Affects Virulence. Mol.
(56) Ton-That, H.; Liu, G.; Mazmanian, S. K.; Faull, K. F.; Microbiol. 2002, 43, 869−881.
Schneewind, O. Purification and Characterization of Sortase, the (76) Kharat, A. S.; Tomasz, A. Inactivation of the SrtA Gene Affects
Transpeptidase That Cleaves Surface Proteins of Staphylococcus aureus Localization of Surface Proteins and Decreases Adhesion of
at the LPXTG Motif. Proc. Natl. Acad. Sci. U. S. A. 1999, 96, 12424− Streptococcus pneumoniae to Human Pharyngeal Cells In Vitro. Infect.
12429. Immun. 2003, 71, 2758−2765.
(57) Bradshaw, W. J.; Davies, A. H.; Chambers, C. J.; Roberts, A. K.; (77) Hu, P.; Huang, P.; Chen, W. M. Curcumin Inhibits the Sortase A
Shone, C. C.; Acharya, K. R. Molecular Features of the Sortase Enzyme Activity of the Streptococcus mutans UA159. Appl. Biochem. Biotechnol.
Family. FEBS J. 2015, 282, 2097−2114. 2013, 171, 396−402.
(58) McCafferty, D. G.; Melvin, J. A. Sortases. In Handbook of (78) Lee, S. F.; Boran, T. L. Roles of Sortase in Surface Expression of
Proteolytic Enzymes; Elsevier, 2013, 2459−2465. DOI: 10.1016/B978- the Major Protein Adhesin P1, Saliva-Induced Aggregation and
0-12-382219-2.00549-4. Adherence, and Cariogenicity of Streptococcus mutans. Infect. Immun.
(59) Comfort, D.; Clubb, R. T. A Comparative Genome Analysis 2003, 71, 676−681.
Identifies Distinct Sorting Pathways in Gram-Positive Bacteria. Infect. (79) Vanier, G.; Sekizaki, T.; Domínguez-Punaro, M. C.; Esgleas, M.;
Immun. 2004, 72, 2710−2722. Osaki, M.; Takamatsu, D.; Segura, M.; Gottschalk, M. Disruption of

13126 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

SrtA Gene in Streptococcus suis Results in Decreased Interactions with (98) Williamson, D. J.; Fascione, M. A.; Webb, M. E.; Turnbull, W. B.
Endothelial Cells and Extracellular Matrix Proteins. Vet. Microbiol. Efficient N-Terminal Labeling of Proteins by Use of Sortase. Angew.
2008, 127, 417−424. Chem., Int. Ed. 2012, 51, 9377−9380.
(80) Cossart, P.; Jonquières, R. Sortase, a Universal Target for (99) Bellucci, J. J.; Bhattacharyya, J.; Chilkoti, A. A Noncanonical
Therapeutic Agents against Gram-Positive Bacteria? Proc. Natl. Acad. Function of Sortase Enables Site-Specific Conjugation of Small
Sci. U. S. A. 2000, 97, 5013−5015. Molecules to Lysine Residues in Proteins. Angew. Chem., Int. Ed.
(81) Ton-That, H.; Schneewind, O. Anchor Structure of Staph- 2015, 54, 441−445.
ylococcal Surface Proteins: IV. Inhibitors of the Cell Wall Sorting (100) Ling, J. J.; Policarpo, R. L.; Rabideau, A. E.; Liao, X.; Pentelute,
Reaction. J. Biol. Chem. 1999, 274, 24316−24320. B. L. Protein Thioester Synthesis Enabled by Sortase. J. Am. Chem. Soc.
(82) Ruzin, A.; Severin, A.; Ritacco, F.; Tabei, K.; Singh, G.; Bradford, 2012, 134, 10749−10752.
P. A.; Siegel, M. M.; Projan, S. J.; Shlaes, D. M. Further Evidence That a (101) Wu, Z.; Guo, X.; Gao, J.; Guo, Z. Sortase A-Mediated
Cell Wall Precursor [C55-MurNAc-(Peptide)-GlcNAc] Serves as an Chemoenzymatic Synthesis of Complex Glycosylphosphatidylinositol-
Acceptor in a Sorting Reaction. J. Bacteriol. 2002, 184, 2141−2147. Anchored Protein. Chem. Commun. 2013, 49, 11689−11691.
(83) Scheffers, D.-J.; Pinho, M. G. Bacterial Cell Wall Synthesis: New (102) Matayoshi, E.; Wang, G.; Krafft, G.; Erickson, J. Novel
Insights from Localization Studies. Microbiol. Mol. Biol. Rev. 2005, 69, Fluorogenic Substrates for Assaying Retroviral Proteases by Resonance
585−607. Energy Transfer. Science 1990, 247, 954−958.
(84) Ilangovan, U.; Ton-That, H.; Iwahara, J.; Schneewind, O.; Clubb, (103) Huang, X.; Aulabaugh, A.; Ding, W.; Kapoor, B.; Alksne, L.;
R. T. Structure of Sortase, the Transpeptidase That Anchors Proteins to Tabei, K.; Ellestad, G. Kinetic Mechanism of Staphylococcus aureus
the Cell Wall of Staphylococcus aureus. Proc. Natl. Acad. Sci. U. S. A. Sortase SrtA. Biochemistry 2003, 42, 11307−11315.
