Next Article in Journal
Comparative Meropenem Pharmacodynamics and Emergence of Resistance against Carbapenem-Susceptible Non-Carbapenemase-Producing and Carbapenemase-Producing Enterobacterales: A Pharmacodynamic Study in a Hollow-Fiber Infection Model
Next Article in Special Issue
Antimicrobial Prophylaxis in Robot-Assisted Laparoscopic Radical Prostatectomy: A Systematic Review
Previous Article in Journal
Antibiotic Potentiation as a Promising Strategy to Combat Macrolide Resistance in Bacterial Pathogens
Previous Article in Special Issue
Primary Care Antibiotic Prescribing and Infection-Related Hospitalisation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Lipid-Centric Approaches in Combating Infectious Diseases: Antibacterials, Antifungals and Antivirals with Lipid-Associated Mechanisms of Action

by
Olga S. Ostroumova
* and
Svetlana S. Efimova
Laboratory of Membrane and Ion Channel Modeling, Institute of Cytology, Russian Academy of Sciences, Tikhoretsky Ave. 4, St. Petersburg 194064, Russia
*
Author to whom correspondence should be addressed.
Antibiotics 2023, 12(12), 1716; https://doi.org/10.3390/antibiotics12121716
Submission received: 31 October 2023 / Revised: 5 December 2023 / Accepted: 8 December 2023 / Published: 11 December 2023
(This article belongs to the Special Issue Antibiotics Use in Infection and Public Health)

Abstract

:
One of the global challenges of the 21st century is the increase in mortality from infectious diseases against the backdrop of the spread of antibiotic-resistant pathogenic microorganisms. In this regard, it is worth targeting antibacterials towards the membranes of pathogens that are quite conservative and not amenable to elimination. This review is an attempt to critically analyze the possibilities of targeting antimicrobial agents towards enzymes involved in pathogen lipid biosynthesis or towards bacterial, fungal, and viral lipid membranes, to increase the permeability via pore formation and to modulate the membranes’ properties in a manner that makes them incompatible with the pathogen’s life cycle. This review discusses the advantages and disadvantages of each approach in the search for highly effective but nontoxic antimicrobial agents. Examples of compounds with a proven molecular mechanism of action are presented, and the types of the most promising pharmacophores for further research and the improvement of the characteristics of antibiotics are discussed. The strategies that pathogens use for survival in terms of modulating the lipid composition and physical properties of the membrane, achieving a balance between resistance to antibiotics and the ability to facilitate all necessary transport and signaling processes, are also considered.

1. Introduction

Here, we provide an overview of antibacterial, antifungal, and antiviral agents that target lipid biosynthesis and modulate the properties of pathogen membranes, including pore formation, the induction of curvature stress, and full disruption. Taking into account the emphasis on the lipid-associated mechanisms of action of potent antimicrobial drugs, to compare various agents that inhibit lipid biosynthesis, the concentrations causing a twofold decrease in the activity of appropriate enzymes are presented. The threshold concentration in the membrane bathing solution is chosen to quantitatively characterize the effectiveness of various pore-forming antibiotics. The modulation of the properties of the host cell plasma membranes, targeting enzymes engaged in the regulation of lipid metabolism and the biosynthesis of pathogens’ cell wall components, are not reviewed.

2. Antibacterials with Lipid-Associated Mechanisms of Action

We focus on two modes of targeting antibacterial agents towards pathogen membranes: (i) indirect action via inhibition of the biosynthesis of membrane lipids; (ii) primary interaction with lipids, which results in the disruption of the functioning of bacterial membranes. These fundamentally different possibilities are considered below.

2.1. Inhibitors of Membrane Lipid Biosynthesis in Bacteria

Despite the fact that bacterial cell wall biosynthesis inhibitors, especially β-lactams and glycopeptide antibiotics, inhibiting the synthesis of the peptidoglycan layer, are the most effective and extensively used classes of antibiotics [1,2], they are not covered in this work, which focuses on targeting the lipid membrane. This part of the review concerns the key enzymes in the biosynthesis of membrane lipids in bacteria and their inhibitors. Possible ways to regulate the biosynthesis, transport, and degradation of lipids are not considered.

2.1.1. Biosynthesis of Fatty Acids of Bacterial Membrane Lipids

The search for selective inhibitors of enzymes participating in bacterial pathways for lipid biosynthesis is a good strategy to find novel antibiotics due to the fact that a certain lipid composition of the bacterial membrane is necessary for its proper functioning, and there is a difference in the principal organization of lipid biosynthesis in bacteria and mammals. In particular, the membrane’s fatty acid composition is very important for metabolic plasticity and the growth rate of bacteria. Bacterial and mammalian fatty acid synthases (FAS) have various types. Bacterial and plant fatty acid synthases belong to type II (FASII), and each reaction is catalyzed by distinct single-functional small proteins. Type I fatty acid synthases (FASI), present in mammals and yeast, are composed of one polypeptide chain, and each stage of FAS is accomplished by a various functional domain of this multidomain protein. Figure 1 summarizes the information about the key enzymes of FASII in different bacteria. Due to another principal organization of mammalian FASI, the specific inhibitors of key enzymes of bacterial FASII are expected to be good candidates for the development of low-toxicity antibacterials. Molecules that effectively inhibit fatty acid synthesis in bacteria are discussed extensively in the text below. Table 1 provides information on some specific inhibitors of each known enzyme, as well as their inhibitory concentrations. A demonstration of the ability of the compound to inhibit the activity of the appropriate enzyme in in vitro tests can be considered as direct evidence in favor of a lipid synthesis-related mechanism of action and a specific molecular target. For this reason, the table presents only those inhibitors for which such information can be found in the available literature. The minimum inhibitory concentrations against various bacteria are not presented in Table 1 due to the high variability depending on the bacterial strain. Examples of the most common/known inhibitors are also shown in Figure 1 and marked with a black box. Next, we analyze the possibility of pharmacologically influencing this bacterial pathway and discuss more promising chemical scaffolds for further optimization. It should be taken into account that the therapeutic strategy, in addition to analyzing the inhibitory concentrations, must take into account the risks of developing resistance to the antibiotic and the side effects of its application.
The acetyl-CoA carboxylase is represented by a multiprotein complex, containing biotin carboxylase (AccC), the biotin carboxyl carrier protein (AccB), and biotin carboxyl transferase (AccAD), and performing the carboxylation of acetyl-CoA to generate malonyl-CoA (Figure 1) [36]. Via virtual screening and optimization of the small molecule library, antibacterials aminooxazoles and benzimidazoles were identified as potential AccC inhibitors (Table 1) [3,4]. Comparing the IC50 values, the optimization of the benzimidazole scaffold seems to be a more promising way to find more potent AccC inhibitors, but the appropriate toxicity tests should be carried out to estimate the safe therapeutic window. The biotin carboxylation step is also known to be inhibited by pyrrolocin C and equisetin [37]. The selectivity of pyrrolocin C and equisetin between bacterial and human cells does not exceed 10 [37], and the mechanisms of their toxic action should be elucidated to increase the therapeutic window for more potent AccC inhibitors.
The broad-spectrum antibacterial activity of pyrrolidinedione derivatives, particularly moiramide B and andrimid, is referred to as targeting AccAD (Table 1) [5,38,39,40]. An in silico evaluation of andrimid showed no systemic toxicity [41]. However, bacterial resistance to andrimid arising from a single amino acid mutation in AccAD was found [42]. A promising approach, including the development of dual-use conjugated inhibitors of acetyl-CoA carboxylase composed of covalently linked motifs of aminooxazole (interacting with AccC) and moiramide B (targeting AccAD), was proposed to lower the frequency of strain resistance [43]. Preliminary studies on the antibacterial mechanism of yanglingmycin exhibited the potent inhibition of AccAD, which led to fatty acid and lipid biosynthesis being blocked, and, as a result, cell membrane destruction [44]. Among herbicides, several haloxyfop derivatives were found to demonstrate antimycobacterial activity via the inhibition of AccAD [45].
A malonyl-CoA:acyl carrier protein (ACP) transacylase (FabD) transfers the malonate group from malonyl-CoA to ACP (Figure 1). It is known that FabD is the target for the antibacterial action of aporphine alkaloids [46].
The malonyl-ACP, produced by FabD, is used by several β-ketoacyl-ACP synthases (KASs) of FASII: KAS I (FabB), KAS II (FabF), and KAS III (FabH). FabH initiates the cycle of elongation by condensing malonyl-ACP and acyl-CoA (Figure 1). The origin of the latter strongly depends on the bacteria. The huge variety in the fatty acid profile produced by different bacteria is defined by the substrate specificity of FabH. FabH of Escherichia coli is most specific for acetyl-CoA and propionyl-CoA and incapable of using branched-chain substrates and longer straight-chain acyl-CoA [47,48]. The specificity of Streptococcus pneumoniae FabH is significantly higher towards short (C2–C4) straight-chain than for branched-chain acyl-CoAs [6]. Rather than synthesizing unsaturated fatty acids (UFA) to fluidify the membranes (as Gram-negative bacteria and streptococci do), a number of Gram-positive bacteria (particularly Bacillus subtilis and Staphylococcus aureus) synthesize branched-chain fatty acids (BCFA) [2]. B. subtilis FabH displays higher efficiency with straight-chain and branched-chain acyl-CoA composed of C4–C8 compared to acetyl-CoA [48]. The substrate binding pocket of S. aureus FabH is substantially larger than that of E. coli FabH, and the activity of S. aureus FabH to elongate different acyl-CoAs decreases in the following order: isobutyryl → hexanoyl → butyryl → isovaleryl → acetyl-CoA [49]. The FabH homolog of M. tuberculosis, mtFabH, to synthesize long-chain fatty acids (LCFA) prefers long-chain acyl-CoA substrates composed of C10–C16 rather than acetyl-CoA, short-chain, or branched-chain primers, due to the long internal acyl-binding channel [7,50]. Many FabH inhibitors have been discovered [51], and the dramatic variance in their activity against FabH from various stains clearly indicates the structural differences in the protein active sites. Thiolactomycin is known to inhibit BCFA and straight-chain fatty acid biosynthesis by targeting the FabH of Streptomyces collinus and Streptomyces glaucescens [52,53]. E. coli and S. pneumoniae FabH are weakly inhibited by thiolactomycin, while the indole compound SB418011 significantly inhibits E. coli, S. pneumoniae, and Haemophilus influenza FabH (Table 1) [6]. A number of promising inhibitors of E.coli FabH have been found, including thiazolidine, chrysin, thiazole, deoxybenzoin, salicylaldehyde, pyrazole, cinnamate, carbamate, benzaldehyde, o-benzylhydroxylamine, vanillic acylhydrazone, nitroimidazole, pyrazoline, and piperidine derivatives, furoxan/sulfonylhydrazone hybrids, and others [54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75]. S. aureus FabH is weakly suppressed by thiolactomycin and is efficiently inhibited by different 1,2-dithiole-3-ones, 1,3,5-oxadiazin-2-ones, and amycomicin [76,77,78]. Selected benzoylaminobenzoic acid derivatives are also potent against FabH of Enterococcus faecalis and Streptococcus pyogenes, demonstrate only moderate activity against S. aureus FabH, and are ineffective against H. influenzae FabH [79,80,81]. Alkylsulfonyl compounds and pyrrole-2-carboxylic acid derivatives are specific inhibitors of mtFabH [82,83,84]. Analyzing Table 1, one can conclude that SB418011 has the lowest IC50 against FabH among the presented chemicals. It should be also noted that it did not demonstrate inhibitory activity against human FAS at 200-times higher concentrations [6]. Despite promising differences in selectivity, the in vivo efficacy and toxicity of SB418011 should be evaluated in further experiments.
Two other elongating KASs, FabB and FabF, operating later in the cycle, use acyl-ACP as the substrate for subsequent condensations, instead of the acetyl-CoA used by FabH [85] (Figure 1). The structural similarities between the active sites of FabB, FabF, and FabH [86,87,88,89] presume the development of antimicrobials that are able to hit several KASs at the same time, preventing the de novo synthesis of the fatty acids required for bacterial growth and survival. Two fungal products, thiolactomycin and cerulenin, are non-selective inhibitors of KASs, blocking FabB, FabF, and FabH with varying degrees of success [8,9,90,91]. It is found that cerulenin is characterized by its preferential selectivity towards FabB and FabF and is a very poor inhibitor of FabH (Table 1) [6,7,8,76]. The alteration in the inhibitory activity of cerulenin against various KASs is suggested to be a result of differences in the catalytic triads of the β-ketoacyl-ACP synthases and the size of the acyl-chain binding pockets [90]. Natural bacterial diterpenoid products platensimycin and platencin also target the KASs of FASII [92]. Platensimycin preferentially inhibits the chain elongation enzyme FabF, whereas platencin inhibits both chain initiation and elongation-condensing KASs, FabH and FabF (Table 1) [10,11,12,93,94,95,96]. According to the IC50 information presented in Table 1, platensimycin is the most potent inhibitor of FabF. Moreover, it has great potential to inhibit methicillin-resistant Staphylococcus aureus and vancomycin-resistant Enterococci [97]. The low mammalian cell toxicity and the lack of antifungal activity indicate that platensimycin acts selectively [12,98]. This makes it extremely promising to search for new platensimycin-based antimicrobials in order to improve its pure pharmacokinetic properties. Thus, several semisynthetic analogs of platensimycin with enhanced in vivo efficacy towards MRSA infection in a mouse peritonitis model and improved pharmacokinetic properties have been developed [99,100]. In silico docking studies clearly demonstrated that the specific structural motif presented in fasamycins is predicted to be one more naturally occurring pharmacophore for the specific inhibition of FabF [101]. A series of N-substituted benzoxazolinones were also shown to be active towards elongating KASs, FabB and FabF, by docking into the thiolactomycin binding site [102].
An NADPH-dependent β-ketoacyl-ACP reductase (FabG) produces the reduction of β-ketoacyl-ACP to β-hydroxyacyl-ACP. β-hydroxyacyl-ACP dehydrases (FabZ or FabA) and dehydrates β-hydroxyacyl-ACP to yield trans-2-enoyl-ACP. Another NADPH-dependent reductase of the FASII, trans-2-enoyl-ACP reductase (FabI), reduces trans-2-enoyl-ACP to form acyl-ACP, which can reenter the elongation cycle as a substrate for elongating KASs, FabB or FabF, or can be used for lipid production via phospholipid acyltransferases (PlsB/PlsC or PlsX/PlsY/PlsC system) [36] (Figure 1).
A broad range of plant polyphenols, including epigallocatechin gallate, gallocatechin gallate, epicatechin gallate, catechin gallate, luteolin-7-O-glucoside, luteolin, quercetin, fisetin, morin, and myricetin, are potent inhibitors of all three crucial FASII enzymes, FabG, FabZ, and FabI (Table 1) [13,14,16]. Hexachlorophene and its anthelmintic bis-(2-hydroxyphenyl)methane/sulfide analogs are believed to exhibit antimalarial activity by inhibiting Plasmodium falciparum FabG [103]. Trans-cinnamic acid derivatives and macrolactins showed the inhibition of E. coli and S. aureus FabG (Table 1) [17,104]. Ethyl-6-bromo-2-((dimethylamino)methyl)-5-hydroxy-1-phenyl-1H-indole-3-carboxylate was demonstrated to be potent against Acinetobacter baumannii FabG [105]. A series of small-molecule Pseudomonas aeruginosa FabG inhibitors were identified [106]. According to Table 1, various catechin gallates demonstrate the most impressive activity (expressed as the lowest levels of IC50) against three key FASII enzymes at once, compared to other inhibitors presented. The triple inhibiting action of these polyphenols, as well as their relatively low toxicity, has kept these naturally occurring compounds as the focus of researchers’ attention in terms of finding more safe antibiotics. For example, a safe intake level of green tea polyphenol epigallocatechin gallate, derived from toxicological and human safety data, is about 300 mg per day for adults [107]. We find another approach that involves combining natural polyphenols with other antibiotics that have alternative mechanisms of action on pathogen metabolism to be the most relevant [108,109,110].
FabZ is a crucial enzyme to elongate both saturated fatty acids (SFA) and UFA, and this is why it is an attractive target for the discovery of new antibacterials. The inhibitors of P. falciparum FabZ, NAS-21 (4,4,4-trifluoro-1-(4-nitrophenyl)-butane-1,3-dione), NAS-91 (4-chloro-2-[(5-chloroquinolin-8-5 yl)oxyl]phenol), and their variants, were identified [111]. NAS-21 and NAS-91 analogs also demonstrated activity against mycobacterial FabZ (Table 1) [18]. The natural anthraquinone emodin (3-methyl-1,6,8-trihydroxyanthraquinone) and several synthetic inhibitors of Helicobacter pylori FabZ, based on two promising chemical scaffolds, namely (3,5-dibromo-2,4-dihydroxy-benzylidene)-hydrazide and 2-chloro-5-{5-[3-(2-methoxy-ethyl)-4-oxo-2-phenylimino-thiazolidin-5-ylidenemethyl]-furan-2-yl}-benzoic acid, were discovered (Table 1) [20,112]. Two novel inhibitors of Francisella tularensis and Yersinia pestis FabZ, mangostin and stictic acid, have been found (Table 1) [19]. Moreover, 1,4-naphthoquinone and juglone (5-hydroxyl-1,4-naphthoquinone) is a dual inhibitor of FabZ and FabD that might be potent against M. catarrhalis and H. pylori (Table 1) [21,22]. The therapeutic application of juglone is limited by its possible toxicity [113], but this pharmacophore might be used to find more safe and potent inhibitors of FabZ and FabD.
In Gram-negative bacteria, two enzymes, FabA and FabB, catalyze the production of UFA [114]. FabA has a dual function as a β-hydroxyacyl-ACP dehydratase, catalyzing the dehydration of β-hydroxyacyl-ACP to the trans-2-enoyl-ACP (as FabZ does), and as a trans-2-,cis-3-decenoyl-ACP isomerase, producing the transformation of the trans-2-decenoyl-ACP to a cis-3-decenoyl-ACP (Figure 1). The size of tunnel in the active site of FabA, which perfectly fits trans-2-decenoyl-ACP, determines the specificity of the isomerization reaction at the 10-carbon stage of the unbranched substrate [115]. FabB specifically elongates the cis-UFA produced by FabA. FabA and FabI rival trans-2-decenoyl-ACP, and this balance determines the UFA/SFA ratio and the fluidity of the bacterial membrane [116]. S. pneumoniae and Streptococcus mutans have only one β-hydroxyacyl-ACP dehydrase (FabZ), while another enzyme, FabM, catalyzes the reaction of double-bond isomerization from trans-C2-C3 to cis-C3-C4 [117,118]. In E. faecalis, FabN performs the role of FabA, and FabF elongates the cis-UFA produced by FabN [119]. FabQ from Aerococcus viridans can act as a monofunctional dehydrase like FabZ or as a bifunctional dehydratase/isomerase like FabA (Figure 1) [120]. To produce UFA, some bacteria, particularly B. subtilis and Pseudomonas aeruginosa, contain fatty acid desaturases, introducing a double bond into saturated acyl chains attached to phospholipids or acyl-CoA [114,121]. S. aureus does not contain FabA (or its analogs) or any desaturases, but can utilize the exogeneous UFA via acyl-ACP synthetase [36].
Furthermore, 3-decynoyl-N-acetyl cysteamine is a substrate-mimicking inhibitor of FabA, which covalently bonds to the active site of the enzyme [115]. N42FTA (3-(pyridin-2-yloxy)aniline and N-(4-chlorobenzyl)-3-(2-furyl)-1H-1,2,4-triazol-5-amine) may be a promising scaffold to design more potent inhibitors of P. aeruginosa FabA [122,123].
Many compounds targeting FabI (including those undergoing clinical trials) are known: the front-line antituberculosis drug isoniazid, the common antiseptic triclosan (which demonstrates strong antimalarial activity via P. falciparum FabI inhibition), diazaborines, CG400462, CG400549, MUT056399, AFN-1252, AFN-1720, xanthorrhizol, benzoxaboroles, and their derivatives (Table 1) [15,24,26,27,28,124,125,126,127,128,129,130,131,132,133,134]. FabI inhibition, in the cases of triclosan, CG400462, CG400549, MUT056399, AFN-1252, xanthorrhizol, and benzoxaboroles, was validated by the isolation of resistant clones (Staphylococci and Chlamydia) containing mutations in the FabI gene [28,127,128,131,134,135]. A series of 2,9-disubstituted 1,2,3,4-tetrahydropyrido [3,4-b]indoles, 1,4-disubstituted imidazoles, 1-benzyl-1H-benzimidazoles, and 4-pyridone derivatives, and piperazine and imidazole coumarin derivatives, as well as N-carboxy pyrrolidine analogs inhibiting S. aureus and/or E. coli FabI, were also reported [136,137,138,139,140]. Some natural macrocyclic compounds (complestatin, neuroprotectin, and chloropeptin), methyl-branched fatty acids (14-methyl-9(Z)-pentadecenoic and 15-methyl-9(Z)-hexadecenoic acids), meleagrin, phellinstatin, chalcomoracin, and moracin C demonstrated a promising ability to target S. aureus FabI (Table 1) [23,25,29,30]. In addition to the isoniazid, the trans-2-enoyl-ACP reductase from M. tuberculosis (InhA), which is involved in the biosynthesis of long-chain fatty acids (LCFA) (mycolic acids), was shown to be a target for several triclosan and benzodiazborine derivatives; more specific inhibitors of InhA were also developed [141,142,143,144,145,146,147,148,149,150]. According to Table 1, triclosan- and complestatin-related compounds are characterized by similar low IC50 values against S. aureus FabI. Taking into account that triclosan’s application is limited by the possibility of bacterial resistance development via different mechanisms, including mutations in the genes of FabI or multidrug efflux pump [151,152,153,154] and significant cytotoxic effects [155], the search for natural macrocyclic compounds that are able to inhibit FabI seems to be a more promising approach to identify novel antibiotics.
It should be noted that four enoyl-ACP reductase isozymes have been reported in bacteria (Figure 1). S. pneumoniae, E. faecalis, and Clostridia have FabK instead of FabI, and FabV was discovered in Vibrio cholerae. Moreover, some pathogens have more than one enoyl-ACP reductase; for example, FabI and FabK in pseudomonads and enterococci, FabI and FabL in B. subtilis, and FabI, FabK, and FabV in P. aeruginosa [156,157,158,159]. Triclosan, inhibiting FabI, is a poor inhibitor of FabL and has no activity against FabK and FabV [157]. FabMG, isolated from the soil metagenome, was predicted to be a novel triclosan-resistant enoyl-ACP reductase, revealing that the main mechanism to develop triclosan resistance is a mutation of FabI [160]. Indole naphthyridinones and aquastatin A are inhibitors of both FabI and FabK [31,161], while AG205, atromentin, and leucomelone are thought to be specific to FabK (Table 1) [32,162]. Carfilzomib showed high binding affinity with Klebsiella pneumoniae FabI and FabV [163]. Despite the fact that atromentin has a low IC50 against S. pneumoniae FabK (Table 1), compounds that are able to inhibit various enoyl-ACP reductase isozymes (in particular, complestatin, aquastatin A, and carfilzomib) are more preferable for the development of broad-spectrum antibacterials.
The length of the hydrocarbon chains of the membrane lipids of bacteria is determined by competition between elongating KASs (FabF/FabB) and acyltransferases for acyl-ACP produced by trans-2-enoyl-ACP reductases (FabI/FabK/FabL/FabV). The upper limit is defined by the substrate specificity of the elongating KASs, while the lower limit is a result of the specificity of acyltransferases [164,165,166]. Glycerol-3-phosphate acyltransferases transfer two acyl chains from two acyl-ACPs to the 1 and 2 positions of glycerol-3-phosphate to produce phosphatidic acid (PA), a universal precursor of bacterial phospholipids (Figure 1). Two various acyltransferase systems, PlsX/PlsY/PlsC and PlsB/PlsC, have been discovered [167,168].
The more widespread bacterial pathway for the formation of PA involves the sequential transfer of acyl from acyl-ACP to acyl-phosphate (acyl-PO4) via phosphate acyltransferase (PlsX) and then to the 1 position of glycerol-3-phosphate to produce lysophosphatidic acid (LPA) through acyl-phosphate:glycerol-3-phosphate acyltransferase (PlsY) [164]. Acyl-PO4 is a single substrate for PlsY; it cannot utilize acyl-ACP or acyl-CoA [168]. A series of stabilized acyl-phosphate mimetics, including acyl-phosphonates, reverse amide-phosphonates, and acyl-sulfamates, demonstrated promising activity against S. pneumoniae, Bacillus anthracis, and S. aureus through the inhibition of PlsY (Table 1) [33,34]. The lead compound, having a low IC50 against S. pneumoniae PlsY, (Z)-1-oxooctadec-11-enylphosphoramidic acid, demonstrated potential toxicity [33]. These data necessitate a further search for ways to expand the therapeutic window of acyl-phosphate mimetics.
A glycerol-3-phosphate acyltransferase (PlsB) catalyzes the ligation of the acyl chain to the 1 position of glycerol-3-phosphate to produce LPA. A 1-acyl-sn-glycerol-3-phosphate acyltransferase (PlsC) ligates the second acyl chain (in the 2 position) to the LPA produced by PlsX/PlsY or PlsB to form PA. The advantage of the existence in bacteria of two distinct pathways, the PlsB/PlsC and PlsX/PlsY/PlsC acyltransferase systems, is the possibility of using exogenous fatty acids, because PlsB/PlsC might utilize not only the acyl-ACP produced by trans-2-enoyl-ACP reductases (FabI/FabK/FabL/FabV) but also the acyl-CoA thioesters derived from exogenous fatty acid metabolism—for example, those produced by FabD [168,169]. A fatty acid-rich environment in the host might facilitate the pathogen strain’s resistance to FASII inhibitors by enhancing the assimilation of exogenous fatty acids [170,171]. S. aureus can become insensitive to the FabI inhibitor triclosan via mutations in FabD, lowering FabD activity and inducing the integration of exogenous fatty acids [171,172]. The UFA-to-SFA ratio, membrane fluidity, and cell growth of Rhodobacter sphaeroides were reinstated upon both the inhibition of FabI with diazaborine and the introduction of exogenous UFA [173].
The cyclopropanation of fatty acids is intended to rigidify the bacterial membranes under stress conditions [174,175,176,177,178]. Cyclopropane fatty acids (CFA) are synthesized by cyclopropane fatty acid acyl-phospholipid synthase (CfaS) via the addition of a methylene group to the cis double bonds of the UFA chains of membrane phospholipids (Figure 1) [179]. Dioctylamine inhibits the CfaS of H. pylori (Table 1), preventing bacterial insensitivity to acid stress, antibiotics, and macrophage killing [35]. M. tuberculosis produces a number of cyclopropanated lipids, including mycolic acids, which are essential components to maintain cell wall integrity (Figure 1). This may indicate that agents capable of inhibiting the lipid cyclopropanation enzymes may be an approach to combatting tuberculosis pathogens [180,181].