2001, 98, 6056−6061. (104) Nitulescu, G.; Nicorescu, I. M.; Olaru, O. T.; Ungurianu, A.;
(85) Suree, N.; Liew, C. K.; Villareal, V. A.; Thieu, W.; Fadeev, E. A.; Mihai, D. P.; Zanfirescu, A.; Nitulescu, G. M.; Margina, D. Molecular
Clemens, J. J.; Jung, M. E.; Clubb, R. T. The Structure of the Docking and Screening Studies of New Natural Sortase A Inhibitors.
Staphylococcus aureus Sortase-Substrate Complex Reveals How the Int. J. Mol. Sci. 2017, 18, 2217.
Universally Conserved LPXTG Sorting Signal Is Recognized. J. Biol. (105) Nitulescu, G.; Mihai, D. P.; Nicorescu, I. M.; Olaru, O. T.;
Chem. 2009, 284, 24465−24477. Ungurianu, A.; Zanfirescu, A.; Nitulescu, G. M.; Margina, D. Discovery
(86) Ton-That, H.; Mazmanian, S. K.; Alksne, L.; Schneewind, O. of Natural Naphthoquinones as Sortase A Inhibitors and Potential Anti-
Anchoring of Surface Proteins to the Cell Wall of Staphylococcus aureus: infective Solutions against Staphylococcus aureus. Drug Dev. Res. 2019,
Cysteine 184 and Histidine 120 of Sortase Form a Thiolate- 80, 1136−1145.
Imidazolium Ion Pair for Catalysis. J. Biol. Chem. 2002, 277, 7447− (106) Li, Y.; Yang, Y.; Zhang, C. Visualization and Quantification of
7452. Sortase Activity at the Single-Molecule Level via Transpeptidation-
(87) Frankel, B. A.; Tong, Y.; Bentley, M. L.; Fitzgerald, M. C.; Directed Intramolecular Förster Resonance Energy Transfer. Anal.
McCafferty, D. G. Mutational Analysis of Active Site Residues in the Chem. 2018, 90, 13007−13012.
Staphylococcus aureus Transpeptidase SrtA. Biochemistry 2007, 46, (107) Xu, A.; Yang, Y.; Zhang, C.-Y. Transpeptidation-Directed
7269−7278. Intramolecular Bipartite Tetracysteine Display for Sortase Activity
(88) Perry, A. M.; Ton-That, H.; Mazmanian, S. K.; Schneewind, O. Assay. Chem. Commun. 2018, 54, 8116−8119.
Anchoring of Surface Proteins to the Cell Wall of Staphylococcus aureus. (108) Kruger, R. G.; Dostal, P.; McCafferty, D. G. Development of a
III. Lipid II Is an In Vivo Peptidoglycan Substrate for Sortase-Catalyzed High-Performance Liquid Chromatography Assay and Revision of
Surface Protein Anchoring. J. Biol. Chem. 2002, 277, 16241−16248. Kinetic Parameters for the Staphylococcus aureus Sortase Trans-
(89) Liew, C. K.; Smith, B. T.; Pilpa, R.; Suree, N.; Ilangovan, U.; peptidase SrtA. Anal. Biochem. 2004, 326, 42−48.
Connolly, K. M.; Jung, M. E.; Clubb, R. T. Localization and (109) Aulabaugh, A.; Ding, W.; Kapoor, B.; Tabei, K.; Alksne, L.;
Mutagenesis of the Sorting Signal Binding Site on Sortase A from Dushin, R.; Zatz, T.; Ellestad, G.; Huang, X. Development of an HPLC
Staphylococcus aureus. FEBS Lett. 2004, 571, 221−226. Assay for Staphylococcus aureus Sortase: Evidence for the Formation of
(90) Zong, Y.; Bice, T. W.; Ton-That, H.; Schneewind, O.; Narayana, the Kinetically Competent Acyl Enzyme Intermediate. Anal. Biochem.
S. V. L. Crystal Structures of Staphylococcus aureus Sortase A and Its 2007, 360, 14−22.
Substrate Complex. J. Biol. Chem. 2004, 279, 31383−31389. (110) Wolz, C.; McDevitt, D.; Foster, T. J.; Cheung, A. L. Influence of
(91) Bentley, M. L.; Lamb, E. C.; McCafferty, D. G. Mutagenesis Agr on Fibrinogen Binding in Staphylococcus aureus Newman. Infect.
Studies of Substrate Recognition and Catalysis in the Sortase A Immun. 1996, 64, 3142−3147.
Transpeptidase from Staphylococcus aureus. J. Biol. Chem. 2008, 283, (111) Elgalai, I.; Foster, H. A. Comparison of Adhesion of Wound
14762−14771. Isolates of Staphylococcus aureus to Immobilized Proteins. J. Appl.
(92) Gao, C.; Uzelac, I.; Gottfries, J.; Eriksson, L. A. Exploration of Microbiol. 2003, 94, 413−420.
Multiple Sortase A Protein Conformations in Virtual Screening. Sci. (112) Zhang, J.; Liu, H.; Zhu, K.; Gong, S.; Dramsi, S.; Wang, Y.-T.;
Rep. 2016, 6, 20413. Li, J.; Chen, F.; Zhang, R.; Zhou, L.; Lan, L.; Jiang, H.; Schneewind, O.;
(93) Kappel, K.; Wereszczynski, J.; Clubb, R. T.; McCammon, J. A. Luo, C.; Yang, C.-G. Antiinfective Therapy with a Small Molecule
The Binding Mechanism, Multiple Binding Modes, and Allosteric Inhibitor of Staphylococcus aureus Sortase. Proc. Natl. Acad. Sci. U. S. A.
Regulation of Staphylococcus aureus Sortase A Probed by Molecular 2014, 111, 13517−13522.
Dynamics Simulations: Staphylococcus aureus SrtA Probed by MD (113) Yang, T.; Zhang, T.; Guan, X.; Dong, Z.; Lan, L.-F.; Yang, S.;
Simulations. Protein Sci. 2012, 21, 1858−1871. Yang, C.-G. Tideglusib and Its Analogues As Inhibitors of Staph-
(94) Tsukiji, S.; Nagamune, T. Sortase-Mediated Ligation: A Gift ylococcus aureus SrtA. J. Med. Chem. 2020, 63, 8442−8457.
from Gram-Positive Bacteria to Protein Engineering. ChemBioChem (114) Rebollo, I. R.; McCallin, S.; Bertoldo, D.; Entenza, J. M.;
2009, 10, 787−798. Moreillon, P.; Heinis, C. Development of Potent and Selective S. aureus
(95) Haridas, V.; Sadanandan, S.; Dheepthi, N. U. Sortase-Based Bio- Sortase A Inhibitors Based on Peptide Macrocycles. ACS Med. Chem.
Organic Strategies for Macromolecular Synthesis. ChemBioChem 2014, Lett. 2016, 7, 606−611.
15, 1857−1867. (115) Hsu, Y.-P.; Hall, E.; Booher, G.; Murphy, B.; Radkov, A. D.;
(96) Gupta, K.; Singh, S.; Gupta, K.; Khan, N.; Sehgal, D.; Haridas, V.; Yablonowski, J.; Mulcahey, C.; Alvarez, L.; Cava, F.; Brun, Y. V.; Kuru,
Roy, R. P. A Bioorthogonal Chemoenzymatic Strategy for Defined E.; VanNieuwenhze, M. S. Fluorogenic D-Amino Acids Enable Real-
Protein Dendrimer Assembly. ChemBioChem 2012, 13, 2489−2494. Time Monitoring of Peptidoglycan Biosynthesis and High-Throughput
(97) Sapra, R.; Verma, R. P.; Maurya, G. P.; Dhawan, S.; Babu, J.; Transpeptidation Assays. Nat. Chem. 2019, 11, 335−341.
Haridas, V. Designer Peptide and Protein Dendrimers: A Cross- (116) Scott, C. J.; McDowell, A.; Martin, S. L.; Lynas, J. F.;
Sectional Analysis. Chem. Rev. 2019, 119, 11391−11441. Vandenbroeck, K.; Walker, B. Irreversible Inhibition of the Bacterial

13127 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

Cysteine Protease-Transpeptidase Sortase (SrtA) by Substrate-Derived (137) Jeon, J.; Na, Z.; Jung, M.; Lee, H.-S.; Sim, C. J.; Nahm, K.; Oh,
Affinity Labels. Biochem. J. 2002, 366, 953−958. K.-B.; Shin, J. Discorhabdins from the Korean Marine Sponge Sceptrella
(117) Kim, S.-W.; Chang, I.-M.; Oh, K.-B. Inhibition of the Bacterial sp. J. Nat. Prod. 2010, 73, 258−262.
Surface Protein Anchoring Transpeptidase Sortase by Medicinal Plants. (138) Bae, J.; Jeon, J.; Lee, Y.-J.; Lee, H.-S.; Sim, C. J.; Oh, K.-B.; Shin,
Biosci., Biotechnol., Biochem. 2002, 66, 2751−2754. J. Sesterterpenes from the Tropical Sponge Coscinoderma sp. J. Nat.
(118) Baell, J. B.; Holloway, G. A. New Substructure Filters for Prod. 2011, 74, 1805−1811.
Removal of Pan Assay Interference Compounds (PAINS) from (139) Won, T. H.; Jeon, J.; Lee, S.-H.; Rho, B. J.; Oh, K.-B.; Shin, J.
Screening Libraries and for Their Exclusion in Bioassays. J. Med. Beta-Carboline Alkaloids Derived from the Ascidian Synoicum sp.
Chem. 2010, 53, 2719−2740. Bioorg. Med. Chem. 2012, 20, 4082−4087.
(119) Dahlin, J. L.; Nissink, J. W. M.; Strasser, J. M.; Francis, S.; (140) Won, T. H.; Jeon, J.; Kim, S.-H.; Lee, S.-H.; Rho, B. J.; Oh, D.-
Higgins, L.; Zhou, H.; Zhang, Z.; Walters, M. A. PAINS in the Assay: C.; Oh, K.-B.; Shin, J. Brominated Aromatic Furanones and Related
Chemical Mechanisms of Assay Interference and Promiscuous Esters from the Ascidian Synoicum sp. J. Nat. Prod. 2012, 75, 2055−
Enzymatic Inhibition Observed during a Sulfhydryl-Scavenging HTS. 2061.