2.1.2. Biosynthesis of Head Groups of Bacterial Lipids

In addition to the enzymes participating in the synthesis of the fatty acids of bacterial membrane lipids, there are unique enzymes that are involved in the synthesis of the heads of lipid molecules. Figure 2 summarizes the information about the synthesis of the lipid heads.
Cytidine disphosphate-diacylglycerol synthase (CdsA) is the critical enzyme catalyzing the production of the key intermediate in phospholipid diversity, CDP-diacylglycerol (CDP-DG), from cytidine triphosphate (CTP) and PA (Figure 2). CDP-DG is a precursor of the major phospholipids, including phosphatidylglycerol (PG), cardiolipin (CL), phosphatidylethanolamine (PE) (produced through the decarboxylation of phosphatidylserine (PS)), and even phosphatidylinositol (PI) and phosphatidylcholine (PC), which is absent in most prokaryotic cells. The fundamental nature of the CDP-DG-dependent pathway, characteristic of both prokaryotic and eukaryotic phospholipid biosynthesis [182,183], makes the majority of CDP-DG-converting enzymes poor targets for antibiotic therapy.
Phosphatidylglycerol phosphate (PGP) is synthesized by phosphatidylglycerophosphate synthase (PgsA) from CDP-DG via the displacement of cytidine monophosphate (CMP) by glycerol-3-phosphate (glycerol-3-P) (Figure 2). Further, phosphatidylglycerolphosphate phosphatase (PgpA) dephosphorylates PGP to yield PG (Figure 2). Two additional genes of phosphatidylglycerolphosphate phosphatases, PgpB and PgpC, were discovered in E. coli [184]. Subsequently, a cardiolipin synthase (ClsA) utilizes two PG molecules to produce CL and glycerol (Figure 2). Two extra CL synthases have been discovered in E. coli, ClsB and ClsC [185]. The first one can use one molecule of PG and the molecule of another phospholipid as the second substrate [186]. Moreover, ClsB of E. coli can convert PE and glycerol into PG in a PgsA-independent manner [187]. To form CL, ClsC uses PG and PE instead of the two PG molecules [185]. The products of the Cls1 and Cls2 genes of S. aureus are CL synthases with various types of stress-activated production [188]. Three genes of CL synthases have been identified in B. subtilis [189].
The phosphatidylserine synthase (PssA) synthesizes PS from CDP-DG via the displacement of CMP by serine (Figure 2). PS is only a minor biosynthetic intermediate in most bacteria and is decarboxylated by phosphatidylserine decarboxylase (Psd) to produce PE.
PC is also absent in most prokaryotic cells, although some Gram-negative bacteria contain phosphatidylcholine synthases (Pcs) to condense choline into the phosphatidyl moiety of CDP-DG, similar to PssA with serine (Figure 2) [190,191].
One more component that is rarely present in bacterial membranes is PI. For example, Mycobacteria are able to form PI using phosphatidylinositol synthase (PIS) via the exchange of the CMP moiety of CDP-PG for inositol (Figure 2) [192]. Due to the lack of sequence homology between bacterial and mammalian PISs, their different kinetic characteristics, and the essential role of PI in mycobacteria, the PIS of mycobacteria seems to be a good potential drug target for antimycobacterial therapy. Structural analogs of inositol were shown to be more potent inhibitors of mycobacterial PIS compared to mammalian PIS [193]. Alternatively, there is a difference in the bacterial and mammalian biosynthetic pathways used to form PI: in Mycobacteria, PI is produced from CDP-DG and inositol 1-phosphate through an intermediate, phosphatidylinositol phosphate (PIP), which is dephosphorylated subsequently to PI, and inositol 1-phosphate analogs serving as inhibitors of PIP synthase can be used as antimycobacterials [194].
In some Gram-positive bacteria, the anionic glycerophospholipids, particularly PG and CL, can be decorated with aminoacyl residues, most often with lysil, to form cationic PG and CL derivatives by lysil phosphatidylglycerol (LPG) synthase and flippase, multiple peptide resistance factors (MprF) (Figure 2). This pathway is crucial for the adaptation of bacteria to cationic antimicrobial peptides [195,196,197]; for this reason, MprF-targeting antibodies or inhibitors of the factors involved in MprF regulation might sensitize resistant strains to antimicrobial agents [198,199].
In Gram-negative bacteria, a phosphoglycerol transferase (MdoB) transfers sn-1-phosphoglycerol from PG to membrane-derived oligosaccharides (MDO) to obtain diacylglycerol (DG). Further, it is phosphorylated by DG kinases (DgkA and DgkB in Gram-negative and Gram-positive bacteria, respectively) to generate PA, which can be recycled in the phospholipid biosynthetic pathway (Figure 2). DgkA presents a large family of prokaryotic DG kinases that are unrelated to the eukaryotic DG kinases and DgkB [200]. Some products of the dgkA gene are undecaprenol kinases [201].
In Gram-positive bacteria, DG is used to form glycolipids. A diacylglycerol β-glucosyltransferase (YpfP) uses uridine diphosphate-glucose (UDP-Glc) to attach one monosaccharide unit to DG to form monoglycosyl-DG (MGDG) and to add one more Glc residue to MGDG to yield diglycosyl-DG (DGDG) (Figure 2). Although YpfP is a viable target for the development of novel antibacterial drugs [202], there are no approved inhibitors for this enzyme. Anionic glycopolymers, called lipoteichoic acids, composed of 1,3-polyglycerol-phosphate attached to DGDG (anchoring lipoteichoic acids in the membrane), are exposed on the cell walls of Gram-positive bacteria.

2.1.3. Biosynthesis of Lipid A

The outer leaflets of the outer membranes of Gram-negative bacteria are formed by specific lipopolysaccharides (LPS). They consist of O-antigen and core sugar domains and a lipid anchor, known as lipid A. This phosphorylated disaccharide lipid is highly conserved and absolutely required for bacterial growth and survival [203,204]. For this reason, many enzymes involved in lipid A biosynthesis have been identified as targets for antibiotic development [205] (Figure 3, Table 2).
A UDP-N-acetylglucosamine acyltransferase (LpxA) induces the first step of lipid A biosynthesis (Raetz pathway). It transfers a β-hydroxyacyl chain from β-hydroxyacyl-ACP generated by FabG to the 3 position of UDP-N-acetyl-glucosamine (UDP-GlcNAc) (Figure 3). It should be noted that LPS-producing enzymes are highly selective towards ACP thioesters; they cannot be substituted by normal fatty acids [220]. LpxA enzymes are highly specific regarding the acyl chain length. For example, E. coli LpxA transfers only β-hydroxymyristoyl chains [221]. Peptide (Peptide 920, RJPXD33) and small-molecule inhibitors (particularly (R)-(3-(2-chloro-6-methoxybenzyl)morpholino)(3-(4-methylpyridin-2-yl)-1H-pyrazol-5-yl)methanone and erythroskyrin) of LpxA were reported to compete with the substrate or interact with the complex product (Table 2) [206,207,208,222,223,224,225,226,227]. Analyzing Table 2, one can conclude that peptide 920 is of interest due to its relatively low IC50 value, while RJPXD33 demonstrates dual targeting of LpxA and LpxD, offering the possibility to develop novel dual-binding antimicrobials. However, systematic studies of the safety of the peptide’s administration must be performed before it can be determined how promising these methods are.
The acyl transfer reaction by LpxA is thermodynamically reversible and unfavorable, and the subsequent second reaction of the Raetz pathway, catalyzed by UDP-3-O-(R-3-hydroxyacyl)-N-acetylglucosamine deacetylase (LpxC), should occur (Figure 3). LpxC splits the acetyl radical from the UDP-3-(β-hydroxyacyl)-N-acetylglucosamine to produce UDP-3-(β-hydroxyacyl)-D-glucosamine (acyl-UDP-GlcN). Small-molecule inhibitors of the LpxC have been discovered, including hydroxamate-based compounds, exemplified by TU-514, BB-78484, BB-78485, L-159,692, L-161,240, L-573,655, CHIR-090, LPC-009, LPC-011, and LpxC-4 (Table 2) [209,210,211,212,213,214,215,228,229,230,231,232,233,234]. Some of them are highly potent and have proven to be active against various multidrug-resistant Gram-negative bacteria. Analyzing Table 1, it can be assumed that the greatest interest regarding the design of new antibacterials targeting LpxC is in the further optimization of the most effective compounds, L-161,240, CHIR-090, and LpxC-4, in order to avoid emerging resistance [214,234,235,236,237].
A UDP-3-O-(R-3-hydroxyacyl)glucosamine N-acyltransferase (LpxD) performs the third reaction of the lipid A biosynthetic pathway; it transfers a second acyl group from β-hydroxyacyl-ACP to acyl-UDP-GlcN to produce UDP-2,3-bis(β-hydroxyacyl)-D-glucosamine (Figure 3). Some LpxA inhibitors, particularly RJPXD33, also bind to and inhibit LpxD [207,222,224]. It is also believed that LpxD is a drug target of natural compounds like curcumin, gallotannin, isoorientin, neral, isovitexin, vitexin, allicin, ajoene, and cinnamaldehyde [237]. Several synthetic compounds related to hydro-pyrazolo-quinolinones were identified as LpxD inhibitors (Table 2) [216].
A UDP-diacylglucosamine pyrophosphohydrolase (LpxH) hydrolyses UDP-2,3-bis(β-hydroxyacyl)-D-glucosamine to split UMP and to generate 2,3-diacylglucosamine-1-phosphate (lipid X) (Figure 3). LpxI and LpxG are functional orthologs of LpxH in α-proteobacteria and in Chlamydiae, respectively [238,239,240]. LpxH is inhibited by sulfonyl piperazine antibiotics (such as AZ1, JH-LPH-28, JH-LPH-33) (Table 2) [217,218,219,241,242,243]. Bacterial efflux pump functioning was found to be a significant deterrent for LpxH-targeting antimicrobials, highlighting the significance of their combination with antibiotics, permeabilizing the outer membrane to fight multidrug-resistant Gram-negative pathogens [218].
A lipid-A-disaccharide synthase (LpxB) combines the substrate and the product of the LpxH-catalyzed reaction to form the lipid A disaccharide (Figure 3). Compounds that target LpxB have not been discovered to date; only antisense pPNA technology is used to block the lpxB gene [244].
A tetraacyldisaccharide-1-phosphate 4′-kinase (LpxK) translocates the gamma-phosphate of ATP to the 4′ position of the lipid A disaccharide to produce lipid IVA (Figure 3). The 5-(4-carbamoylbenzenesulfonyl)-N-hydroxy-1H-imidazole-2-carboxamide analogs (STOCK6S-33288, 35740, 37164, 39892, and 43621) are believed to be a promising template to develop novel potent LpxK inhibitors [245].
A 3-deoxy-D-manno-oct-2-ulosonic acid (Kdo) transferase (WaaA/KdtA) adds two Kdo residues to lipid IVA to form Kdo2-lipid IVA (Figure 3). Lysophospholipid acyltransferases, LpxL and LpxM, incorporate two additional acyl chains at positions 2′ and 3′ of Kdo2-lipid IVA to yield a hexa-acylated Kdo2-lipid A (Figure 3). At lower temperatures, LpxP might partially perform the function of LpxL. The structure of the active sites of LpxA and LpxD permits the incorporation of myristoyl residues, while acyltransferases LpxL, LpxP, and LpxM transfer lauroyl, palmitoleoyl, and myristoyl chains, respectively [246,247]. LpxN is an ortholog of LpxM in V. cholerae [248]. No information about the specific compounds inhibiting the enzymes in the last steps of LPS biosynthesis are available in the literature.