J. Med. Chem. 2015, 58, 2091−2113. (141) Bae, J.; Cho, E.; Park, J. S.; Won, T. H.; Seo, S.-Y.; Oh, D.-C.;
(120) Baell, J.; Walters, M. A. Chemistry: Chemical Con Artists Foil Oh, K.-B.; Shin, J. Isocadiolides A-H: Polybrominated Aromatics from a
Drug Discovery. Nature 2014, 513, 481−483. Synoicum sp. Ascidian. J. Nat. Prod. 2020, 83, 429−437.
(121) Baell, J. B.; Nissink, J. W. M. Seven Year Itch: Pan-Assay (142) Julianti, E.; Lee, J.-H.; Liao, L.; Park, W.; Park, S.; Oh, D.-C.;
Interference Compounds (PAINS) in 2017  Utility and Limitations. Oh, K.-B.; Shin, J. New Polyaromatic Metabolites from a Marine-
ACS Chem. Biol. 2018, 13, 36−44. Derived Fungus Penicillium sp. Org. Lett. 2013, 15, 1286−1289.
(122) Trabocchi, A. Principles and Applications of Small Molecule (143) Chung, B.; Kwon, O.-S.; Shin, J.; Oh, K.-B. Inhibitory Effects of
Peptidomimetics. In Small Molecule Drug Discovery; Elsevier, 2020, Streptomyces sp. MBTH32 Metabolites on Sortase A and Sortase A-
163−195. DOI: 10.1016/B978-0-12-818349-6.00006-6. Mediated Cell Clumping of Staphylococcus aureus to Fibrinogen. J.
(123) Lenci, E.; Trabocchi, A. Peptidomimetic Toolbox for Drug Microbiol. Biotechnol. 2019, 29, 1603−1606.
Discovery. Chem. Soc. Rev. 2020, 49, 3262−3277. (144) Park, S. C.; Chung, B.; Lee, J.; Cho, E.; Hwang, J.-Y.; Oh, D.-C.;
(124) Connolly, K. M.; Smith, B. T.; Pilpa, R.; Ilangovan, U.; Jung, M. Shin, J.; Oh, K.-B. Sortase A-Inhibitory Metabolites from a Marine-
E.; Clubb, R. T. Sortase from Staphylococcus aureus Does Not Contain a Derived Fungus Aspergillus sp. Mar. Drugs 2020, 18, 359.
Thiolate-Imidazolium Ion Pair in Its Active Site. J. Biol. Chem. 2003, (145) Che, C.-T.; Zhang, H. Plant Natural Products for Human
278, 34061−34065. Health. Int. J. Mol. Sci. 2019, 20, 830.
(125) Kruger, R. G.; Barkallah, S.; Frankel, B. A.; McCafferty, D. G. (146) Silva, L. N.; Zimmer, K. R.; Macedo, A. J.; Trentin, D. S. Plant
Inhibition of the Staphylococcus aureus Sortase Transpeptidase SrtA by Natural Products Targeting Bacterial Virulence Factors. Chem. Rev.
Phosphinic Peptidomimetics. Bioorg. Med. Chem. 2004, 12, 3723−
2016, 116, 9162−9236.
3729.
(147) Kim, S.-H.; Shin, D.-S.; Oh, M.-N.; Chung, S.-C.; Lee, J.-S.; Oh,
(126) Jung, M. E.; Clemens, J. J.; Suree, N.; Liew, C. K.; Pilpa, R.;
K.-B. Inhibition of the Bacterial Surface Protein Anchoring Trans-
Campbell, D. O.; Clubb, R. T. Synthesis of (2R,3S) 3-Amino-4-
peptidase Sortase by Isoquinoline Alkaloids. Biosci., Biotechnol.,
mercapto-2-butanol, a Threonine Analogue for Covalent Inhibition of
Biochem. 2004, 68, 421−424.
Sortases. Bioorg. Med. Chem. Lett. 2005, 15, 5076−5079.
(148) Park, B.-S.; Kim, J.-G.; Kim, M.-R.; Lee, S.-E.; Takeoka, G. R.;
(127) Wang, J.; Li, H.; Pan, J.; Dong, J.; Zhou, X.; Niu, X.; Deng, X.
Oh, K.-B.; Kim, J.-H. Curcuma Longa L. Constituents Inhibit Sortase A
Oligopeptide Targeting Sortase A as Potential Anti-Infective Therapy
for Staphylococcus aureus. Front. Microbiol. 2018, 9, 245. and Staphylococcus aureus Cell Adhesion to Fibronectin. J. Agric. Food
(128) Dai, J.; Han, R.; Xu, Y.; Li, N.; Wang, J.; Dan, W. Recent Chem. 2005, 53, 9005−9009.
Progress of Antibacterial Natural Products: Future Antibiotics (149) Niu, X.; Gao, Y.; Yu, Y.; Yang, Y.; Wang, G.; Sun, L.; Wang, H.
Candidates. Bioorg. Chem. 2020, 101, 103922. Molecular Modelling Reveals the Inhibition Mechanism and Structure-
(129) Blunt, J. W.; Carroll, A. R.; Copp, B. R.; Davis, R. A.; Keyzers, R. Activity Relationship of Curcumin and Its Analogues to Staphylococcal
A.; Prinsep, M. R. Marine Natural Products. Nat. Prod. Rep. 2018, 35, aureus Sortase A. J. Biomol. Struct. Dyn. 2019, 37, 1220−1230.
8−53. (150) van Houte, J. Role of Micro-Organisms in Caries Etiology. J.
(130) Donia, M.; Hamann, M. T. Marine Natural Products and Their Dent. Res. 1994, 73, 672−681.
Potential Applications as Anti-Infective Agents. Lancet Infect. Dis. 2003, (151) Socransky, S. S.; Haffajee, A. D. Dental Biofilms: Difficult
3, 338−348. Therapeutic Targets. Periodontol. 2000 2002, 28, 12−55.
(131) Tortorella, E.; Tedesco, P.; Palma Esposito, F.; January, G.; (152) Marsh, P. D. Dental Plaque as a Biofilm and a Microbial
Fani, R.; Jaspars, M.; de Pascale, D. Antibiotics from Deep-Sea Community - Implications for Health and Disease. BMC Oral Health
Microorganisms: Current Discoveries and Perspectives. Mar. Drugs 2006, 6, S14.
2018, 16, 355. (153) Igarashi, I.; Asaga, E.; Goto, N. The Sortase of Streptococcus
(132) Carroll, A. R.; Copp, B. R.; Davis, R. A.; Keyzers, R. A.; Prinsep, mutans Mediates Cell Wall Anchoring of a Surface Protein Antigen.
M. R. Marine Natural Products. Nat. Prod. Rep. 2019, 36, 122−173. Oral Microbiol. Immunol. 2003, 18, 266−269.
(133) Oh, K.-B.; Mar, W.; Kim, S.; Kim, J.-Y.; Oh, M.-N.; Kim, J.-G.; (154) Russell, M. W.; Mansson-Rahemtulla, B. Interaction Between
Shin, D.; Sim, C. J.; Shin, J. Bis(Indole) Alkaloids as Sortase A Surface Protein Antigen of Streptococcus mutans and Human Salivary
Inhibitors from the Sponge Spongosorites sp. Bioorg. Med. Chem. Lett. Components. Oral Microbiol. Immunol. 1989, 4, 106−111.
2005, 15, 4927−4931. (155) Scharnow, A. M.; Solinski, A. E.; Wuest, W. M. Targeting S.
(134) Oh, K.-B.; Oh, M.-N.; Kim, J.-G.; Shin, D.-S.; Shin, J. Inhibition mutans Biofilms: A Perspective on Preventing Dental Caries.
of Sortase-Mediated Staphylococcus aureus Adhesion to Fibronectin via MedChemComm 2019, 10, 1057−1067.
Fibronectin-Binding Protein by Sortase Inhibitors. Appl. Microbiol. (156) Lévesque, C. M.; Voronejskaia, E.; Huang, Y.-C. C.; Mair, R.
Biotechnol. 2006, 70, 102−106. W.; Ellen, R. P.; Cvitkovitch, D. G. Involvement of Sortase Anchoring of
(135) Jang, K. H.; Chung, S.-C.; Shin, J.; Lee, S.-H.; Kim, T.-I.; Lee, Cell Wall Proteins in Biofilm Formation by Streptococcus mutans. Infect.
H.-S.; Oh, K.-B. Aaptamines as Sortase A Inhibitors from the Tropical Immun. 2005, 73, 3773−3777.
Sponge Aaptos aaptos. Bioorg. Med. Chem. Lett. 2007, 17, 5366−5369. (157) Hu, P.; Huang, P.; Chen, M. W. Curcumin Reduces
(136) Lv, Z.; Wang, H. S.; Niu, X. D. Molecular Dynamics Simulations Streptococcus mutans Biofilm Formation by Inhibiting Sortase A
Reveal Insight into Key Structural Elements of Aaptamines as Sortase Activity. Arch. Oral Biol. 2013, 58, 1343−1348.
Inhibitors with Free Energy Calculations. Chem. Phys. Lett. 2013, 585, (158) Karak, P. Biological Activities of Flavonoids: An Overview. Int. J.
171−177. Pharm. Sci. Res. 2019, 10, 1567−1574.

13128 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

(159) Kang, S. S.; Kim, J.-G.; Lee, T.-H.; Oh, K.-B. Flavonols Inhibit (177) Mu, D.; Luan, Y.; Wang, L.; Gao, Z.; Yang, P.; Jing, S.; Wang, Y.;
Sortases and Sortase-Mediated Staphylococcus aureus Clumping to Xiang, H.; Wang, T.; Wang, D. The Combination of Salvianolic Acid A
Fibrinogen. Biol. Pharm. Bull. 2006, 29, 1751−1755. with Latamoxef Completely Protects Mice against Lethal Pneumonia
(160) Huang, P.; Hu, P.; Zhou, S. Y.; Li, Q.; Chen, W. M. Morin Caused by Methicillin-Resistant Staphylococcus aureus. Emerging
Inhibits Sortase A and Subsequent Biofilm Formation in Streptococcus Microbes Infect. 2020, 9, 169−179.
mutans. Curr. Microbiol. 2014, 68, 47−52. (178) Park, J.-S.; Chung, B.; Lee, W.-H.; Lee, J.; Suh, Y.; Oh, D.-C.;
(161) Liu, B.; Chen, F.; Bi, C.; Wang, L.; Zhong, X.; Cai, H.; Deng, X.; Oh, K.-B.; Shin, J. Sortase A-Inhibitory Coumarins from the Folk
Niu, X.; Wang, D. Quercitrin, an Inhibitor of Sortase A, Interferes with Medicinal Plant Poncirus trifoliata. J. Nat. Prod. 2020, 83, 3004−3011.
the Adhesion of Staphylococcal aureus. Molecules 2015, 20, 6533−6543. (179) Palmer, J. T.; Rasnick, D.; Klaus, J. L.; Bromme, D. Vinyl
(162) Silva, L. N.; Da Hora, G. C. A.; Soares, T. A.; Bojer, M. S.; Sulfones as Mechanism-Based Cysteine Protease Inhibitors. J. Med.