2.2. Agents with Direct Action on Bacterial Lipid Membranes

Figure 4 summarizes the major mechanisms of the direct action of antibacterial agents on target lipid membranes. The mechanisms include pore formation and a detergent-like manner of action [249]. In the first case, the bacterium dies due to a violation in the water–salt balance via the formation of unauthorized transport pathways for water, ions, and small organic molecules. In the second case, the cause of death is the destruction of the membrane after reaching a critical detergent concentration, and a dramatic enhancement in the membrane fluidity and micellization of membrane lipids.
Antimicrobial peptides are synthetized as components of the immune system in higher eukaryotes to defend them against a wide variety of invasive pathogens [250,251]. Antimicrobial lipopeptides are produced in bacteria or fungi as metabolites and/or to gain a competitive advantage over other species. A number of natural antimicrobial agents exert their defending activities primarily via pathogens’ membrane disruption due to pore formation or the disordering of membrane lipids, and they are characterized by a lower probability of inducing microbial resistance. Thus, owing to the high efficiency of these compounds, their broad-spectrum bactericidal effects, and the low rate of pathogens’ resistance to them, the use of antimicrobial peptides and lipopeptides in clinical practice, as well as in the search for new “natural” antibiotics, seems to be a productive anti-infective therapeutic strategy [252,253]. As a rule, antimicrobial peptides and lipopeptides share common structural features, such as molecular amphiphilicity and a net-positive electrical charge, which govern the binding and permeabilization of the negatively charged bacterial membranes through the mechanisms indicated above.
Table 3 presents examples of natural antimicrobial peptides and lipopeptides, their possible lipid targets, and the threshold concentrations needed to form pores and disintegrate lipid bilayers, mimicking the membranes of sensitive bacteria. Most of the antimicrobial peptides—gramicidin A from Bacillus brevis; alamethicin produced by the fungus Trichoderma viride; pardaxin isolated from secretions of the Red Sea Moses sole; melittin and mastoparan isolated from bee and wasp venom, respectively; protegrin-1 found in porcine leukocytes; magainin found in frog skin; ceratotoxins and cecropins discovered in the accessory gland secretion fluid of the insect Ceratitis capitata and the hemolymph of Hyalophora cecropia, respectively; nisin from Streptococcus lactis; cinnamycin and its close analog duramycin from Streptomyces sp.; mammalian defensins; human cathelicidin LL-37 [254,255,256,257,258,259,260,261,262,263,264,265,266,267,268,269,270,271,272,273,274]; lipopeptides; colistins (polymyxins) from Bacillus polymyxa; daptomycin from Streptomyces roseosporus; and gausemycin from Streptomyces sp. [108,275,276,277,278,279]—manifest their action via the pore formation mechanism (Table 3). The pores formed by antimicrobial agents are characterized by their different architectures [280]. For example, alamethicin, pardaxin, and seratotoxin A pores are believed to be “barrels” composed of peptide aggregates [271,272,281,282], while mellitin, magainin, and polymyxin B form (lipo)peptide–lipid toroidal pores [108,270,283,284,285] (Figure 4). Some antimicrobial agents are not shown to form transmembrane pores; they act as detergents by forming a peptide “carpet” on the membrane surface (Figure 4). Such properties are exhibited by cecropin P1, lasioglossin III, and aurein 1.2 (Table 3) [264,286,287,288]. The peptide thanatin disrupts the bacterial outer membrane [289]. In the case of cecropins and protegrins, dual activity was found, including pore formation and detergent-like model action [259,264,290]. In any case, it is likely that when a critical antibiotic concentration is reached, regardless of whether the agent can form pores and in what way, an irreversible change in the rheological properties of the lipid bilayer occurs and it will be destroyed [291,292] (Figure 4).
Despite the lower bacterial resistance to the naturally occurring antibiotics acting on the microbial membranes compared to classical antibiotics, including those inhibiting lipid biosynthesis, antimicrobial peptides and lipopeptides are not a panacea for the emergence of resistance in pathogenic bacteria. One of the evolutionary mechanisms by which to develop pathogenic resistance to cationic antibacterial agents is a reduction in the total negative charge of the cell surface of a microorganism to reduce the initial electrostatic binding. Thus, the resistance of S. aureus to defensins and protegrins is determined by the activity of MprF, an enzyme that modifies phosphatidylglycerol with L-lysine (1.1.2), which, in turn, leads to a decrease in the surface membrane charge and the repulsion of cationic peptides [299]. Pseudomonas fluorescens was proposed to diminish the net anionic charge of the cytoplasmic membrane by reducing the content of anionic phospholipids and increasing the concentration of positively charged ornithine–amide lipids that lead to the resistance to the cationic polymyxin B [300]. According to the literature data, a change in the structure of the LPS of Gram-negative bacteria E. coli, Salmonella enterica, Salmonella typhimurium, K. pneumoniae, and P. aeruginosa induced by the attachment of L-arabinose or phosphatidylethanolamine to the phosphate residues of lipid A leads to the emergence of resistance among these microorganisms to polymyxins due to changes in the membrane surface charge [301,302,303,304,305]. In turn, daptomycin is recommended for application as a therapy against β-lactam-resistant Streptococcus mitis. The target of daptomycin is thought to be phosphoglycerol [306]. However, S. mitis can rapidly develop resistance to daptomycin via loss-of-function mutations in the gene of CdsA, which catalyzes the formation of a common phospholipid precursor, CDP-DG (1.1.2); moreover, daptomicin-resistant strains exhibit the absence of anionic phospholipid membrane microdomains composed of CL and PG [307]. Daptomicin resistance in E. faecalis was found to be associated with changes in the genes of cardiolipin synthase, Cls, and cyclopropane fatty acid synthase, CfaS (1.1.2) [308]. The latter indicates that reducing the level of negatively charged lipids is not the only strategy for resistance development by changing the membrane properties; the fatty acid profile is also of fundamental importance. For example, the development of resistance of S. aureus to gausemycin A is accompanied by growth in the ratio between the levels of anteiso- and iso-BCFA [309]. The membrane fluidity is significantly enhanced when anteiso acyl chains replace iso acyl chains. In contrast, the resistance of S. aureus to daptomycin and Listeria monocytogenes to nisin develops with an increase in the percentage of SFA compared to BCFA, which should lead to a decrease in membrane fluidity [310,311]. Thus, alterations in the fatty acid profile and rheological properties of the membrane may be another important factor determining the sensitivity of pathogens to antibiotics. Moreover, whether the fluidity of the membrane should be increased or decreased depends on the architecture of the pores formed by a specific antimicrobial agent.

3. Antifungal Agents with Lipid-Related Mechanisms of Action

3.1. Inhibition of Biosynthesis of Fungal Cell Membrane Components

Fundamentally, fungal walls are all engineered in a similar way and contain the cell membrane and cell wall [312]. The absence of a cell wall in mammalian cells provides an opportunity for the development of antifungal agents that target the enzymes involved in the biosynthesis of cell wall components in fungi, chitin synthase (Chs) and β-1,3-glucan synthase (Fks) [313,314,315]. The resistance of Aspergillus fumigatus and Candida glabrata to semisynthetic ehinocandin and caspofungin might arise from not only mutations in the Fks gene but also from alterations in the lipid microenvironment of the enzyme due to an increase in dihydrosphingosine and phytosphingosine content [316,317]. Thus, it should be taken into account that although the cell wall is an essential structure, maintaining the integrity and viability of fungal cells, the fungal lipid membrane serves as both a second barrier and a platform for the functioning of the enzymes that are responsible for the cell wall’s biosynthesis (Chs and Fks) [318,319]. The fungal cell membrane is composed of various glycerophospholipids, sphingolipids, and ergosterol. The latter component is particularly interesting in terms of antifungal targeting, since mammalian cell membranes include another sterol, cholesterol.

3.1.1. Biosynthesis of Fatty Acids of Fungal Membrane Lipids

Recently, a discrepancy between the human and fungal FASI has been discovered [320]. The human FAS encoded by the FASN gene is a type Ib FAS. It consists of one polypeptide chain, including seven domains that assemble into homodimers [321]. Yeast FAS belongs to type Ia FAS and includes a heterododecameric complex composed of six subunits α and six subunits β, which are encoded by the genes Fas1 and Fas2 [322]. It was shown that the deletion of the FAS genes in Cryptococcus neoformans significantly reduced the growth and virulence of the fungi [323,324,325]. Thus, the differences in fungal and human FAS [320] can, in Candida albicans, potentially be used to target broad-spectrum antifungals towards the products of the Fas1 and Fas2 genes. However, the fungal mutants for the corresponding FAS genes could survive due to the utilization of exogenous fatty acids [326], which might significantly reduce the possibilities of anti-FAS therapy. There have been few fruitful efforts to repurpose antibacterial FAS inhibitors. FAS inhibition in C. neoformans with the FASII inhibitor cerulenin (1.1) drastically reduced the inhibitory concentration of the inhibitor of ergosterol synthesis, fluconazole (2.4) [323]. Cerulenin (but not platensimycin and thiolactomycin) was shown to inhibit Saccharomyces cerevisiae FAS [327]. The attempts to inhibit a product of the OLE1 gene, fatty acid Δ9 desaturase, were more successful in terms of targeting fatty acid biosynthesis in C. albicans [328,329].

3.1.2. Biosynthesis of Phospholipid Head Groups

Similar to bacteria (Figure 2), the biosynthesis of the fungal phospholipids begins with the common precursor CDP-DG, which is produced from PA by CdsA [330,331]. Further, PI, PGP, and PS are generated from CDP-DG by PIS, PgsA, and PssA, respectively [330,332]. PGP is dephosphorylated by PgpA to form PG, and PG is condensed to CL by ClsA [330,332]. It is important that the major phospholipid present in most eukaryotic membranes is PC, and PS is a key substrate for PC synthesis in yeast and fungi [333]. The Psd enzyme converts PS to PE. The main pathway for PC synthesis in yeast involves the three-step methylation of PE (Figure 5). The first stage includes the methylation of PE by phosphatidylethanolamine N-methyltransferase (Pems) to form the phosphatidyl-N-monomethylethanolamine (PMME) and the methylation of PMME to form phosphatidyl-N,N-dimethylethanolamine (PDME) [330]. PDME is converted to PC by PgpA. It was shown that the disruption of PS and PE biosynthesis within the CDP-DG pathway causes the avirulence of C. albicans [334]. Moreover, the action of some fungicides is associated with Pems inhibition [335]. As, in the cells of higher eukaryotes, PC is mainly synthesized from exogenous ethanolamine and choline via the Kennedy pathway [336,337] (Figure 5), one might suggest that the enzymes that perform the methylation reactions in PC biosynthesis by the CDP-DG pathway can be potential targets for antifungals. However, the alternative Kennedy pathway can be used by lower eukaryotes to produce PC and determines the possibility of developing resistance to the action of such antibiotics.

3.1.3. Biosynthesis of Sphingolipids

Figure 6 demonstrates the pathway for the synthesis of sphingolipids in S. cerevisiae [338,339,340]. Serine palmitoyltransferase (SPT) performs the condensation of L-serine and palmitoyl-CoA to lead to 3-ketodihydrosphingosine [341]. Meanwhile, 3-ketodihydrosphingosine reductase (KDSR) converts 3-ketodihydrosphingosine to dihydrosphingosine. Phytoceramide can be synthesized from dihydrosphingosine by ceramide synthase (CerS) and sphingosine C4-hydroxylase (SCH) through two alternative intermediates, ceramide and phytosphingosine (Figure 6). The next reaction involves inositol-phosphoceramide synthase (IPCS), which converts phytoceramide to inositolphosphatyl-ceramide (IPC). IPC can be further mannosylated by mannosylinositol phosphorylceramide synthase (MIPCS) and condensed with additional inositolphosphate using inositolphosphotransferase (IPS) to yield the more complex sphingolipids MIPC and M(IP)2C, respectively [341,342,343].
Comparing the enzymes in the fungal and mammalian cells required for sphingolipid biosynthesis, it can be concluded that they are homologous and, consequently, are not suitable as targets for the development of low-toxicity antifungal drugs. Nevertheless, potent inhibitors of fungal sphingolipid biosynthesis are described in the literature, although a proper assessment of their possible toxicity has not been completed yet. Mainly, the compounds target SPT and IPCS. Table 4 summarizes the available information about the agents targeting the synthesis of sphingolipids in different fungi, their chemical structures, and the IC50 values against appropriate fungal enzymes. In particular, a variety of different SPT inhibitors have been isolated, including lipoxamycin [344,345], myriocin [346], sphingofungins [347,348,349], and viridiofungins [350]. It is interesting that sphingofungins are characterized by similar activity against C. albicans and S. cerevisiae SPT, while viridiofungins show 70-200-fold higher selectivity towards C. albicans SPT (Table 4). Unfortunately, SPT inhibitors demonstrated high toxicity towards mammalian cells due to the inhibition of human SPT1 [351]. Interestingly, the S. cerevisiae SPT is composed of three different subunits, known as Lcb1, Lcb2, and Tsc3, and a homologue of Tsc3 has not been found in mammals [352]. Thus, the question of whether Tsc3 inhibition would be sufficient to effectively suppress the fungal SPT activity in pathogenic fungi awaits elucidation. Fumonisins are effective inhibitors of CerS, but also demonstrate toxicity to mammalian cells [353]. Australifungin is a very potent inhibitor of fungal CerS from several species; however, the α-diketone and β-ketoaldehyde functional groups present in this compound have high chemical reactivity, which seriously limits australifungin’s use [348]. The inositol-phosphoceramide synthase (IPCS) identified in S. cerevisiae and other different fungi [354] is a potential target for antifungals. IPCS inhibitors include aureobasidin A [355], khafrefungin [356], haplofungins [357], and pleofungin [358]. Galbonolide A and B inhibit the IPCS of B. cinerea and C. neoformans [359,360,361]. An analysis of Table 4 shows that khafrefungin, pleofungin A, and galbonolide A, also known as rustmicin, are the most potent inhibitors of C. albicans, A. fumigatus, and C. neoformans IPCS, respectively. Khafrefungin seems to be a more promising candidate due to its relevant selectivity between fungal and mammalian IPCS [356,362], while rustmicin’s application is limited by its low metabolic stability and drug efflux in fungi [359]. Thus, a further search may accelerate the discovery of selective low-toxicity natural inhibitors for fungi.

3.1.4. Ergosterol Synthesis

Contrary to the cholesterol-containing membranes of mammalian cells, fungal cell membranes are enriched with ergosterol [364]. Figure 7 demonstrates the ergosterol biosynthesis pathway in S. cerevisiae [365].
The ergosterol biosynthesis in S. cerevisiae includes three different modules, mevalonate, farnesyl-PP, and ergosterol biosynthesis [366]. Table 5 provides the available information about the potential inhibitors of enzymes participating in the ergosterol biosynthetic pathway.
Acetyl-CoA C-acetyltransferase (ERG10) catalyzes the additional acetylation of acetyl-CoA molecules to produce acetoacetyl-CoA, which is further transformed by hydroxymethylglutaryl-CoA synthase (ERG13) to 3-hydroxy-3-methylglutaryl-CoA. Mevalonate is synthesized by NADPH-dependent hydroxymethylglutaryl-CoA reductase (HMG1/2) [366,367]. As the synthesis of mevalonate is the critical step in the ergosterol biosynthetic pathway, it is believed that HMG1/2 might be a good target for antifungals. It is well known that statins competitively bind to human 3-hydroxy-3-methylglutaryl coenzyme-A reductase, preventing the conversion of 3-hydroxy-3-methylglutaryl-CoA into mevalonate [368,369], and the prospects for the repurposing of statins to treat fungal infections should be estimated. Supporting this theory, simvastatin, lovastatin, atorvastatin, pravastatin, fluvastatin, and related compounds were reported to decrease the intracellular ergosterol level via the inhibition of HMG1/2 in C. glabrata, C. albicans, Ustilago maydis, Trichothecium roseum, S. cerevisiae, C. neoformans, Zygomycetes, and Aspergillus spp. [370,371,372,373,374,375,376,377,378,379,380,381]. Moreover, statin therapy is associated with a reduction in oral Candida carriage in hyperlipidemic patients [382]. Statins also reduced mortality due to the diminishing risk of fungal-related complications in patients with diabetes, hematologic malignancies, and COVID-19 [368,383,384,385].
Further, mevalonate is successively phosphorylated to mevalonate-PP by two different kinases, mevalonate kinase (ERG12) and phosphomevalonate kinase (ERG8). Diphosphomevalonate decarboxylase (ERG19) performs the transformation of mevalonate-PP to isopentenyl-PP [386]. Farnesyl-PP is a product of two successive reactions catalyzed by farnesyl diphosphate synthase (ERG20) [387].
Using NADPH, squalene synthase (ERG9) produces squalene from farnesyl-PP. Natural fungal metabolites, such as zaragozic acids, are potent inhibitors of ERG9 [388]; however, due to their high toxicity, the compounds failed to reach the clinical trial phase [389].
Squalene epoxidase (ERG1) and 2,3-oxidosqualene cyclase (ERG7) catalyze the synthesis of squalene epoxide and lanosterol, respectively. It was found that allylamines, naftifine, and terbinafine are reversible inhibitors of the Candida ERG1 (Table 5) [390,391]. In addition to allylamines, the thiocarbamates tolciclate and tolnaftate were also shown to be potent inhibitors of ERG1 (Table 5) [392]. Table 5 clearly demonstrates that all presented allylamines and thiocarbamates are more effective against T. rubrum ERG1 than against C. albicans squalene epoxidase. However, the resistance of Trichophyton spp. to terbinafine, licensed for the treatment of dermatophytic infections, increases dramatically [393,394,395], creating a serious limitation to its further clinical application. Moreover, there is evidence in favor of terbinafine-induced hepatotoxicity [396]. The emerging resistance of dermatophytes to terbinafine and the moderate activity of both allylamines and thiocarbamates against C. albicans show the need for a further search for highly effective ERG1 inhibitors.
Lanosterol is converted to zymosterol via two intermediates, 4,4-dimethyl-zymosterol-8,14,24-trienol and 4,4-dimethyl-zymosterol, by lanosterol 14α-demethylase (ERG11) and sterol C14-reductase (ERG24). Azoles were identified as effective inhibitors of ERG11 (Table 5) via selective coordination with heme iron [397,398] and demonstrated striking antifungal activity against a variety of human fungal pathogens [399,400]. Importantly, azoles inhibit human lanosterol 14α-demethylase at substantially higher concentrations than the fungal enzyme [397]. A series of steroidal 1,4-dihydropyridines also showed promising activity against various Candida spp. via the inhibition of ERG11 [401]. Despite the pronounced activity of azoles against Candida spp. ERG11 (Table 5), the enlarged resistance to azoles by Candida species is a serious threat in their clinical use [402,403]. Therefore, the novel azole-based derivatives could attract attention as ERG11 inhibitors.
In the fungal cell, sterol C24-methyltransferase (ERG6) transforms zymosterol to fecosterol. It is converted to episterol by sterol C8,7-isomerase (ERG2). Sterol C5(6)-desaturase (ERG3) converts episterol to ergosta-5,7,24(28)-trienol [404,405,406]. NADPH-dependent sterol C22-desaturase (ERG5) catalyzes the formation of the next intermediate, ergosta-5,7,22,24(28)-tetraenol. At the final step, NADPH-dependent sterol C24-reductase (ERG4) converts ergosta-5,7,22,24(28)-tetraenol to ergosterol molecules [365]. It was demonstrated that amorolfine, fenpropidin, fenpropimorph, and the related morpholines and piperidines act as dual inhibitors of ERG24 and ERG2 [407,408]. Thus, these compounds seem to be ideal antifungals, as acquiring resistance against them will be difficult for pathogens because of the requirement to mutate two enzyme genes at once. Among the aminopiperidine derivatives, as presented in Table 5, compound 1b has a lower IC50 value, and the substantial prolongation of survival in infected mice with its oral administration [409] indicates the potential clinical benefits. However, presently, only amorolfine is used clinically to treat nail infections.
It should be noted that zymosterol is a precursor in cholesterol biosynthesis in mammalian cells [410] (this branch of the biosynthetic pathway is marked with an orange color in Figure 7). Ergosta-5,7,24(28)-trienol is a precursor of phytosterols, campesterol, β-sitosterol, and stigmasterol [411] (this branch of the biosynthetic pathway is indicated with a green color in Figure 7).
Table 5. Major inhibitors of fungal sterol biosynthesis.
Table 5. Major inhibitors of fungal sterol biosynthesis.
InhibitorStructureEnzyme IC50, μMReferences
terbinafineAntibiotics 12 01716 i098ERG1C. albicans0.03[390,392]
C. parapsilosis0.02–0.04[390]
C. glabrata0.137[390]
Trichophyton rubrum0.002–0.016[390,392]
A. fumigatus0.24[390]
naftifineAntibiotics 12 01716 i099ERG1C. albicans1.1[390]
C. parapsilosis0.34[390]
T. rubrum0.115 ± 0.030[392]
SDZ 87-469Antibiotics 12 01716 i100ERG1T. rubrum0.020 ± 0.005[392]
C. albicans0.011[392]
tolciclateAntibiotics 12 01716 i101ERG1T. rubrum0.028 ± 0.003[392]
C. albicans0.12[392]
tolnaftateAntibiotics 12 01716 i102ERG1T. rubrum0.052± 0.009[392]
C. albicans1.04[392]
bifonazoleAntibiotics 12 01716 i103ERG11C. albicans0.3[397]
clotrimazoleAntibiotics 12 01716 i104ERG11C. albicans0.091[397]
miconazoleAntibiotics 12 01716 i105ERG11C. albicans0.072[397]
fluconazoleAntibiotics 12 01716 i106ERG11C. albicans0.051–0.6[397,412]
C. neoformans0.17[413]
Malassezia globosa0.206  ±  0.008[414]
itraconazoleAntibiotics 12 01716 i107ERG11C. albicans0.039–0.4[397,412]
C. neoformans0.17[413]
M. globosa0.188  ±  0.008[414]
voriconazoleAntibiotics 12 01716 i108ERG11C. neoformans0.17[413]
VT-1129Antibiotics 12 01716 i109ERG11C. neoformans0.16[413]
ketoconazoleAntibiotics 12 01716 i110ERG11C. albicans0.064–0.5[397,412]
M. globosa0.176  ±  0.016[414]
ketaminazoleAntibiotics 12 01716 i111ERG11M. globosa0.321  ±  0.042[414]
compound 1aAntibiotics 12 01716 i112ERG24C. albicans0.063[415]
compound 1bAntibiotics 12 01716 i113ERG24C. albicans0.016[415]