Ingmer, H.; Macedo, A. J.; Trentin, D. S. Myricetin Protects Galleria Chem. 1995, 38, 3193−3196.
Mellonella against Staphylococcus aureus Infection and Inhibits Multiple (180) Frankel, B. A.; Bentley, M.; Kruger, R. G.; McCafferty, D. G.
Virulence Factors. Sci. Rep. 2017, 7, 2823. Vinyl Sulfones: Inhibitors of SrtA, a Transpeptidase Required for Cell
(163) Oh, I.; Yang, W.-Y.; Chung, S.-C.; Kim, T.-Y.; Oh, K.-B.; Shin, J. Wall Protein Anchoring and Virulence in Staphylococcus aureus. J. Am.
In Vitro Sortase A Inhibitory and Antimicrobial Activity of Flavonoids Chem. Soc. 2004, 126, 3404−3405.
Isolated from the Roots of Sophora f lavescens. Arch. Pharmacal Res. (181) Oh, K.-B.; Kim, S.-H.; Lee, J.; Cho, W.-J.; Lee, T.; Kim, S.
2011, 34, 217−222. Discovery of Diarylacrylonitriles as a Novel Series of Small Molecule
(164) Bi, C.; Dong, X.; Zhong, X.; Cai, H.; Wang, D.; Wang, L. Sortase A Inhibitors. J. Med. Chem. 2004, 47, 2418−2421.
Acacetin Protects Mice from Staphylococcus aureus Bloodstream (182) Oh, K.-B.; Nam, K.-W.; Ahn, H.; Shin, J.; Kim, S.; Mar, W.
Infection by Inhibiting the Activity of Sortase A. Molecules 2016, 21, Therapeutic Effect of (Z)-3-(2,5-Dimethoxyphenyl)-2-(4-Methoxy-
1285. phenyl) Acrylonitrile (DMMA) against Staphylococcus aureus Infection
(165) Zhang, B.; Wang, X.; Wang, L.; Chen, S.; Shi, D.; Wang, H. in a Murine Model. Biochem. Biophys. Res. Commun. 2010, 396, 440−
Molecular Mechanism of the Flavonoid Natural Product Dryocrassin 444.
ABBA against Staphylococcus aureus Sortase A. Molecules 2016, 21, (183) Kudryavtsev, K. V.; Bentley, M. L.; McCafferty, D. G. Probing of
1428. the Cis-5-Phenyl Proline Scaffold as a Platform for the Synthesis of
(166) Cho, H.; Chung, B.; Kim, C.-K.; Oh, D.-C.; Oh, K.-B.; Shin, J. Mechanism-Based Inhibitors of the Staphylococcus aureus Sortase SrtA
Spatholobus suberectus Dunn. Constituents Inhibit Sortase A and Isoform. Bioorg. Med. Chem. 2009, 17, 2886−2893.
Staphylococcus aureus Cell Clumping to Fibrinogen. Arch. Pharmacal (184) Maresso, A. W.; Wu, R.; Kern, J. W.; Zhang, R.; Janik, D.;
Res. 2017, 40, 518−523. Missiakas, D. M.; Duban, M.-E.; Joachimiak, A.; Schneewind, O.
(167) Mu, D.; Xiang, H.; Dong, H.; Wang, D.; Wang, T. Isovitexin, a Activation of Inhibitors by Sortase Triggers Irreversible Modification of
Potential Candidate Inhibitor of Sortase A of Staphylococcus aureus the Active Site. J. Biol. Chem. 2007, 282, 23129−23139.
USA300. J. Microbiol. Biotechnol. 2018, 28, 1426−1432. (185) Suree, N.; Yi, S. W.; Thieu, W.; Marohn, M.; Damoiseaux, R.;
(168) Tritsch, D.; Zinglé, C.; Rohmer, M.; Grosdemange-Billiard, C. Chan, A.; Jung, M. E.; Clubb, R. T. Discovery and Structure-Activity
Flavonoids: True or Promiscuous Inhibitors of Enzyme? The Case of Relationship Analysis of Staphylococcus aureus Sortase A Inhibitors.
Deoxyxylulose Phosphate Reductoisomerase. Bioorg. Chem. 2015, 59, Bioorg. Med. Chem. 2009, 17, 7174−7185.
140−144. (186) Uddin, R.; Lodhi, M. U.; Ul-Haq, Z. Combined Pharmacophore
(169) Kim, S.-H.; Shin, D.-S.; Oh, M.-N.; Chung, S.-C.; Lee, J.-S.; and 3D-QSAR Study on A Series of Staphylococcus aureus Sortase A
Chang, I.-M.; Oh, K.-B. Inhibition of Sortase, a Bacterial Surface Inhibitors. Chem. Biol. Drug Des. 2012, 80, 300−314.
Protein Anchoring Transpeptidase, by β-Sitosterol-3-O -Glucopyrano- (187) Chan, A. H.; Yi, S. W.; Weiner, E. M.; Amer, B. R.; Sue, C. K.;
side from Fritillaria verticillata. Biosci., Biotechnol., Biochem. 2003, 67, Wereszczynski, J.; Dillen, C. A.; Senese, S.; Torres, J. Z.; McCammon, J.
2477−2479. A.; Miller, L. S.; Jung, M. E.; Clubb, R. T. NMR Structure-Based
(170) Wang, L.; Bi, C.; Cai, H.; Liu, B.; Zhong, X.; Deng, X.; Wang, T.; Optimization of Staphylococcus aureus Sortase A Pyridazinone
Xiang, H.; Niu, X.; Wang, D. The Therapeutic Effect of Chlorogenic Inhibitors. Chem. Biol. Drug Des. 2017, 90, 327−344.
Acid against Staphylococcus aureus Infection through Sortase A (188) Chenna, B. C.; Shinkre, B. A.; King, J. R.; Lucius, A. L.;
Inhibition. Front. Microbiol. 2015, 6, 1031. Narayana, S. V. L.; Velu, S. E. Identification of Novel Inhibitors of
(171) Bi, C.; Wang, L.; Niu, X.; Cai, H.; Zhong, X.; Deng, X.; Wang, Bacterial Surface Enzyme Staphylococcus aureus Sortase A. Bioorg. Med.
T.; Wang, D. The Use of Chlorogenic Acid and Its Analogues as Chem. Lett. 2008, 18, 380−385.
Inhibitors: An Investigation of the Inhibition of Sortase A of (189) Chenna, B. C.; King, J. R.; Shinkre, B. A.; Glover, A. L.; Lucius,
Staphylococcus aureus Using Molecular Docking and Dynamic A. L.; Velu, S. E. Synthesis and Structure Activity Relationship Studies
Simulation. Biotechnol. Lett. 2016, 38, 1341−1347. of Novel Staphylococcus aureus Sortase A Inhibitors. Eur. J. Med. Chem.
(172) Lee, S.; Song, I.-H.; Lee, J.-H.; Yang, W.-Y.; Oh, K.-B.; Shin, J. 2010, 45, 3752−3761.
Sortase A Inhibitory Metabolites from the Roots of Pulsatilla koreana. (190) Lee, H.-S.; Yoon, K.-M.; Han, Y.-R.; Lee, K. J.; Chung, S.-C.;
Bioorg. Med. Chem. Lett. 2014, 24, 44−48. Kim, T.-I.; Lee, S.-H.; Shin, J.; Oh, K.-B. 5-Hydroxyindole-Type
(173) Wu, Y.-P.; Liu, X.-Y.; Bai, J.-R.; Xie, H.-C.; Ye, S.-L.; Zhong, K.; Alkaloids, as Candida albicans Isocitrate Lyase Inhibitors, from the
Huang, Y.-N.; Gao, H. Inhibitory Effect of a Natural Phenolic Tropical Sponge Hyrtios sp. Bioorg. Med. Chem. Lett. 2009, 19, 1051−
Compound, 3-p-Trans-Coumaroyl-2-Hydroxyquinic Acid against the 1053.
Attachment Phase of Biofilm Formation of Staphylococcus aureus (191) Lee, Y.-J.; Han, Y.-R.; Park, W.; Nam, S.-H.; Oh, K.-B.; Lee, H.-
through Targeting Sortase A. RSC Adv. 2019, 9, 32453−32461. S. Synthetic Analogs of Indole-Containing Natural Products as
(174) Hou, X.; Wang, M.; Wen, Y.; Ni, T.; Guan, X.; Lan, L.; Zhang, Inhibitors of Sortase A and Isocitrate Lyase. Bioorg. Med. Chem. Lett.
N.; Zhang, A.; Yang, C.-G. Quinone Skeleton as a New Class of 2010, 20, 6882−6885.
Irreversible Inhibitors against Staphylococcus aureus Sortase A. Bioorg. (192) Zhulenkovs, D.; Rudevica, Z.; Jaudzems, K.; Turks, M.;
Med. Chem. Lett. 2018, 28, 1864−1869. Leonchiks, A. Discovery and Structure-Activity Relationship Studies of
(175) Ouyang, P.; He, X.; Yuan, Z.-W.; Yin, Z.-Q.; Fu, H.; Lin, J.; He, Irreversible Benzisothiazolinone-Based Inhibitors against Staphylococ-
C.; Liang, X.; Lv, C.; Shu, G.; Yuan, Z.-X.; Song, X.; Li, L.; Yin, L. cus aureus Sortase A Transpeptidase. Bioorg. Med. Chem. 2014, 22,
Erianin against Staphylococcus aureus Infection via Inhibiting Sortase A. 5988−6003.
Toxins 2018, 10, 385. (193) Jaudzems, K.; Kurbatska, V.; Jekabsons, A.; Bobrovs, R.;
(176) Zhang, B.; Teng, Z.; Li, X.; Lu, G.; Deng, X.; Niu, X.; Wang, J. Rudevica, Z.; Leonchiks, A. Targeting Bacterial Sortase A with Covalent
Chalcone Attenuates Staphylococcus aureus Virulence by Targeting Inhibitors: 27 New Starting Points for Structure-Based Hit-to-Lead
Sortase A and Alpha-Hemolysin. Front. Microbiol. 2017, 8, 1715. Optimization. ACS Infect. Dis. 2020, 6, 186−194.