3.2. Agents with Direct Action on Fungal Lipid Membrane

The principles of action of naturally occurring antibiotics on fungal membranes are similar to those of antibacterial peptides and lipopeptides (Figure 4).
Table 6 summarizes the data concerning the effect of antifungal lipopeptides and polyene macrolides on the permeability of lipid bilayers that mimic the cell membranes of target fungi. One of the most attractive groups is the cyclic lipopeptides, which are secondary metabolites of certain bacteria and are used to combat plant fungal pathogens. It is well known that the syringomycins and syringopeptines from Pseudomonas syringae, and surfactins, fengycins, iturins, bacillomycins, and mycosubtilin from B. subtilis, form the transmembrane pores in model lipid membranes [416,417,418,419,420,421,422,423,424,425,426,427].
Another clinically important group is antifungal macrolides and polyene antibiotics. Amphotericin B, nystatin, and fillipin demonstrate antimicrobial activity via the formation of transmembrane pores in the target-sterol-containing membranes [432,433,434,438,439,440,441,442,443]. The three-dimensional structure of the amphotericin B channel was proposed as an asymmetric heptameric complex of polyene and sterol molecules penetrating the membrane [444]. In addition to natamycin’s inhibitory effect on transport proteins, it was suggested to specifically interact with the sterol- and sphingomyelin-enriched ordered phase and disrupt lipid packing [437,445]. Amphidinol 3, an antifungal polyene isolated from a marine dinoflagellate, is also able to induce pore-like defects in model membranes [446].
Antimicrobial peptides also demonstrate antifungal efficiency. Piscidins identified in the mast cells of fish exert their fungicidal effects on C. albicans by disrupting the fungal membranes through pore formation [436,447]. An antimicrobial peptide from the tree frog Hyla punctata, hylaseptin P1-NH2, demonstrates strong antifungal potential by promoting membrane disruption [448].
Saccharomyces cerevisiae strains, resistant to syringomycin E, are characterized by a decrease in the length of fatty acid chains and sphingolipid content. Mutants were defective in two key enzymes of the terminal sphingolipid biosynthetic pathway, IPCS and IPS (2.1.3) [449]. The sensitivity of S. cerevisiae towards syringomycin E was also shown to depend on the C4-hydroxylation of sphingoid bases to form phytoceramide, catalyzed by SCH [450]. The relevance of M(IP)2C [451] and sphingolipid C4-hydroxylation [452] for the lateral segregation of lipids in S. cerevisiae membranes might suggest that syringomycin E may interact with sphingolipid-enriched microdomains, and its pore-forming ability is sensitive to their composition [453,454].
As expected for the sterol-dependent mechanism of pore formation by polyene antibiotics, the decline in the ergosterol content in the plasma membranes of target fungi results in the development of resistance [455]. In fact, it was found that the minimal inhibitory concentration of amphotericin B against C. albicans was increased by the deletion of ERG2, ERG6, ERG3, and ERG11, the enzymes participating in the ergosterol biosynthetic pathway (2.1.4). It should be noted that the emergence of resistance to amphotericin B through a decrease in ergosterol content makes resistant strains extremely sensitive to osmotic and other types of stress. The reduced level of ergosterol in clinical strains of Candida lusitaniae, which is resistant to amphotericin B, might arise from mutations in the ERG3 gene [456].

4. Antivirals Targeting Lipid Envelope

Since we have narrowed our focus to reviewing only compounds that directly target pathogen membranes, antivirals that have been shown to affect the membranes of virions, which lead to the destruction of the lipid envelope or suppression of virus fusion with the host cell, are discussed below.
Many socially significant viruses are enveloped, i.e., the virions are surrounded by a supercapsid composed of a lipid bilayer. Despite the fact that the origin of the lipid envelope is the host cell membrane, in some cases, a quantitative difference has been found in the content of various lipids in the viral envelope and the host cell membrane from which the virions have been budded [457,458,459,460,461]. This may also be due to virus-induced changes in the host cell’s lipid metabolism [462,463]. Thus, the lipid membranes of enveloped viruses might be considered a target for innovative antiviral drugs. The compounds are thought to break the lipid envelopes of virions or dramatically change the properties of the viral membrane in order to prevent fusion with the cell membrane. A significant advantage of using such an approach is the broadening of the spectrum of antiviral activity and a decrease in the resistance to viral pathogens.

4.1. Disrupting Agents

4.1.1. Photosensitizing Antivirals

Photosensitizers are compounds that can absorb light and generate reactive oxygen species, which, in turn, leads to the peroxidation of membrane lipids and damage to the lipid bilayers of both viral and cellular membranes [464] (Figure 8A). In the absence of virus systems for reparation, the photodamage of the lipid envelope causes a dramatic reduction in the infectivity of virions due to the inactivation of viral fusion. Among the photosensitizers, compounds with absorption in the infrared region are of particular interest in the search for new broad-spectrum antivirals due to their substantially higher tissue transparency for the radiation of this spectrum [465].
Hypericin, a plant-occurring polycyclic quinone, demonstrated broad-spectrum activity against enveloped viruses such as human immunodeficiency virus type 1 (HIV-1), Moloney murine leukemia virus, equine infectious anemia virus, vesicular stomatitis virus (VSV), herpes simplex virus types 1 (HSV-1) and 2 (HSV-2), parainfluenza virus (PIV), vaccinia virus, murine cytomegalovirus (mCMV), and Sindbis virus (SINV) [466,467,468,469,470,471] (Table 7). Hypericin did not alter non-enveloped viruses [467]. Halogen derivatives of hypericin were shown to be effective against HSV-1 [472,473]. Gymnochromes isolated from Gymnocrinus richer were shown to be highly potent antiviral agents against dengue virus, HSV-1, and influenza virus type A (IVA) [474,475] (Table 7). Hypocrellins from Hypocrella bambuase also demonstrated light-dependent anti-HIV, anti-mCMV, anti-HSV-1, anti-VSV, and anti-IVA efficacy [469,476,477,478] (Table 7), while the non-enveloped virus was not inactivated [477].
Initially, rigid amphipathic perylene-containing nucleoside derivatives, particularly 5- (perylen-3-yl)ethynyl-2′-deoxy-uridine (dUY11) and 5-(perylen-3-yl)ethynyl-arabino-uridine (aUY11), having considerable activity against IVA, hepatitis C (HCV), HSV-1, HSV-2, mCMV, VSV, SINV, tick-borne encephalitis virus (TBEV), yellow fever virus (YFV), Chikungunya virus (CHIKV), African swine fever virus, PIV, and human respiratory syncytial virus (RSV) (Table 7), were believed to inhibit viral fusion by affecting membrane curvature stress (Section 3.2) [479,480,481]; however, later, their action through the photosensitization of viruses was postulated [482,483,484,485,486,487]. No activity or the non-specific activity of perylene derivatives against non-enveloped viruses was found [480,488]. Non-nucleoside perylene derivatives also showed promising antiviral activity [489].
BODIPY-based photosensitizer 2,6-diiodo-1,3,5,7-tetramethyl-8-(N-methyl-4-pyridyl)-4,4′-difluoroboradiazaindacene (DIMPy-BODIPY) exhibited the photodynamic inactivation of dengue virus and VSV at nanomolar concentrations [490].
A class of thiazolidine-based lipophilic inhibitors of LJ and JL series demonstrating high activity against a variety of enveloped viruses (Table 7), with no effect on the infection of non-enveloped viruses, was also originally described as curvature-induced antivirals (Section 3.2), but it was later shown that the compounds act as membrane-targeted photosensitizers [491,492,493].
Cationic imidazolyl and pyridyl porphyrins were characterized as promising photosensitizing antivirals against SARS-CoV-2 and HSV-1 [494,495]. A natural chlorine photosensitizer, pheophorbide a, inactivates HSV-1, HSV-2, MERS-CoV, SARS-CoV-2, YFV, HCV, and SINV, by targeting the lipid envelope [496,497] (Table 7).
Table 7. Photosensitizers and their antiviral activity.
Table 7. Photosensitizers and their antiviral activity.
PhotosensitizerStructureVirusIC50, µMReference
hypericinAntibiotics 12 01716 i127HIV-10.44[466]
HSV-10.006[469]
gymnochrome BAntibiotics 12 01716 i128dengue0.029[475]
hypocrellin AAntibiotics 12 01716 i129HSV-10.015[469]
hypocrellin BAntibiotics 12 01716 i130HSV-10.025[469]
5-(perylen-3-yl)ethynyl-2′-deoxy-uridine (dUY11)Antibiotics 12 01716 i131IVA0.097–2.7[480,498]
HSV-10.048–0.131[479]
HSV-20.031–0.055[479,480]
HCV0.183–0.187[479,480]
mCMV0.037 ± 0.016[480]
SINV0.006 ± 0.001[480]
TBEV0.024 ± 0.013[483,499]
PIV2.2 ± 0.5[498]
RSV1.8 ± 0.2[498]
SARS-CoV-20.2564[487]
5-(perylen-3-yl)ethynyl-arabino-uridine (aUY11)Antibiotics 12 01716 i132IVA0.078–5.2[480,498]
HSV-10.048 ± 0.012[479]
HSV-20.052 ± 0.003[480]
HCV0.107 ± 0.041[480]
mCMV0.013 ± 0.004[480]
SINV0.011 ± 0.005[479]
TBEV0.018 ± 0.010[483]
YFV0.0086 ± 0.0007[484]
CHIKV<0.78[484]
PIV1.3 ± 0.3[498]
RSV2.3 ± 0.1[498]
SARS-CoV-20.4058[487]
(5Z)-5-[(5-phenylfuran-2-yl)methylidene]-3-prop-2-enyl-2-sulfanylidene-1,3-thiazolidin-4-one (LJ-001)Antibiotics 12 01716 i133HIV0.133[492]
Newcastle disease virus0.095[492]
Ebola virus0.9[492]
IVA0.026[492]
Nipah virus0.048[492]
Hendra virus0.018[492]
Rift valley fever virus0.02[492]
Semliki forest virus0.537[492]
HSV-10.02[492]
hCMV0.13[492]
VSV0.298[492]
(Z) 3-ethyl-5-[5-(2-methoxyphenyl)-furan-2-ylmethylene]oxazolid-ine-2,4-dithione (JL-103)Antibiotics 12 01716 i134HIV0.013[492]
Newcastle disease virus0.004[492]
Ebola virus0.185[492]
IVA0.002[492]
Nipah virus0.004[492]
Hendra viru0.0005[492]
Rift valley fever virus0.003[492]
Semliki forest virus0.044[492]
HSV-10.002[492]
hCMV0.004[492]
VSV0.011[492]
5,15-bis(1,3-dimethylimidazol-2-yl)chlorin (ICH-Me2+)Antibiotics 12 01716 i135SARS-CoV-20.12[494]
pheophorbide aAntibiotics 12 01716 i136SARS-CoV-20.18[497]
MERS-CoV0.18[497]
IC50 is determined at photoactivation.

4.1.2. Tweezers

Molecular tweezers are membrane-destabilizing agents that can disrupt the virus lipid envelope and can be used as broad-spectrum antivirals against influenza A virus, respiratory syncytial virus, human immunodeficiency virus, herpes simplex viruses, human cytomegalovirus, Ebola and Marburg viruses, SARS-CoV, SARS-CoV-2, MERS-CoV, and other enveloped viruses. These small molecules act as pincers that bind lipid head groups and disrupt lipid ordering and packing in the virus lipid envelope, which results in the virions being unable to infect the cells [500] (Figure 8B).
A basing compound, CLR01, was shown to inhibit HIV-1, Ebola, Zika, herpes simplex (HSV-1, HSV-2), measles, influenza virus, and SARS-CoV-2 infection by directly targeting the viral membrane [501,502,503] (Table 8). Its close analog, CLR05, also possessed broad-spectrum antiviral activity [500]. CLR01 and CLR05 did not reduce infection by the non-enveloped adenovirus and encephalomyocarditis virus [500]. The membrane-disrupting and, consequently, the antiviral activity of CLR01 might be substantially potentiated by the introduction of C4, C7, or aromatic radicals to each phosphate group [500,503] (Table 8).

4.1.3. Antimicrobial Peptides

Although many antimicrobial peptides, especially defensins and cathelicidins, have been shown to possess antiviral effects [273,504,505], the mechanisms of antiviral action are highly pleiotropic and involve more than simply a direct effect on the viral membrane. Regarding the focus of this section, bomidin, a naturally occurring antimicrobial peptide that is active against a variety of enveloped viruses, including SARS-CoV-2, HSV, dengue virus, and CHIKV, was supposed to disrupt the viral membrane [506]. Plantaricin NC8 αβ, a two-peptide bacteriocin produced by Lactobacillus plantarum strains, was shown to inhibit SARS-CoV-2, IVA, flaviviruses Langat and Kunjin, and HIV-1, via permeabilizing and destroying their envelopes [507] (Table 9). It was demonstrated that the anti-HIV activity of a cyclic peptide from plants, kalata B1, resulted from the disruption of the membranes of HIV particles due to their raft-like lipid density and enrichment with PE [508].

4.2. Fusion Inhibitors Affecting Membrane Fluidity and/or Curvature Stress

An essential step in the fusion of an enveloped virus with a cell is the fusion of their lipid membranes. It is believed that this occurs in several successive stages, one of which includes the assembly of the contiguous outer lipid leaflets of the membranes to constitute an intermediate stalk. The stalk is characterized by a negative spontaneous curvature, corresponding to this formation via the cone-shaped lipids of an inverted hexagonal phase (HII) [509,510]. The induction of positive curvature stress by putative antiviral agents is believed to prevent the generation of fusion intermediates of negative curvature (Figure 8C).
Lipophosphoglycan dramatically reduced the fusion of Sendai virus and IVA with host cells [511,512], while it raised the bilayer-to-HII-phase transition temperature of phosphatidyletanolamine, indicating the elevation of positive curvature stress by lipophosphoglycan [511].
Naturally occurring and synthetic lipopeptides appear to be the most promising candidates when taking into account their amphiphilicity and cone shape, which suggests the induction of positive curvature when incorporated into a lipid monolayer. A simple lipopeptide sequence, myr-WD, was shown to successfully combat IVA and murine coronavirus infections by modulating the membrane lipid packing and surface potential [513]. Surfactin, a cyclic lipopeptide from B. subtilis, was found to inhibit porcine epidemic diarrhea virus and transmissible gastroenteritis virus infections via affecting curvature stress [514]. The dependence of the efficiency of surfactins to inhibit VSV, HSV-1, and Semliki forest virus on the length of the hydrocarbon “tail” was in good agreement with their membrane targeting [515]. Recently, the ability of several lipopeptides to successfully inhibit SARS-CoV-2 fusion with Vero cells was shown [516] (Table 10). The most effective compounds were also characterized by their marked ability to increase the transition temperature of phosphatidylethanolamine from the lamellar to HII phase, i.e., induce a positive curvature stress [516]. Interestingly, simpler molecules, such as black pepper alkaloid piperine, show a similar ability to suppress SARS-CoV-2 infection [517].

5. Conclusions

(i)
Due to principal differences in the organization of fatty acid synthase systems in bacteria and mammals, the specific inhibitors of bacterial key enzymes, especially the acetyl-CoA-carboxylase complex, various β-ketoacyl-ACP synthases, different NADPH-dependent reductases, β-hydroxyacyl-ACP dehydrases, and acyl-phosphate:glycerol-3-phosphate acyltransferase, are attractive targets for the development of low-toxicity antibacterials.
(ii)
The pathway for the synthesis of the lipid fatty acid tails in fungi is similar to that in mammalian cells and, therefore, is not very promising in the search for potential antifungals.
(iii)
The presence of a single fundamental pathway for the synthesis of the phospholipid heads in both prokaryotes and eukaryotes makes the majority of the involved enzymes poor targets for antibiotic therapy in bacterial and fungal infections.
(iv)
Many enzymes of the lipopolysaccharide (Kdo2-lipid A) biosynthetic pathway in Gram-negative bacteria (UDP-N-acetylglucosamine acyltransferase, UDP-3-O-(R-3-hydroxyacyl)glucosamine N-acyltransferase, UDP-3-O-(R-3-hydroxyacyl)-N-acetylglucosamine deacetylase, and UDP-diacylglucosamine pyrophosphohydrolase) are identified as targets for antibiotic development.
(v)
Sphingolipid biosynthetic pathways are conserved from yeast to humans, and the enzymes cannot serve as targets for low-toxicity antifungals. Some inhibitors of inositol-phosphoceramide synthase demonstrate promisingly low effective concentrations.
(vi)
The most effective approach when targeting fungal lipid biosynthesis is to search for inhibitors of enzymes in the ergosterol pathway, especially squalene epoxidase, lanosterol 14α-demethylase, and sterol C14-reductase/sterol C8,7-isomerase.
(vii)
A preference given to inhibitors that simultaneously act on two enzymes of the lipid biosynthetic pathway or the combination of inhibitors with agents directly affecting the pathogen membrane should reduce the risk of developing antibiotic resistance in pathogenic strains.
(viii)
Natural antimicrobial agents exert their defensive activities via pathogen membrane disruption due to pore formation or the disordering of membrane lipids. Due to the high efficiency of naturally occurring antimicrobial agents, their broad-spectrum antibacterial/antifungal/antiviral effect, and their low rate of resistance in pathogen strains, the use of antimicrobial peptides, lipopeptides, and polyenes is a good anti-infective therapeutic strategy.
(ix)
The lipid envelope of viruses should be considered as a target for innovative antivirals, disrupting the membranes of virions or inducing curvature stress and inhibiting viral entry.