13129 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130
Journal of Medicinal Chemistry pubs.acs.org/jmc Perspective

(194) Chan, A. H.; Wereszczynski, J.; Amer, B. R.; Yi, S. W.; Jung, M.
E.; McCammon, J. A.; Clubb, R. T. Discovery of Staphylococcus aureus
Sortase A Inhibitors Using Virtual Screening and the Relaxed Complex
Scheme. Chem. Biol. Drug Des. 2013, 82, 418−428.
(195) Maggio, B.; Raffa, D.; Raimondi, M. V.; Cascioferro, S.; Plescia,
F.; Schillaci, D.; Cusimano, M. G.; Leonchiks, A.; Zhulenkovs, D.;
Basile, L.; Daidone, G. Discovery of a New Class of Sortase A
Transpeptidase Inhibitors to Tackle Gram-Positive Pathogens: 2-(2-
Phenylhydrazinylidene)Alkanoic Acids and Related Derivatives.
Molecules 2016, 21, 241.
(196) Cascioferro, S.; Maggio, B.; Raffa, D.; Raimondi, M. V.;
Cusimano, M. G.; Schillaci, D.; Manachini, B.; Leonchiks, A.; Daidone,
G. A New Class of Phenylhydrazinylidene Derivatives as Inhibitors of
Staphylococcus aureus Biofilm Formation. Med. Chem. Res. 2016, 25,
870−878.
(197) Zhang, Y.; Bao, J.; Deng, X.-X.; He, W.; Fan, J.-J.; Jiang, F.-Q.;
Fu, L. Synthesis, Biological Evaluation and Molecular Docking of 2-
Phenyl-Benzo[d]Oxazole-7-Carboxamide Derivatives as Potential
Staphylococcus aureus Sortase A Inhibitors. Bioorg. Med. Chem. Lett.
2016, 26, 4081−4085.
(198) He, W.; Zhang, Y.; Bao, J.; Deng, X.; Batara, J.; Casey, S.; Guo,
Q.; Jiang, F.; Fu, L. Synthesis, Biological Evaluation and Molecular
Docking Analysis of 2-Phenyl-Benzofuran-3-Carboxamide Derivatives
as Potential Inhibitors of Staphylococcus aureus Sortase A. Bioorg. Med.
Chem. 2017, 25, 1341−1351.
(199) Wehrli, P. M.; Uzelac, I.; Olsson, T.; Jacso, T.; Tietze, D.;
Gottfries, J. Discovery and Development of Substituted Thiadiazoles as
Inhibitors of Staphylococcus aureus Sortase A. Bioorg. Med. Chem. 2019,
27, 115043.
(200) Domínguez, J. M.; Fuertes, A.; Orozco, L.; del Monte-Millán,
M.; Delgado, E.; Medina, M. Evidence for Irreversible Inhibition of
Glycogen Synthase Kinase-3β by Tideglusib. J. Biol. Chem. 2012, 287,
893−904.
(201) Parrino, B.; Carbone, D.; Cascioferro, S.; Pecoraro, C.;
Giovannetti, E.; Deng, D.; Di Sarno, V.; Musella, S.; Auriemma, G.;
Cusimano, M. G.; Schillaci, D.; Cirrincione, G.; Diana, P. 1,2,4-
Oxadiazole Topsentin Analogs as Staphylococcal Biofilm Inhibitors
Targeting the Bacterial Transpeptidase Sortase A. Eur. J. Med. Chem.
2021, 209, 112892.
(202) Barthels, F.; Marincola, G.; Marciniak, T.; Konhäuser, M.;
Hammerschmidt, S.; Bierlmeier, J.; Distler, U.; Wich, P. R.; Tenzer, S.;
Schwarzer, D.; Ziebuhr, W.; Schirmeister, T. Asymmetric Disulfanyl-
benzamides as Irreversible and Selective Inhibitors of Staphylococcus
aureus Sortase A. ChemMedChem 2020, 15, 839−850.
(203) Dickey, S. W.; Cheung, G. Y. C.; Otto, M. Different Drugs for
Bad Bugs: Antivirulence Strategies in the Age of Antibiotic Resistance.
Nat. Rev. Drug Discovery 2017, 16, 457−471.
(204) Cragg, G. M.; Newman, D. J.; Snader, K. M. Natural Products in
Drug Discovery and Development. J. Nat. Prod. 1997, 60, 52−60.
(205) Breinbauer, R.; Vetter, I. R.; Waldmann, H. From Protein
Domains to Drug Candidates-Natural Products as Guiding Principles
in the Design and Synthesis of Compound Libraries. Angew. Chem., Int.
Ed. 2002, 41, 2878−2890.
(206) Wu, S.-C.; Liu, F.; Zhu, K.; Shen, J.-Z. Natural Products That
Target Virulence Factors in Antibiotic-Resistant Staphylococcus. aureus.
J. Agric. Food Chem. 2019, 67, 13195−13211.

13130 https://doi.org/10.1021/acs.jmedchem.1c00386
J. Med. Chem. 2021, 64, 13097−13130

You might also like