Author Contributions

Conceptualization, S.S.E. and O.S.O.; formal analysis, S.S.E. and O.S.O.; investigation, S.S.E. and O.S.O.; resources, S.S.E. and O.S.O.; data curation, S.S.E. and O.S.O.; writing—original draft preparation, S.S.E. and O.S.O.; writing—review and editing, O.S.O.; supervision, O.S.O.; project administration, O.S.O.; funding acquisition, O.S.O. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the Russian Foundation of Science # 22-15-00417.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kim, D.; Kim, S.; Kwon, Y.; Kim, Y.; Park, H.; Kwak, K.; Lee, H.; Lee, J.H.; Jang, K.M.; Kim, D.; et al. Structural Insights for β-Lactam Antibiotics. Biomol. Ther. 2023, 31, 141–147. [Google Scholar] [CrossRef] [PubMed]
  2. Dumas, F.; Haanappel, E. Lipids in infectious diseases—The case of AIDS and tuberculosis. Biochim. Biophys. Acta Biomembr. 2017, 1859, 1636–1647. [Google Scholar] [CrossRef]
  3. Mochalkin, I.; Miller, J.R.; Narasimhan, L.; Thanabal, V.; Erdman, P.; Cox, P.B.; Prasad, J.V.; Lightle, S.; Huband, M.D.; Stover, C.K. Discovery of antibacterial biotin carboxylase inhibitors by virtual screening and fragment-based approaches. ACS Chem. Biol. 2009, 4, 473–483. [Google Scholar] [CrossRef]
  4. Cheng, C.C.; Shipps, G.W., Jr.; Yang, Z.; Sun, B.; Kawahata, N.; Soucy, K.A.; Soriano, A.; Orth, P.; Xiao, L.; Mann, P.; et al. Discovery and optimization of antibacterial AccC inhibitors. Bioorg. Med. Chem. Lett. 2009, 19, 6507–6514. [Google Scholar] [CrossRef]
  5. Freiberg, C.; Pohlmann, J.; Nell, P.G.; Endermann, R.; Schuhmacher, J.; Newton, B.; Otteneder, M.; Lampe, T.; Häbich, D.; Ziegelbauer, K. Novel bacterial acetyl coenzyme A carboxylase inhibitors with antibiotic efficacy in vivo. Antimicrob. Agents Chemother. 2006, 50, 2707–2712. [Google Scholar] [CrossRef]
  6. Khandekar, S.S.; Gentry, D.R.; Van Aller, G.S.; Warren, P.; Xiang, H.; Silverman, C.; Doyle, M.L.; Chambers, P.A.; Konstantinidis, A.K.; Brandt, M.; et al. Identification, substrate specificity, and inhibition of the Streptococcus pneumoniae beta-ketoacyl-acyl carrier protein synthase III (FabH). J. Biol. Chem. 2001, 276, 30024–30030. [Google Scholar] [CrossRef]
  7. Choi, K.H.; Kremer, L.; Besra, G.S.; Rock, C.O. Identification and substrate specificity of beta -ketoacyl (acyl carrier protein) synthase III (mtFabH) from Mycobacterium tuberculosis. J. Biol. Chem. 2000, 275, 28201–28207. [Google Scholar] [CrossRef] [PubMed]
  8. Price, A.C.; Choi, K.H.; Heath, R.J.; Li, Z.; White, S.W.; Rock, C.O. Inhibition of beta-ketoacyl-acyl carrier protein synthases by thiolactomycin and cerulenin. Structure and mechanism. J. Biol. Chem. 2001, 276, 6551–6559. [Google Scholar] [CrossRef] [PubMed]
  9. Tsay, J.T.; Rock, C.O.; Jackowski, S. Overproduction of beta-ketoacyl-acyl carrier protein synthase I imparts thiolactomycin resistance to Escherichia coli K-12. J. Bacteriol. 1992, 174, 508–513. [Google Scholar] [CrossRef]
  10. Wang, J.; Kodali, S.; Lee, S.H.; Galgoci, A.; Painter, R.; Dorso, K.; Racine, F.; Motyl, M.; Hernandez, L.; Tinney, E.; et al. Discovery of platencin, a dual FabF and FabH inhibitor with in vivo antibiotic properties. Proc. Natl. Acad. Sci. USA 2007, 104, 7612–7616. [Google Scholar] [CrossRef] [PubMed]
  11. Jayasuriya, H.; Herath, K.B.; Zhang, C.; Zink, D.L.; Basilio, A.; Genilloud, O.; Diez, M.T.; Vicente, F.; Gonzalez, I.; Salazar, O.; et al. Isolation and structure of platencin: A FabH and FabF dual inhibitor with potent broad-spectrum antibiotic activity. Angew. Chem. Int. Ed. Engl. 2007, 46, 4684–4688. [Google Scholar] [CrossRef]
  12. Wang, J.; Soisson, S.M.; Young, K.; Shoop, W.; Kodali, S.; Galgoci, A.; Painter, R.; Parthasarathy, G.; Tang, Y.S.; Cummings, R.; et al. Platensimycin is a selective FabF inhibitor with potent antibiotic properties. Nature 2006, 441, 358–361. [Google Scholar] [CrossRef]
  13. Zhang, Y.M.; Rock, C.O. Evaluation of epigallocatechin gallate and related plant polyphenols as inhibitors of the FabG and FabI reductases of bacterial type II fatty-acid synthase. J. Biol. Chem. 2004, 279, 30994–31001. [Google Scholar] [CrossRef]
  14. Tasdemir, D.; Lack, G.; Brun, R.; Rüedi, P.; Scapozza, L.; Perozzo, R. Inhibition of Plasmodium falciparum fatty acid biosynthesis: Evaluation of FabG, FabZ, and FabI as drug targets for flavonoids. J. Med. Chem. 2006, 49, 3345–3353. [Google Scholar] [CrossRef] [PubMed]
  15. Belluti, F.; Perozzo, R.; Lauciello, L.; Colizzi, F.; Kostrewa, D.; Bisi, A.; Gobbi, S.; Rampa, A.; Bolognesi, M.L.; Recanatini, M.; et al. Design, synthesis, and biological and crystallographic evaluation of novel inhibitors of Plasmodium falciparum enoyl-ACP-reductase (PfFabI). J. Med. Chem. 2013, 56, 7516–7526. [Google Scholar] [CrossRef] [PubMed]
  16. Kirmizibekmez, H.; Calis, I.; Perozzo, R.; Brun, R.; Dönmez, A.A.; Linden, A.; Rüedi, P.; Tasdemir, D. Inhibiting activities of the secondary metabolites of Phlomis brunneogaleata against parasitic protozoa and plasmodial enoyl-ACP Reductase, a crucial enzyme in fatty acid biosynthesis. Planta Med. 2004, 70, 711–717. [Google Scholar] [CrossRef]
  17. Sohn, M.J.; Zheng, C.J.; Kim, W.G. Macrolactin S, a new antibacterial agent with FabG-inhibitory activity from Bacillus sp. AT28. J. Antibiot. 2008, 61, 687–691. [Google Scholar] [CrossRef] [PubMed]
  18. Bhowruth, V.; Brown, A.K.; Besra, G.S. Synthesis and biological evaluation of NAS-21 and NAS-91 analogues as potential inhibitors of the mycobacterial FAS-II dehydratase enzyme Rv0636. Microbiology 2008, 154, 1866–1875. [Google Scholar] [CrossRef]
  19. McGillick, B.E.; Kumaran, D.; Vieni, C.; Swaminathan, S. β-Hydroxyacyl-acyl Carrier Protein Dehydratase (FabZ) from Francisella tularensis and Yersinia pestis: Structure Determination, Enzymatic Characterization, and Cross-Inhibition Studies. Biochemistry 2016, 55, 1091–1099. [Google Scholar] [CrossRef]
  20. Chen, J.; Zhang, L.; Zhang, Y.; Zhang, H.; Du, J.; Ding, J.; Guo, Y.; Jiang, H.; Shen, X. Emodin targets the beta-hydroxyacyl-acyl carrier protein dehydratase from Helicobacter pylori: Enzymatic inhibition assay with crystal structural and thermodynamic characterization. BMC Microbiol. 2009, 9, 91. [Google Scholar] [CrossRef]
  21. Kumar, V.; Sharma, A.; Pratap, S.; Kumar, P. Biochemical and biophysical characterization of 1,4-naphthoquinone as a dual inhibitor of two key enzymes of type II fatty acid biosynthesis from Moraxella catarrhalis. Biochim. Biophys. Acta Proteins Proteom. 2018, 1866, 1131–1142. [Google Scholar] [CrossRef] [PubMed]
  22. Kong, Y.H.; Zhang, L.; Yang, Z.Y.; Han, C.; Hu, L.H.; Jiang, H.L.; Shen, X. Natural product juglone targets three key enzymes from Helicobacter pylori: Inhibition assay with crystal structure characterization. Acta Pharmacol. Sin. 2008, 29, 870–876. [Google Scholar] [CrossRef] [PubMed]
  23. Zheng, C.J.; Sohn, M.J.; Lee, S.; Kim, W.G. Meleagrin, a new FabI inhibitor from Penicillium chryosogenum with at least one additional mode of action. PLoS ONE 2013, 8, e78922. [Google Scholar] [CrossRef] [PubMed]
  24. Kim, Y.G.; Seo, J.H.; Kwak, J.H.; Shin, K.J. Discovery of a potent enoyl-acyl carrier protein reductase (FabI) inhibitor suitable for antistaphylococcal agent. Bioorg. Med. Chem. Lett. 2015, 25, 4481–4486. [Google Scholar] [CrossRef] [PubMed]
  25. Kwon, Y.J.; Kim, H.J.; Kim, W.G. Complestatin exerts antibacterial activity by the inhibition of fatty acid synthesis. Biol. Pharm. Bull. 2015, 38, 715–721. [Google Scholar] [CrossRef]
  26. Surolia, N.; Surolia, A. Triclosan offers protection against blood stages of malaria by inhibiting enoyl-ACP reductase of Plasmodium falciparum. Nat. Med. 2001, 7, 167–173. [Google Scholar] [CrossRef]
  27. Yao, J.; Abdelrahman, Y.M.; Robertson, R.M.; Cox, J.V.; Belland, R.J.; White, S.W.; Rock, C.O. Type II fatty acid synthesis is essential for the replication of Chlamydia trachomatis. J. Biol. Chem. 2014, 289, 22365–22376. [Google Scholar] [CrossRef]
  28. Yogiara; Mordukhova, E.A.; Kim, D.; Kim, W.G.; Hwang, J.K.; Pan, J.G. The food-grade antimicrobial xanthorrhizol targets the enoyl-ACP reductase (FabI) in Escherichia coli. Bioorg. Med. Chem. Lett. 2020, 30, 127651. [Google Scholar] [CrossRef]
  29. Cho, J.Y.; Kwon, Y.J.; Sohn, M.J.; Seok, S.J.; Kim, W.G. Phellinstatin, a new inhibitor of enoyl-ACP reductase produced by the medicinal fungus Phellinus linteus. Bioorg. Med. Chem. Lett. 2011, 21, 1716–1718. [Google Scholar] [CrossRef]
  30. Kim, Y.J.; Sohn, M.J.; Kim, W.G. Chalcomoracin and moracin C, new inhibitors of Staphylococcus aureus enoyl-acyl carrier protein reductase from Morus alba. Biol. Pharm. Bull. 2012, 35, 791–795. [Google Scholar] [CrossRef]
  31. Kwon, Y.J.; Fang, Y.; Xu, G.H.; Kim, W.G. Aquastatin A, a new inhibitor of enoyl-acyl carrier protein reductase from Sporothrix sp. FN611. Biol. Pharm. Bull. 2009, 32, 2061–2064. [Google Scholar] [CrossRef]
  32. Zheng, C.J.; Sohn, M.J.; Kim, W.G. Atromentin and leucomelone, the first inhibitors specific to enoyl-ACP reductase (FabK) of Streptococcus pneumoniae. J. Antibiot. 2006, 59, 808–812. [Google Scholar] [CrossRef] [PubMed]
  33. Grimes, K.D.; Lu, Y.J.; Zhang, Y.M.; Luna, V.A.; Hurdle, J.G.; Carson, E.I.; Qi, J.; Kudrimoti, S.; Rock, C.O.; Lee, R.E. Novel acyl phosphate mimics that target PlsY, an essential acyltransferase in gram-positive bacteria. ChemMedChem 2008, 3, 1936–1945. [Google Scholar] [CrossRef] [PubMed]
  34. Cherian, P.T.; Yao, J.; Leonardi, R.; Maddox, M.M.; Luna, V.A.; Rock, C.O.; Lee, R.E. Acyl-sulfamates target the essential glycerol-phosphate acyltransferase (PlsY) in Gram-positive bacteria. Bioorg. Med. Chem. 2012, 20, 4985–4994. [Google Scholar] [CrossRef] [PubMed]
  35. Jiang, X.; Duan, Y.; Zhou, B.; Guo, Q.; Wang, H.; Hang, X.; Zeng, L.; Jia, J.; Bi, H. The Cyclopropane Fatty Acid Synthase Mediates Antibiotic Resistance and Gastric Colonization of Helicobacter pylori. J. Bacteriol. 2019, 201, e00374-19. [Google Scholar] [CrossRef]
  36. Parsons, J.B.; Rock, C.O. Bacterial lipids: Metabolism and membrane homeostasis. Prog. Lipid Res. 2013, 52, 249–276. [Google Scholar] [CrossRef]
  37. Larson, E.C.; Lim, A.L.; Pond, C.D.; Craft, M.; Čavužić, M.; Waldrop, G.L.; Schmidt, E.W.; Barrows, L.R. Pyrrolocin C and equisetin inhibit bacterial acetyl-CoA carboxylase. PLoS ONE 2020, 15, e0233485. [Google Scholar] [CrossRef] [PubMed]
  38. Freiberg, C.; Brunner, N.A.; Schiffer, G.; Lampe, T.; Pohlmann, J.; Brands, M.; Raabe, M.; Häbich, D.; Ziegelbauer, K. Identification and characterization of the first class of potent bacterial acetyl-CoA carboxylase inhibitors with antibacterial activity. J. Biol. Chem. 2004, 279, 26066–26073. [Google Scholar] [CrossRef] [PubMed]
  39. Freiberg, C.; Fischer, H.P.; Brunner, N.A. Discovering the mechanism of action of novel antibacterial agents through transcriptional profiling of conditional mutants. Antimicrob. Agents Chemother. 2005, 49, 749–759. [Google Scholar] [CrossRef]
  40. Pohlmann, J.; Lampe, T.; Shimada, M.; Nell, P.G.; Pernerstorfer, J.; Svenstrup, N.; Brunner, N.A.; Schiffer, G.; Freiberg, C. Pyrrolidinedione derivatives as antibacterial agents with a novel mode of action. Bioorg. Med. Chem. Lett. 2005, 15, 1189–1192. [Google Scholar] [CrossRef]
  41. Toor, H.G.; Banerjee, D.I.; Chauhan, J.B. In Silico Evaluation of Human Cathelicidin LL-37 as a Novel Therapeutic Inhibitor of Panton-Valentine Leukocidin Toxin of Methicillin-Resistant Staphylococcus aureus. Microb. Drug Resist. 2021, 27, 602–615. [Google Scholar] [CrossRef]
  42. Liu, X.; Fortin, P.D.; Walsh, C.T. Andrimid producers encode an acetyl-CoA carboxyltransferase subunit resistant to the action of the antibiotic. Proc. Natl. Acad. Sci. USA 2008, 105, 13321–13326. [Google Scholar] [CrossRef]
  43. Silvers, M.A.; Robertson, G.T.; Taylor, C.M.; Waldrop, G.L. Design, synthesis, and antibacterial properties of dual-ligand inhibitors of acetyl-CoA carboxylase. J. Med. Chem. 2014, 57, 8947–8959. [Google Scholar] [CrossRef] [PubMed]
  44. Zhang, W.; Wei, S.; Wu, W. Preliminary studies on the antibacterial mechanism of Yanglingmycin. Pestic. Biochem. Physiol. 2018, 147, 27–31. [Google Scholar] [CrossRef] [PubMed]
  45. Kuldeep, J.; Sharma, S.K.; Singh, B.N.; Siddiqi, M.I. Computational exploration and anti-mycobacterial activity of potential inhibitors of Mycobacterium tuberculosis acetyl coenzyme A carboxylase as anti-tubercular agents. SAR QSAR Environ. Res. 2021, 32, 191–205. [Google Scholar] [CrossRef] [PubMed]
  46. Kumar, V.; Sharma, A.; Pratap, S.; Kumar, P. Biophysical and in silico interaction studies of aporphine alkaloids with Malonyl-CoA: ACP transacylase (FabD) from drug resistant Moraxella catarrhalis. Biochimie 2018, 149, 18–33. [Google Scholar] [CrossRef] [PubMed]
  47. Heath, R.J.; Rock, C.O. Inhibition of beta-ketoacyl-acyl carrier protein synthase III (FabH) by acyl-acyl carrier protein in Escherichia coli. J. Biol. Chem. 1996, 271, 10996–11000. [Google Scholar] [CrossRef] [PubMed]
  48. Choi, K.H.; Heath, R.J.; Rock, C.O. Beta-ketoacyl-acyl carrier protein synthase III (FabH) is a determining factor in branched-chain fatty acid biosynthesis. J. Bacteriol. 2000, 182, 365–370. [Google Scholar] [CrossRef] [PubMed]
  49. Qiu, X.; Choudhry, A.E.; Janson, C.A.; Grooms, M.; Daines, R.A.; Lonsdale, J.T.; Khandekar, S.S. Crystal structure and substrate specificity of the beta-ketoacyl-acyl carrier protein synthase III (FabH) from Staphylococcus aureus. Protein Sci. 2005, 14, 2087–2094. [Google Scholar] [CrossRef]
  50. Musayev, F.; Sachdeva, S.; Scarsdale, J.N.; Reynolds, K.A.; Wright, H.T. Crystal structure of a substrate complex of Mycobacterium tuberculosis beta-ketoacyl-acyl carrier protein synthase III (FabH) with lauroyl-coenzyme A. J. Mol. Biol. 2005, 346, 1313–1321. [Google Scholar] [CrossRef]
  51. Luo, Y.; Yang, Y.S.; Fu, J.; Zhu, H.L. Novel FabH inhibitors: A patent and article literature review (2000—2012). Expert. Opin. Ther. Pat. 2012, 22, 1325–1336. [Google Scholar] [CrossRef]
  52. Wallace, K.K.; Lobo, S.; Han, L.; McArthur, H.A.; Reynolds, K.A. In vivo and In vitro effects of thiolactomycin on fatty acid biosynthesis in Streptomyces collinus. J. Bacteriol. 1997, 179, 3884–3891. [Google Scholar] [CrossRef]
  53. Han, L.; Lobo, S.; Reynolds, K.A. Characterization of beta-ketoacyl-acyl carrier protein synthase III from Streptomyces glaucescens and its role in initiation of fatty acid biosynthesis. J. Bacteriol. 1998, 180, 4481–4486. [Google Scholar] [CrossRef]
  54. Alhamadsheh, M.M.; Waters, N.C.; Huddler, D.P.; Kreishman-Deitrick, M.; Florova, G.; Reynolds, K.A. Synthesis and biological evaluation of thiazolidine-2-one 1,1-dioxide as inhibitors of Escherichia coli beta-ketoacyl-ACP-synthase III (FabH). Bioorg. Med. Chem. Lett. 2007, 17, 879–883. [Google Scholar] [CrossRef]
  55. Li, H.Q.; Shi, L.; Li, Q.S.; Liu, P.G.; Luo, Y.; Zhao, J.; Zhu, H.L. Synthesis of C(7) modified chrysin derivatives designing to inhibit beta-ketoacyl-acyl carrier protein synthase III (FabH) as antibiotics. Bioorg. Med. Chem. 2009, 17, 6264–6269. [Google Scholar] [CrossRef] [PubMed]
  56. Lv, P.C.; Wang, K.R.; Yang, Y.; Mao, W.J.; Chen, J.; Xiong, J.; Zhu, H.L. Design, synthesis and biological evaluation of novel thiazole derivatives as potent FabH inhibitors. Bioorg. Med. Chem. Lett. 2009, 19, 6750–6754. [Google Scholar] [CrossRef] [PubMed]
  57. Li, H.Q.; Luo, Y.; Lv, P.C.; Shi, L.; Liu, C.H.; Zhu, H.L. Design and synthesis of novel deoxybenzoin derivatives as FabH inhibitors and anti-inflammatory agents. Bioorg. Med. Chem. Lett. 2010, 20, 2025–2028. [Google Scholar] [CrossRef] [PubMed]
  58. Cheng, K.; Zheng, Q.Z.; Qian, Y.; Shi, L.; Zhao, J.; Zhu, H.L. Synthesis, antibacterial activities and molecular docking studies of peptide and Schiff bases as targeted antibiotics. Bioorg. Med. Chem. 2009, 17, 7861–7871. [Google Scholar] [CrossRef] [PubMed]
  59. Shi, L.; Fang, R.Q.; Zhu, Z.W.; Yang, Y.; Cheng, K.; Zhong, W.Q.; Zhu, H.L. Design and synthesis of potent inhibitors of beta-ketoacyl-acyl carrier protein synthase III (FabH) as potential antibacterial agents. Eur. J. Med. Chem. 2010, 45, 4358–4364. [Google Scholar] [CrossRef] [PubMed]
  60. Cheng, K.; Zheng, Q.Z.; Hou, J.; Zhou, Y.; Liu, C.H.; Zhao, J.; Zhu, H.L. Synthesis, molecular modeling and biological evaluation of PSB as targeted antibiotics. Bioorg. Med. Chem. 2010, 18, 2447–2455. [Google Scholar] [CrossRef] [PubMed]
  61. Lv, P.C.; Sun, J.; Luo, Y.; Yang, Y.; Zhu, H.L. Design, synthesis, and structure-activity relationships of pyrazole derivatives as potential FabH inhibitors. Bioorg. Med. Chem. Lett. 2010, 20, 4657–4660. [Google Scholar] [CrossRef]
  62. Zhang, H.J.; Zhu, D.D.; Li, Z.L.; Sun, J.; Zhu, H.L. Synthesis, molecular modeling and biological evaluation of β-ketoacyl-acyl carrier protein synthase III (FabH) as novel antibacterial agents. Bioorg. Med. Chem. 2011, 19, 4513–4519. [Google Scholar] [CrossRef]
  63. Li, H.Q.; Luo, Y.; Zhu, H.L. Discovery of vinylogous carbamates as a novel class of β-ketoacyl-acyl carrier protein synthase III (FabH) inhibitors. Bioorg. Med. Chem. 2011, 19, 4454–4459. [Google Scholar] [CrossRef]
  64. Li, Z.L.; Li, Q.S.; Zhang, H.J.; Hu, Y.; Zhu, D.D.; Zhu, H.L. Design, synthesis and biological evaluation of urea derivatives from o-hydroxybenzylamines and phenylisocyanate as potential FabH inhibitors. Bioorg. Med. Chem. 2011, 19, 4413–4420. [Google Scholar] [CrossRef]
  65. Luo, Y.; Zhang, L.R.; Hu, Y.; Zhang, S.; Fu, J.; Wang, X.M.; Zhu, H.L. Synthesis and antimicrobial activities of oximes derived from O-benzylhydroxylamine as FabH inhibitors. ChemMedChem 2012, 7, 1587–1593. [Google Scholar] [CrossRef] [PubMed]
  66. Zhou, Y.; Du, Q.R.; Sun, J.; Li, J.R.; Fang, F.; Li, D.D.; Qian, Y.; Gong, H.B.; Zhao, J.; Zhu, H.L. Novel Schiff-base-derived FabH inhibitors with dioxygenated rings as antibiotic agents. ChemMedChem 2013, 8, 433–441. [Google Scholar] [CrossRef] [PubMed]
  67. Wang, X.L.; Zhang, Y.B.; Tang, J.F.; Yang, Y.S.; Chen, R.Q.; Zhang, F.; Zhu, H.L. Design, synthesis and antibacterial activities of vanillic acylhydrazone derivatives as potential β-ketoacyl-acyl carrier protein synthase III (FabH) inhibitors. Eur. J. Med. Chem. 2012, 57, 373–382. [Google Scholar] [CrossRef] [PubMed]
  68. Li, Y.; Luo, Y.; Hu, Y.; Zhu, D.D.; Zhang, S.; Liu, Z.J.; Gong, H.B.; Zhu, H.L. Design, synthesis and antimicrobial activities of nitroimidazole derivatives containing 1,3,4-oxadiazole scaffold as FabH inhibitors. Bioorg. Med. Chem. 2012, 20, 4316–4322. [Google Scholar] [CrossRef] [PubMed]
  69. Yang, Y.S.; Zhang, F.; Gao, C.; Zhang, Y.B.; Wang, X.L.; Tang, J.F.; Sun, J.; Gong, H.B.; Zhu, H.L. Discovery and modification of sulfur-containing heterocyclic pyrazoline derivatives as potential novel class of β-ketoacyl-acyl carrier protein synthase III (FabH) inhibitors. Bioorg. Med. Chem. Lett. 2012, 22, 4619–4624. [Google Scholar] [CrossRef]
  70. Li, Y.; Zhao, C.P.; Ma, H.P.; Zhao, M.Y.; Xue, Y.R.; Wang, X.M.; Zhu, H.L. Design, synthesis and antimicrobial activities evaluation of Schiff base derived from secnidazole derivatives as potential FabH inhibitors. Bioorg. Med. Chem. 2013, 21, 3120–3126. [Google Scholar] [CrossRef] [PubMed]
  71. Li, J.R.; Li, D.D.; Wang, R.R.; Sun, J.; Dong, J.J.; Du, Q.R.; Fang, F.; Zhang, W.M.; Zhu, H.L. Design and synthesis of thiazole derivatives as potent FabH inhibitors with antibacterial activity. Eur. J. Med. Chem. 2014, 75, 438–447. [Google Scholar] [CrossRef] [PubMed]
  72. Duan, Y.T.; Wang, Z.C.; Sang, Y.L.; Tao, X.X.; Teraiya, S.B.; Wang, P.F.; Wen, Q.; Zhou, X.J.; Ding, L.; Yang, Y.H.; et al. Design and synthesis of 2-styryl of 5-Nitroimidazole derivatives and antimicrobial activities as FabH inhibitors. Eur. J. Med. Chem. 2014, 76, 387–396. [Google Scholar] [CrossRef] [PubMed]
  73. Song, X.; Yang, Y.; Zhao, J.; Chen, Y. Synthesis and antibacterial activity of cinnamaldehyde acylhydrazone with a 1,4-benzodioxan fragment as a novel class of potent β-ketoacyl-acyl carrier protein synthase III (FabH) inhibitor. Chem. Pharm. Bull. 2014, 62, 1110–1118. [Google Scholar] [CrossRef]
  74. Segretti, N.D.; Serafim, R.A.; Segretti, M.C.; Miyata, M.; Coelho, F.R.; Augusto, O.; Ferreira, E.I. New antibacterial agents: Hybrid bioisoster derivatives as potential E. coli FabH inhibitors. Bioorg. Med. Chem. Lett. 2016, 26, 3988–3993. [Google Scholar] [CrossRef]
  75. Zhou, Y.; Yang, Y.S.; Song, X.D.; Lu, L.; Zhu, H.L. Study of Schiff-Base-Derived with Dioxygenated Rings and Nitrogen Heterocycle as Potential β-Ketoacyl-acyl Carrier Protein Synthase III (FabH) Inhibitors. Chem. Pharm. Bull. 2017, 65, 178–185. [Google Scholar] [CrossRef] [PubMed]
  76. He, X.; Reynolds, K.A. Purification, characterization, and identification of novel inhibitors of the beta-ketoacyl-acyl carrier protein synthase III (FabH) from Staphylococcus aureus. Antimicrob. Agents Chemother. 2002, 46, 1310–1318. [Google Scholar] [CrossRef]
  77. He, X.; Reeve, A.M.; Desai, U.R.; Kellogg, G.E.; Reynolds, K.A. 1,2-dithiole-3-ones as potent inhibitors of the bacterial 3-ketoacyl acyl carrier protein synthase III (FabH). Antimicrob. Agents Chemother. 2004, 48, 3093–3102. [Google Scholar] [CrossRef]
  78. Pishchany, G.; Mevers, E.; Ndousse-Fetter, S.; Horvath, D.J., Jr.; Paludo, C.R.; Silva-Junior, E.A.; Koren, S.; Skaar, E.P.; Clardy, J.; Kolter, R. Amycomicin is a potent and specific antibiotic discovered with a targeted interaction screen. Proc. Natl. Acad. Sci. USA 2018, 115, 10124–10129. [Google Scholar] [CrossRef]
  79. Singh, S.; Soni, L.K.; Gupta, M.K.; Prabhakar, Y.S.; Kaskhedikar, S.G. QSAR studies on benzoylaminobenzoic acid derivatives as inhibitors of beta-ketoacyl-acyl carrier protein synthase III. Eur. J. Med. Chem. 2008, 43, 1071–1080. [Google Scholar] [CrossRef]
  80. Nie, Z.; Perretta, C.; Lu, J.; Su, Y.; Margosiak, S.; Gajiwala, K.S.; Cortez, J.; Nikulin, V.; Yager, K.M.; Appelt, K.; et al. Structure-based design, synthesis, and study of potent inhibitors of beta-ketoacyl-acyl carrier protein synthase III as potential antimicrobial agents. J. Med. Chem. 2005, 48, 1596–1609. [Google Scholar] [CrossRef]
  81. Ashek, A.; Cho, S.J. A combined approach of docking and 3D QSAR study of beta-ketoacyl-acyl carrier protein synthase III (FabH) inhibitors. Bioorg. Med. Chem. 2006, 14, 1474–1482. [Google Scholar] [CrossRef]
  82. Jones, P.B.; Parrish, N.M.; Houston, T.A.; Stapon, A.; Bansal, N.P.; Dick, J.D.; Townsend, C.A. A new class of antituberculosis agents. J. Med. Chem. 2000, 43, 3304–3314. [Google Scholar] [CrossRef] [PubMed]
  83. Alhamadsheh, M.M.; Musayev, F.; Komissarov, A.A.; Sachdeva, S.; Wright, H.T.; Scarsdale, N.; Florova, G.; Reynolds, K.A. Alkyl-CoA disulfides as inhibitors and mechanistic probes for FabH enzymes. Chem. Biol. 2007, 14, 513–524. [Google Scholar] [CrossRef] [PubMed]
  84. Liu, Y.; Zhong, W.; Li, R.J.; Li, S. Synthesis of potent inhibitors of β-ketoacyl-acyl carrier protein synthase III as potential antimicrobial agents. Molecules 2012, 17, 4770–4781. [Google Scholar] [CrossRef]
  85. Borgaro, J.G.; Chang, A.; Machutta, C.A.; Zhang, X.; Tonge, P.J. Substrate recognition by β-ketoacyl-ACP synthases. Biochemistry 2011, 50, 10678–10686. [Google Scholar] [CrossRef] [PubMed]
  86. Huang, W.; Jia, J.; Edwards, P.; Dehesh, K.; Schneider, G.; Lindqvist, Y. Crystal structure of beta-ketoacyl-acyl carrier protein synthase II from E.coli reveals the molecular architecture of condensing enzymes. EMBO J. 1998, 17, 1183–1191. [Google Scholar] [CrossRef]
  87. Qiu, X.; Janson, C.A.; Konstantinidis, A.K.; Nwagwu, S.; Silverman, C.; Smith, W.W.; Khandekar, S.; Lonsdale, J.; Abdel-Meguid, S.S. Crystal structure of beta-ketoacyl-acyl carrier protein synthase III. A key condensing enzyme in bacterial fatty acid biosynthesis. J. Biol. Chem. 1999, 274, 36465–36471. [Google Scholar] [CrossRef]
  88. Olsen, J.G.; Kadziola, A.; von Wettstein-Knowles, P.; Siggaard-Andersen, M.; Lindquist, Y.; Larsen, S. The X-ray crystal structure of beta-ketoacyl [acyl carrier protein] synthase I. FEBS Lett. 1999, 460, 46–52. [Google Scholar] [CrossRef]
  89. Davies, C.; Heath, R.J.; White, S.W.; Rock, C.O. The 1.8 A crystal structure and active-site architecture of beta-ketoacyl-acyl carrier protein synthase III (FabH) from Escherichia coli. Structure 2000, 8, 185–195. [Google Scholar] [CrossRef]
  90. Heath, R.J.; White, S.W.; Rock, C.O. Inhibitors of fatty acid synthesis as antimicrobial chemotherapeutics. Appl. Microbiol. Biotechnol. 2002, 58, 695–703. [Google Scholar] [CrossRef]
  91. Bommineni, G.R.; Kapilashrami, K.; Cummings, J.E.; Lu, Y.; Knudson, S.E.; Gu, C.; Walker, S.G.; Slayden, R.A.; Tonge, P.J. Thiolactomycin-Based Inhibitors of Bacterial β-Ketoacyl-ACP Synthases with in Vivo Activity. J. Med. Chem. 2016, 59, 5377–5390. [Google Scholar] [CrossRef] [PubMed]
  92. Rudolf, J.D.; Dong, L.B.; Manoogian, K.; Shen, B. Biosynthetic Origin of the Ether Ring in Platensimycin. J. Am. Chem. Soc. 2016, 138, 16711–16721. [Google Scholar] [CrossRef] [PubMed]
  93. Manallack, D.T.; Crosby, I.T.; Khakham, Y.; Capuano, B. Platensimycin: A promising antimicrobial targeting fatty acid synthesis. Curr. Med. Chem. 2008, 15, 705–710. [Google Scholar] [CrossRef] [PubMed]
  94. Moustafa, G.A.I.; Nojima, S.; Yamano, Y.; Aono, A.; Arai, M.; Mitarai, S.; Tanaka, T.; Yoshimitsu, T. Potent growth inhibitory activity of (±)-platencin towards multi-drug-resistant and extensively drug-resistant Mycobacterium tuberculosis. Med. Chem. Commun. 2013, 4, 720–723. [Google Scholar] [CrossRef]
  95. Brown, A.K.; Taylor, R.C.; Bhatt, A.; Fütterer, K.; Besra, G.S. Platensimycin activity against mycobacterial beta-ketoacyl-ACP synthases. PLoS ONE 2009, 4, e6306. [Google Scholar] [CrossRef] [PubMed]
  96. Das, M.; Sakha Ghosh, P.; Manna, K. A Review on Platensimycin: A Selective FabF Inhibitor. Int. J. Med. Chem. 2016, 2016, 9706753. [Google Scholar] [CrossRef] [PubMed]
  97. Martens, E.; Demain, A.L. Platensimycin and platencin: Promising antibiotics for future application in human medicine. J. Antibiot. 2011, 64, 705–710. [Google Scholar] [CrossRef]
  98. Shang, R.; Liang, J.; Yi, Y.; Liu, Y.; Wang, J. Review of Platensimycin and Platencin: Inhibitors of β-Ketoacyl-acyl Carrier Protein (ACP) Synthase III (FabH). Molecules 2015, 20, 16127–16141. [Google Scholar] [CrossRef]
  99. Su, M.; Qiu, L.; Deng, Y.; Ruiz, C.H.; Rudolf, J.D.; Dong, L.B.; Feng, X.; Cameron, M.D.; Shen, B.; Duan, Y.; et al. Evaluation of Platensimycin and Platensimycin-Inspired Thioether Analogues against Methicillin-Resistant Staphylococcus aureus in Topical and Systemic Infection Mouse Models. Mol. Pharm. 2019, 16, 3065–3071. [Google Scholar] [CrossRef]
  100. Deng, Y.; Weng, X.; Li, Y.; Su, M.; Wen, Z.; Ji, X.; Ren, N.; Shen, B.; Duan, Y.; Huang, Y. Late-Stage Functionalization of Platensimycin Leading to Multiple Analogues with Improved Antibacterial Activity In vitro and in Vivo. J. Med. Chem. 2019, 62, 6682–6693. [Google Scholar] [CrossRef]
  101. Feng, Z.; Chakraborty, D.; Dewell, S.B.; Reddy, B.V.; Brady, S.F. Environmental DNA-encoded antibiotics fasamycins A and B inhibit FabF in type II fatty acid biosynthesis. J. Am. Chem. Soc. 2012, 134, 2981–2987. [Google Scholar] [CrossRef]
  102. Zheng, Z.; Parsons, J.B.; Tangallapally, R.; Zhang, W.; Rock, C.O.; Lee, R.E. Discovery of novel bacterial elongation condensing enzyme inhibitors by virtual screening. Bioorg. Med. Chem. Lett. 2014, 24, 2585–2588. [Google Scholar] [CrossRef]
  103. Wickramasinghe, S.R.; Inglis, K.A.; Urch, J.E.; Müller, S.; van Aalten, D.M.; Fairlamb, A.H. Kinetic, inhibition and structural studies on 3-oxoacyl-ACP reductase from Plasmodium falciparum, a key enzyme in fatty acid biosynthesis. Biochem. J. 2006, 393, 447–457. [Google Scholar] [CrossRef]
  104. Kristan, K.; Bratkovic, T.; Sova, M.; Gobec, S.; Prezelj, A.; Urleb, U. Novel inhibitors of beta-ketoacyl-ACP reductase from Escherichia coli. Chem. Biol. Interact. 2009, 178, 310–316. [Google Scholar] [CrossRef] [PubMed]
  105. Vella, P.; Rudraraju, R.S.; Lundbäck, T.; Axelsson, H.; Almqvist, H.; Vallin, M.; Schneider, G.; Schnell, R. A FabG inhibitor targeting an allosteric binding site inhibits several orthologs from Gram-negative ESKAPE pathogens. Bioorg. Med. Chem. 2021, 30, 115898. [Google Scholar] [CrossRef] [PubMed]
  106. Cukier, C.D.; Hope, A.G.; Elamin, A.A.; Moynie, L.; Schnell, R.; Schach, S.; Kneuper, H.; Singh, M.; Naismith, J.H.; Lindqvist, Y.; et al. Discovery of an allosteric inhibitor binding site in 3-Oxo-acyl-ACP reductase from Pseudomonas aeruginosa. ACS Chem. Biol. 2013, 8, 2518–2527. [Google Scholar] [CrossRef]
  107. Hu, J.; Webster, D.; Cao, J.; Shao, A. The safety of green tea and green tea extract consumption in adults—Results of a systematic review. Regul. Toxicol. Pharmacol. 2018, 95, 412–433. [Google Scholar] [CrossRef] [PubMed]
  108. Zakharova, A.A.; Efimova, S.S.; Ostroumova, O.S. Lipid Microenvironment Modulates the Pore-Forming Ability of Polymyxin B. Antibiotics 2022, 11, 1445. [Google Scholar] [CrossRef]
  109. Chernyshova, D.N.; Tyulin, A.A.; Ostroumova, O.S.; Efimova, S.S. Discovery of the Potentiator of the Pore-Forming Ability of Lantibiotic Nisin: Perspectives for Anticancer Therapy. Membranes 2022, 12, 1166. [Google Scholar] [CrossRef] [PubMed]
  110. Efimova, S.S.; Malykhina, A.I.; Ostroumova, O.S. Triggering the Amphotericin B Pore-Forming Activity by Phytochemicals. Membranes 2023, 13, 670. [Google Scholar] [CrossRef]
  111. Sharma, S.K.; Kapoor, M.; Ramya, T.N.; Kumar, S.; Kumar, G.; Modak, R.; Sharma, S.; Surolia, N.; Surolia, A. Identification, characterization, and inhibition of Plasmodium falciparum beta-hydroxyacyl-acyl carrier protein dehydratase (FabZ). J. Biol. Chem. 2003, 278, 45661–45671. [Google Scholar] [CrossRef]
  112. He, L.; Zhang, L.; Liu, X.; Li, X.; Zheng, M.; Li, H.; Yu, K.; Chen, K.; Shen, X.; Jiang, H.; et al. Discovering potent inhibitors against the beta-hydroxyacyl-acyl carrier protein dehydratase (FabZ) of Helicobacter pylori: Structure-based design, synthesis, bioassay, and crystal structure determination. J. Med. Chem. 2009, 52, 2465–2481. [Google Scholar] [CrossRef]
  113. Saling, S.C.; Comar, J.F.; Mito, M.S.; Peralta, R.M.; Bracht, A. Actions of juglone on energy metabolism in the rat liver. Toxicol. Appl. Pharmacol. 2011, 257, 319–327. [Google Scholar] [CrossRef]
  114. Zhu, K.; Choi, K.H.; Schweizer, H.P.; Rock, C.O.; Zhang, Y.M. Two aerobic pathways for the formation of unsaturated fatty acids in Pseudomonas aeruginosa. Mol. Microbiol. 2006, 60, 260–273. [Google Scholar] [CrossRef] [PubMed]
  115. Leesong, M.; Henderson, B.S.; Gillig, J.R.; Schwab, J.M.; Smith, J.L. Structure of a dehydratase-isomerase from the bacterial pathway for biosynthesis of unsaturated fatty acids: Two catalytic activities in one active site. Structure 1996, 4, 253–264. [Google Scholar] [CrossRef] [PubMed]
  116. Clark, D.P.; DeMendoza, D.; Polacco, M.L.; Cronan, J.E., Jr. Beta-hydroxydecanoyl thio ester dehydrase does not catalyze a rate-limiting step in Escherichia coli unsaturated fatty acid synthesis. Biochemistry 1983, 22, 5897–5902. [Google Scholar] [CrossRef] [PubMed]
  117. Marrakchi, H.; Choi, K.H.; Rock, C.O. A new mechanism for anaerobic unsaturated fatty acid formation in Streptococcus pneumoniae. J. Biol. Chem. 2002, 277, 44809–44816. [Google Scholar] [CrossRef]
  118. Fozo, E.M.; Quivey, R.G., Jr. The fabM gene product of Streptococcus mutans is responsible for the synthesis of monounsaturated fatty acids and is necessary for survival at low pH. J. Bacteriol. 2004, 186, 4152–4158. [Google Scholar] [CrossRef]
  119. Wang, H.; Cronan, J.E. Functional replacement of the FabA and FabB proteins of Escherichia coli fatty acid synthesis by Enterococcus faecalis FabZ and FabF homologues. J. Biol. Chem. 2004, 279, 34489–34495. [Google Scholar] [CrossRef]
  120. Bi, H.; Wang, H.; Cronan, J.E. FabQ, a dual-function dehydratase/isomerase, circumvents the last step of the classical fatty acid synthesis cycle. Chem. Biol. 2013, 20, 1157–1167. [Google Scholar] [CrossRef]
  121. Altabe, S.G.; Aguilar, P.; Caballero, G.M.; de Mendoza, D. The Bacillus subtilis acyl lipid desaturase is a delta5 desaturase. J. Bacteriol. 2003, 185, 3228–3231. [Google Scholar] [CrossRef]
  122. Moynié, L.; Leckie, S.M.; McMahon, S.A.; Duthie, F.G.; Koehnke, A.; Taylor, J.W.; Alphey, M.S.; Brenk, R.; Smith, A.D.; Naismith, J.H. Structural insights into the mechanism and inhibition of the β-hydroxydecanoyl-acyl carrier protein dehydratase from Pseudomonas aeruginosa. J. Mol. Biol. 2013, 425, 365–377. [Google Scholar] [CrossRef]
  123. Moynié, L.; Hope, A.G.; Finzel, K.; Schmidberger, J.; Leckie, S.M.; Schneider, G.; Burkart, M.D.; Smith, A.D.; Gray, D.W.; Naismith, J.H. A Substrate Mimic Allows High-Throughput Assay of the FabA Protein and Consequently the Identification of a Novel Inhibitor of Pseudomonas aeruginosa FabA. J. Mol. Biol. 2016, 428, 108–120. [Google Scholar] [CrossRef] [PubMed]
  124. McMurry, L.M.; Oethinger, M.; Levy, S.B. Triclosan targets lipid synthesis. Nature 1998, 394, 531–532. [Google Scholar] [CrossRef] [PubMed]
  125. Perozzo, R.; Kuo, M.; Sidhu, A.; Valiyaveettil, J.T.; Bittman, R.; Jacobs, W.R., Jr.; Fidock, D.A.; Sacchettini, J.C. Structural elucidation of the specificity of the antibacterial agent triclosan for malarial enoyl acyl carrier protein reductase. J. Biol. Chem. 2002, 277, 13106–13114. [Google Scholar] [CrossRef] [PubMed]
  126. Sivaraman, S.; Sullivan, T.J.; Johnson, F.; Novichenok, P.; Cui, G.; Simmerling, C.; Tonge, P.J. Inhibition of the bacterial enoyl reductase FabI by triclosan: A structure-reactivity analysis of FabI inhibition by triclosan analogues. J. Med. Chem. 2004, 47, 509–518. [Google Scholar] [CrossRef] [PubMed]
  127. Park, H.S.; Yoon, Y.M.; Jung, S.J.; Yun, I.N.; Kim, C.M.; Kim, J.M.; Kwak, J.H. CG400462, a new bacterial enoyl-acyl carrier protein reductase (FabI) inhibitor. Int. J. Antimicrob. Agents 2007, 30, 446–451. [Google Scholar] [CrossRef]
  128. Park, H.S.; Yoon, Y.M.; Jung, S.J.; Kim, C.M.; Kim, J.M.; Kwak, J.H. Antistaphylococcal activities of CG400549, a new bacterial enoyl-acyl carrier protein reductase (FabI) inhibitor. J. Antimicrob. Chemother. 2007, 60, 568–574. [Google Scholar] [CrossRef]
  129. Sampson, P.B.; Picard, C.; Handerson, S.; McGrath, T.E.; Domagala, M.; Leeson, A.; Romanov, V.; Awrey, D.E.; Thambipillai, D.; Bardouniotis, E.; et al. Spiro-naphthyridinone piperidines as inhibitors of S. aureus and E. coli enoyl-ACP reductase (FabI). Bioorg. Med. Chem. Lett. 2009, 19, 5355–5358. [Google Scholar] [CrossRef]
  130. Ramnauth, J.; Surman, M.D.; Sampson, P.B.; Forrest, B.; Wilson, J.; Freeman, E.; Manning, D.D.; Martin, F.; Toro, A.; Domagala, M.; et al. 2,3,4,5-Tetrahydro-1H-pyrido[2,3-b and e][1,4]diazepines as inhibitors of the bacterial enoyl ACP reductase, FabI. Bioorg. Med. Chem. Lett. 2009, 19, 5359–5362. [Google Scholar] [CrossRef]
  131. Escaich, S.; Prouvensier, L.; Saccomani, M.; Durant, L.; Oxoby, M.; Gerusz, V.; Moreau, F.; Vongsouthi, V.; Maher, K.; Morrissey, I.; et al. The MUT056399 inhibitor of FabI is a new antistaphylococcal compound. Antimicrob. Agents Chemother. 2011, 55, 4692–4697. [Google Scholar] [CrossRef]
  132. Banevicius, M.A.; Kaplan, N.; Hafkin, B.; Nicolau, D.P. Pharmacokinetics, pharmacodynamics and efficacy of novel FabI inhibitor AFN-1252 against MSSA and MRSA in the murine thigh infection model. J. Chemother. 2013, 25, 26–31. [Google Scholar] [CrossRef]
  133. Schiebel, J.; Chang, A.; Shah, S.; Lu, Y.; Liu, L.; Pan, P.; Hirschbeck, M.W.; Tareilus, M.; Eltschkner, S.; Yu, W.; et al. Rational design of broad spectrum antibacterial activity based on a clinically relevant enoyl-acyl carrier protein (ACP) reductase inhibitor. J. Biol. Chem. 2014, 289, 15987–16005. [Google Scholar] [CrossRef]
  134. Mandal, S.; Parish, T. A Novel Benzoxaborole Is Active against Escherichia coli and Binds to FabI. Antimicrob. Agents Chemother. 2021, 65, e0262220. [Google Scholar] [CrossRef] [PubMed]
  135. Yao, J.; Maxwell, J.B.; Rock, C.O. Resistance to AFN-1252 arises from missense mutations in Staphylococcus aureus enoyl-acyl carrier protein reductase (FabI). J. Biol. Chem. 2013, 288, 36261–36271. [Google Scholar] [CrossRef]
  136. Mehboob, S.; Song, J.; Hevener, K.E.; Su, P.C.; Boci, T.; Brubaker, L.; Truong, L.; Mistry, T.; Deng, J.; Cook, J.L.; et al. Structural and biological evaluation of a novel series of benzimidazole inhibitors of Francisella tularensis enoyl-ACP reductase (FabI). Bioorg. Med. Chem. Lett. 2015, 25, 1292–1296. [Google Scholar] [CrossRef] [PubMed]
  137. Takahata, S.; Iida, M.; Yoshida, T.; Kumura, K.; Kitagawa, H.; Hoshiko, S. Discovery of 4-Pyridone derivatives as specific inhibitors of enoyl-acyl carrier protein reductase (FabI) with antibacterial activity against Staphylococcus aureus. J. Antibiot. 2007, 60, 123–128. [Google Scholar] [CrossRef]
  138. Wang, S.F.; Yin, Y.; Wu, X.; Qiao, F.; Sha, S.; Lv, P.C.; Zhao, J.; Zhu, H.L. Synthesis, molecular docking and biological evaluation of coumarin derivatives containing piperazine skeleton as potential antibacterial agents. Bioorg. Med. Chem. 2014, 22, 5727–5737. [Google Scholar] [CrossRef]
  139. Hu, Y.; Shen, Y.; Wu, X.; Tu, X.; Wang, G.X. Synthesis and biological evaluation of coumarin derivatives containing imidazole skeleton as potential antibacterial agents. Eur. J. Med. Chem. 2018, 143, 958–969. [Google Scholar] [CrossRef] [PubMed]
  140. Kwon, J.; Mistry, T.; Ren, J.; Johnson, M.E.; Mehboob, S. A novel series of enoyl reductase inhibitors targeting the ESKAPE pathogens, Staphylococcus aureus and Acinetobacter baumannii. Bioorg. Med. Chem. 2018, 26, 65–76. [Google Scholar] [CrossRef]
  141. Davis, M.C.; Franzblau, S.G.; Martin, A.R. Syntheses and evaluation of benzodiazaborine compounds against M. tuberculosis H37Rv In vitro. Bioorg. Med. Chem. Lett. 1998, 8, 843–846. [Google Scholar] [CrossRef] [PubMed]
  142. Rodriguez, F.; Saffon, N.; Sammartino, J.C.; Degiacomi, G.; Pasca, M.R.; Lherbet, C. First triclosan-based macrocyclic inhibitors of InhA enzyme. Bioorg. Chem. 2020, 95, 103498. [Google Scholar] [CrossRef] [PubMed]
  143. Manjunatha, U.H.; Rao, S.P.S.; Kondreddi, R.R.; Noble, C.G.; Camacho, L.R.; Tan, B.H.; Ng, S.H.; Ng, P.S.; Ma, N.L.; Lakshminarayana, S.B.; et al. Direct inhibitors of InhA are active against Mycobacterium tuberculosis. Sci. Transl. Med. 2015, 7, 269ra3. [Google Scholar] [CrossRef] [PubMed]
  144. Shirude, P.S.; Madhavapeddi, P.; Naik, M.; Murugan, K.; Shinde, V.; Nandishaiah, R.; Bhat, J.; Kumar, A.; Hameed, S.; Holdgate, G.; et al. Methyl-thiazoles: A novel mode of inhibition with the potential to develop novel inhibitors targeting InhA in Mycobacterium tuberculosis. J. Med. Chem. 2013, 56, 8533–8542. [Google Scholar] [CrossRef]
  145. Šink, R.; Sosič, I.; Živec, M.; Fernandez-Menendez, R.; Turk, S.; Pajk, S.; Alvarez-Gomez, D.; Lopez-Roman, E.M.; Gonzales-Cortez, C.; Rullas-Triconado, J.; et al. Design, synthesis, and evaluation of new thiadiazole-based direct inhibitors of enoyl acyl carrier protein reductase (InhA) for the treatment of tuberculosis. J. Med. Chem. 2015, 58, 613–624. [Google Scholar] [CrossRef] [PubMed]
  146. Joshi, S.D.; Dixit, S.R.; Kulkarni, V.H.; Lherbet, C.; Nadagouda, M.N.; Aminabhavi, T.M. Synthesis, biological evaluation and in silico molecular modeling of pyrrolyl benzohydrazide derivatives as enoyl ACP reductase inhibitors. Eur. J. Med. Chem. 2017, 126, 286–297. [Google Scholar] [CrossRef]
  147. Rotta, M.; Pissinate, K.; Villela, A.D.; Back, D.F.; Timmers, L.F.; Bachega, J.F.; de Souza, O.N.; Santos, D.S.; Basso, L.A.; Machado, P. Piperazine derivatives: Synthesis, inhibition of the Mycobacterium tuberculosis enoyl-acyl carrier protein reductase and SAR studies. Eur. J. Med. Chem. 2015, 90, 436–447. [Google Scholar] [CrossRef]
  148. Chollet, A.; Mori, G.; Menendez, C.; Rodriguez, F.; Fabing, I.; Pasca, M.R.; Madacki, J.; Korduláková, J.; Constant, P.; Quémard, A.; et al. Design, synthesis and evaluation of new GEQ derivatives as inhibitors of InhA enzyme and Mycobacterium tuberculosis growth. Eur. J. Med. Chem. 2015, 101, 218–235. [Google Scholar] [CrossRef]
  149. Hartkoorn, R.C.; Sala, C.; Neres, J.; Pojer, F.; Magnet, S.; Mukherjee, R.; Uplekar, S.; Boy-Röttger, S.; Altmann, K.H.; Cole, S.T. Towards a new tuberculosis drug: Pyridomycin—Nature’s isoniazid. EMBO Mol. Med. 2012, 4, 1032–1042. [Google Scholar] [CrossRef]
  150. Karioti, A.; Skaltsa, H.; Zhang, X.; Tonge, P.J.; Perozzo, R.; Kaiser, M.; Franzblau, S.G.; Tasdemir, D. Inhibiting enoyl-ACP reductase (FabI) across pathogenic microorganisms by linear sesquiterpene lactones from Anthemis auriculata. Phytomedicine 2008, 15, 1125–1129. [Google Scholar] [CrossRef]
  151. Brenwald, N.P.; Fraise, A.P. Triclosan resistance in methicillin-resistant Staphylococcus aureus (MRSA). J. Hosp. Infect. 2003, 55, 141–144. [Google Scholar] [CrossRef] [PubMed]
  152. Yu, B.J.; Kim, J.A.; Pan, J.G. Signature gene expression profile of triclosan-resistant Escherichia coli. J. Antimicrob. Chemother. 2010, 65, 1171–1177. [Google Scholar] [CrossRef] [PubMed]
  153. Ciusa, M.L.; Furi, L.; Knight, D.; Decorosi, F.; Fondi, M.; Raggi, C.; Coelho, J.R.; Aragones, L.; Moce, L.; Visa, P.; et al. A novel resistance mechanism to triclosan that suggests horizontal gene transfer and demonstrates a potential selective pressure for reduced biocide susceptibility in clinical strains of Staphylococcus aureus. Int. J. Antimicrob. Agents 2012, 40, 210–220. [Google Scholar] [CrossRef] [PubMed]
  154. Chuanchuen, R.; Beinlich, K.; Hoang, T.T.; Becher, A.; Karkhoff-Schweizer, R.R.; Schweizer, H.P. Cross-resistance between triclosan and antibiotics in Pseudomonas aeruginosa is mediated by multidrug efflux pumps: Exposure of a susceptible mutant strain to triclosan selects nfxB mutants overexpressing MexCD-OprJ. Antimicrob. Agents Chemother. 2001, 45, 428–432. [Google Scholar] [CrossRef]
  155. Wang, L.; Mao, B.; He, H.; Shang, Y.; Zhong, Y.; Yu, Z.; Yang, Y.; Li, H.; An, J. Comparison of hepatotoxicity and mechanisms induced by triclosan (TCS) and methyl-triclosan (MTCS) in human liver hepatocellular HepG2 cells. Toxicol. Res. 2018, 8, 38–45. [Google Scholar] [CrossRef]
  156. Marrakchi, H.; Dewolf, W.E., Jr.; Quinn, C.; West, J.; Polizzi, B.J.; So, C.Y.; Holmes, D.J.; Reed, S.L.; Heath, R.J.; Payne, D.J.; et al. Characterization of Streptococcus pneumoniae enoyl-(acyl-carrier protein) reductase (FabK). Biochem. J. 2003, 370, 1055–1062. [Google Scholar] [CrossRef]
  157. Heath, R.J.; Rock, C.O. A triclosan-resistant bacterial enzyme. Nature 2000, 406, 145–146. [Google Scholar] [CrossRef]
  158. Heath, R.J.; Su, N.; Murphy, C.K.; Rock, C.O. The enoyl-[acyl-carrier-protein] reductases FabI and FabL from Bacillus subtilis. J. Biol. Chem. 2000, 275, 40128–40133. [Google Scholar] [CrossRef]
  159. Huang, Y.H.; Lin, J.S.; Ma, J.C.; Wang, H.H. Functional Characterization of Triclosan-Resistant Enoyl-acyl-carrier Protein Reductase (FabV) in Pseudomonas aeruginosa. Front. Microbiol. 2016, 7, 1903. [Google Scholar] [CrossRef]
  160. Kim, S.H.; Khan, R.; Choi, K.; Lee, S.W.; Rhee, S. A triclosan-resistance protein from the soil metagenome is a novel enoyl-acyl carrier protein reductase: Structure-guided functional analysis. FEBS J. 2020, 287, 4710–4728. [Google Scholar] [CrossRef]
  161. Seefeld, M.A.; Miller, W.H.; Newlander, K.A.; Burgess, W.J.; DeWolf, W.E., Jr.; Elkins, P.A.; Head, M.S.; Jakas, D.R.; Janson, C.A.; Keller, P.M.; et al. Indole naphthyridinones as inhibitors of bacterial enoyl-ACP reductases FabI and FabK. J. Med. Chem. 2003, 46, 1627–1635. [Google Scholar] [CrossRef]
  162. Takahata, S.; Iida, M.; Osaki, Y.; Saito, J.; Kitagawa, H.; Ozawa, T.; Yoshida, T.; Hoshiko, S. AG205, a novel agent directed against FabK of Streptococcus pneumoniae. Antimicrob. Agents Chemother. 2006, 50, 2869–2871. [Google Scholar] [CrossRef] [PubMed]
  163. Mahfuz, A.M.U.B.; Stambuk Opazo, F.; Aguilar, L.F.; Iqbal, M.N. Carfilzomib as a potential inhibitor of NADH-dependent enoyl-acyl carrier protein reductases of Klebsiella pneumoniae and Mycobacterium tuberculosis as a drug target enzyme: Insights from molecular docking and molecular dynamics. J. Biomol. Struct. Dyn. 2022, 40, 4021–4037. [Google Scholar] [CrossRef] [PubMed]
  164. Zhang, Y.M.; Rock, C.O. Thematic review series: Glycerolipids. Acyltransferases in bacterial glycerophospholipid synthesis. J. Lipid Res. 2008, 49, 1867–1874. [Google Scholar] [CrossRef] [PubMed]
  165. De Mendoza, D.; Klages Ulrich, A.; Cronan, J.E., Jr. Thermal regulation of membrane fluidity in Escherichia coli. Effects of overproduction of beta-ketoacyl-acyl carrier protein synthase I. J. Biol. Chem. 1983, 258, 2098–2101. [Google Scholar] [CrossRef]
  166. Cronan, J.E., Jr.; Weisberg, L.J.; Allen, R.G. Regulation of membrane lipid synthesis in Escherichia coli. Accumulation of free fatty acids of abnormal length during inhibition of phospholipid synthesis. J. Biol. Chem. 1975, 250, 5835–5840. [Google Scholar] [CrossRef]
  167. Bell, R.M. Mutants of Escherichia coli defective in membrane phospholipid synthesis. Properties of wild type and Km defective sn-glycerol-3-phosphate acyltransferase activities. J. Biol. Chem. 1975, 250, 7147–7152. [Google Scholar] [CrossRef]
  168. Lu, Y.J.; Zhang, Y.M.; Grimes, K.D.; Qi, J.; Lee, R.E.; Rock, C.O. Acyl-phosphates initiate membrane phospholipid synthesis in Gram-positive pathogens. Mol. Cell 2006, 23, 765–772. [Google Scholar] [CrossRef]
  169. Greenway, D.L.; Silbert, D.F. Altered acyltransferase activity in Escherichia coli associated with mutations in acyl coenzyme A synthetase. J. Biol. Chem. 1983, 258, 13034–13042. [Google Scholar] [CrossRef]
  170. Brinster, S.; Lamberet, G.; Staels, B.; Trieu-Cuot, P.; Gruss, A.; Poyart, C. Type II fatty acid synthesis is not a suitable antibiotic target for Gram-positive pathogens. Nature 2009, 458, 83–86. [Google Scholar] [CrossRef]
  171. Morvan, C.; Halpern, D.; Kénanian, G.; Hays, C.; Anba-Mondoloni, J.; Brinster, S.; Kennedy, S.; Trieu-Cuot, P.; Poyart, C.; Lamberet, G.; et al. Environmental fatty acids enable emergence of infectious Staphylococcus aureus resistant to FASII-targeted antimicrobials. Nat. Commun. 2016, 7, 12944. [Google Scholar] [CrossRef] [PubMed]
  172. Gloux, K.; Guillemet, M.; Soler, C.; Morvan, C.; Halpern, D.; Pourcel, C.; Vu Thien, H.; Lamberet, G.; Gruss, A. Clinical Relevance of Type II Fatty Acid Synthesis Bypass in Staphylococcus aureus. Antimicrob. Agents Chemother. 2017, 61, e02515-16. [Google Scholar] [CrossRef] [PubMed]
  173. Kim, E.J.; Lee, J.K. Effect of changes in the composition of cellular fatty acids on membrane fluidity of Rhodobacter sphaeroides. J. Microbiol. Biotechnol. 2015, 25, 162–173. [Google Scholar] [CrossRef] [PubMed]
  174. Grogan, D.W.; Cronan, J.E., Jr. Characterization of Escherichia coli mutants completely defective in synthesis of cyclopropane fatty acids. J. Bacteriol. 1986, 166, 872–877. [Google Scholar] [CrossRef]
  175. Harley, J.B.; Santangelo, G.M.; Rasmussen, H.; Goldfine, H. Dependence of Escherichia coli hyperbaric oxygen toxicity on the lipid acyl chain composition. J. Bacteriol. 1978, 134, 808–820. [Google Scholar] [CrossRef]
  176. Dufourc, E.J.; Smith, I.C.; Jarrell, H.C. A 2H-NMR analysis of dihydrosterculoyl-containing lipids in model membranes: Structural effects of a cyclopropane ring. Chem. Phys. Lipids 1983, 33, 153–177. [Google Scholar] [CrossRef] [PubMed]
  177. Choi, T.R.; Park, Y.L.; Song, H.S.; Lee, S.M.; Park, S.L.; Lee, H.S.; Kim, H.J.; Bhatia, S.K.; Gurav, R.; Lee, Y.K.; et al. Effects of a Δ-9-fatty acid desaturase and a cyclopropane-fatty acid synthase from the novel psychrophile Pseudomonas sp. B14-6 on bacterial membrane properties. J. Ind. Microbiol. Biotechnol. 2020, 47, 1045–1057. [Google Scholar] [CrossRef] [PubMed]
  178. Choi, T.R.; Song, H.S.; Han, Y.H.; Park, Y.L.; Park, J.Y.; Yang, S.Y.; Bhatia, S.K.; Gurav, R.; Kim, H.J.; Lee, Y.K.; et al. Enhanced tolerance to inhibitors of Escherichia coli by heterologous expression of cyclopropane-fatty acid-acyl-phospholipid synthase (cfa) from Halomonas socia. Bioprocess Biosyst. Eng. 2020, 43, 909–918. [Google Scholar] [CrossRef]
  179. Wang, A.Y.; Grogan, D.W.; Cronan, J.E., Jr. Cyclopropane fatty acid synthase of Escherichia coli: Deduced amino acid sequence, purification, and studies of the enzyme active site. Biochemistry 1992, 31, 11020–11028. [Google Scholar] [CrossRef]
  180. Barkan, D.; Liu, Z.; Sacchettini, J.C.; Glickman, M.S. Mycolic acid cyclopropanation is essential for viability, drug resistance, and cell wall integrity of Mycobacterium tuberculosis. Chem. Biol. 2009, 16, 499–509. [Google Scholar] [CrossRef]
  181. Barkan, D.; Hedhli, D.; Yan, H.G.; Huygen, K.; Glickman, M.S. Mycobacterium tuberculosis lacking all mycolic acid cyclopropanation is viable but highly attenuated and hyperinflammatory in mice. Infect. Immun. 2012, 80, 1958–1968. [Google Scholar] [CrossRef]
  182. Blunsom, N.J.; Cockcroft, S. CDP-Diacylglycerol Synthases (CDS): Gateway to Phosphatidylinositol and Cardiolipin Synthesis. Front. Cell Dev. Biol. 2020, 8, 63. [Google Scholar] [CrossRef]
  183. Jennings, W.; Epand, R.M. CDP-diacylglycerol, a critical intermediate in lipid metabolism. Chem. Phys. Lipids 2020, 230, 104914. [Google Scholar] [CrossRef] [PubMed]
  184. Lu, Y.H.; Guan, Z.; Zhao, J.; Raetz, C.R. Three phosphatidylglycerol-phosphate phosphatases in the inner membrane of Escherichia coli. J. Biol. Chem. 2011, 286, 5506–5518. [Google Scholar] [CrossRef] [PubMed]
  185. Tan, B.K.; Bogdanov, M.; Zhao, J.; Dowhan, W.; Raetz, C.R.; Guan, Z. Discovery of a cardiolipin synthase utilizing phosphatidylethanolamine and phosphatidylglycerol as substrates. Proc. Natl. Acad. Sci. USA 2012, 109, 16504–16509. [Google Scholar] [CrossRef] [PubMed]
  186. Guo, D.; Tropp, B.E. A second Escherichia coli protein with CL synthase activity. Biochim. Biophys. Acta 2000, 1483, 263–274. [Google Scholar] [CrossRef]
  187. Li, C.; Tan, B.K.; Zhao, J.; Guan, Z. In Vivo and In vitro Synthesis of Phosphatidylglycerol by an Escherichia coli Cardiolipin Synthase. J. Biol. Chem. 2016, 291, 25144–25153. [Google Scholar] [CrossRef]
  188. Tsai, M.; Ohniwa, R.L.; Kato, Y.; Takeshita, S.L.; Ohta, T.; Saito, S.; Hayashi, H.; Morikawa, K. Staphylococcus aureus requires cardiolipin for survival under conditions of high salinity. BMC Microbiol. 2011, 11, 13. [Google Scholar] [CrossRef]
  189. Kawai, F.; Shoda, M.; Harashima, R.; Sadaie, Y.; Hara, H.; Matsumoto, K. Cardiolipin domains in Bacillus subtilis marburg membranes. J. Bacteriol. 2004, 186, 1475–1483. [Google Scholar] [CrossRef] [PubMed]
  190. Sohlenkamp, C.; de Rudder, K.E.; Geiger, O. Phosphatidylethanolamine is not essential for growth of Sinorhizobium meliloti on complex culture media. J. Bacteriol. 2004, 186, 1667–1677. [Google Scholar] [CrossRef]
  191. Sohlenkamp, C.; López-Lara, I.M.; Geiger, O. Biosynthesis of phosphatidylcholine in bacteria. Prog. Lipid Res. 2003, 42, 115–162. [Google Scholar] [CrossRef] [PubMed]
  192. Jackson, M.; Crick, D.C.; Brennan, P.J. Phosphatidylinositol is an essential phospholipid of mycobacteria. J. Biol. Chem. 2000, 275, 30092–30099. [Google Scholar] [CrossRef] [PubMed]
  193. Salman, M.; Lonsdale, J.T.; Besra, G.S.; Brennan, P.J. Phosphatidylinositol synthesis in mycobacteria. Biochim. Biophys. Acta 1999, 1436, 437–450. [Google Scholar] [CrossRef] [PubMed]
  194. Morii, H.; Okauchi, T.; Nomiya, H.; Ogawa, M.; Fukuda, K.; Taniguchi, H. Studies of inositol 1-phosphate analogues as inhibitors of the phosphatidylinositol phosphate synthase in mycobacteria. J. Biochem. 2013, 153, 257–266. [Google Scholar] [CrossRef] [PubMed]
  195. Roy, H.; Dare, K.; Ibba, M. Adaptation of the bacterial membrane to changing environments using aminoacylated phospholipids. Mol. Microbiol. 2009, 71, 547–550. [Google Scholar] [CrossRef] [PubMed]
  196. Andrä, J.; Goldmann, T.; Ernst, C.M.; Peschel, A.; Gutsmann, T. Multiple peptide resistance factor (MprF)-mediated Resistance of Staphylococcus aureus against antimicrobial peptides coincides with a modulated peptide interaction with artificial membranes comprising lysyl-phosphatidylglycerol. J. Biol. Chem. 2011, 286, 18692–18700. [Google Scholar] [CrossRef] [PubMed]
  197. Ernst, C.M.; Peschel, A. Broad-spectrum antimicrobial peptide resistance by MprF-mediated aminoacylation and flipping of phospholipids. Mol. Microbiol. 2011, 80, 290–299. [Google Scholar] [CrossRef]
  198. Slavetinsky, C.J.; Hauser, J.N.; Gekeler, C.; Slavetinsky, J.; Geyer, A.; Kraus, A.; Heilingbrunner, D.; Wagner, S.; Tesar, M.; Krismer, B.; et al. Sensitizing Staphylococcus aureus to antibacterial agents by decoding and blocking the lipid flippase MprF. Elife 2022, 11, e66376. [Google Scholar] [CrossRef]
  199. Dhankhar, P.; Dalal, V.; Kotra, D.G.; Kumar, P. In-silico approach to identify novel potent inhibitors against GraR of S. aureus. Front. Biosci. Landmark Ed. 2020, 25, 1337–1360. [Google Scholar] [CrossRef]
  200. Van Horn, W.D.; Sanders, C.R. Prokaryotic diacylglycerol kinase and undecaprenol kinase. Annu. Rev. Biophys. 2012, 41, 81–101. [Google Scholar] [CrossRef]
  201. Baker, B.R.; Ives, C.M.; Bray, A.; Caffrey, M.; Cochrane, S.A. Undecaprenol kinase: Function, mechanism and substrate specificity of a potential antibiotic target. Eur. J. Med. Chem. 2021, 210, 113062. [Google Scholar] [CrossRef]
  202. Yeo, W.S.; Jeong, B.; Ullah, N.; Shah, M.A.; Ali, A.; Kim, K.K.; Bae, T. Ftsh Sensitizes Methicillin-Resistant Staphylococcus aureus to β-Lactam Antibiotics by Degrading YpfP, a Lipoteichoic Acid Synthesis Enzyme. Antibiotics 2021, 10, 1198. [Google Scholar] [CrossRef] [PubMed]
  203. Galloway, S.M.; Raetz, C.R. A mutant of Escherichia coli defective in the first step of endotoxin biosynthesis. J. Biol. Chem. 1990, 265, 6394–6402. [Google Scholar] [CrossRef]
  204. Kelly, T.M.; Stachula, S.A.; Raetz, C.R.; Anderson, M.S. The firA gene of Escherichia coli encodes UDP-3-O-(R-3-hydroxymyristoyl)-glucosamine N-acyltransferase. The third step of endotoxin biosynthesis. J. Biol. Chem. 1993, 268, 19866–19874. [Google Scholar] [CrossRef]
  205. Raetz, C.R.; Whitfield, C. Lipopolysaccharide endotoxins. Annu. Rev. Biochem. 2002, 71, 635–700. [Google Scholar] [CrossRef] [PubMed]
  206. Williams, A.H.; Immormino, R.M.; Gewirth, D.T.; Raetz, C.R. Structure of UDP-N-acetylglucosamine acyltransferase with a bound antibacterial pentadecapeptide. Proc. Natl. Acad. Sci. USA 2006, 103, 10877–10882. [Google Scholar] [CrossRef]
  207. Jenkins, R.J.; Dotson, G.D. Dual targeting antibacterial peptide inhibitor of early lipid A biosynthesis. ACS Chem. Biol. 2012, 7, 1170–1177. [Google Scholar] [CrossRef] [PubMed]
  208. Han, W.; Ma, X.; Balibar, C.J.; Baxter Rath, C.M.; Benton, B.; Bermingham, A.; Casey, F.; Chie-Leon, B.; Cho, M.K.; Frank, A.O.; et al. Two Distinct Mechanisms of Inhibition of LpxA Acyltransferase Essential for Lipopolysaccharide Biosynthesis. J. Am. Chem. Soc. 2020, 142, 4445–4455. [Google Scholar] [CrossRef]
  209. Clements, J.M.; Coignard, F.; Johnson, I.; Chandler, S.; Palan, S.; Waller, A.; Wijkmans, J.; Hunter, M.G. Antibacterial activities and characterization of novel inhibitors of LpxC. Antimicrob. Agents Chemother. 2002, 46, 1793–1799. [Google Scholar] [CrossRef]
  210. Mdluli, K.E.; Witte, P.R.; Kline, T.; Barb, A.W.; Erwin, A.L.; Mansfield, B.E.; McClerren, A.L.; Pirrung, M.C.; Tumey, L.N.; Warrener, P.; et al. Molecular validation of LpxC as an antibacterial drug target in Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 2006, 50, 2178–2184. [Google Scholar] [CrossRef]
  211. Chen, M.H.; Steiner, M.G.; de Laszlo, S.E.; Patchett, A.A.; Anderson, M.S.; Hyland, S.A.; Onishi, H.R.; Silver, L.L.; Raetz, C.R. Carbohydroxamido-oxazolidines: Antibacterial agents that target lipid A biosynthesis. Bioorg. Med. Chem. Lett. 1999, 9, 313–318. [Google Scholar] [CrossRef]
  212. McClerren, A.L.; Endsley, S.; Bowman, J.L.; Andersen, N.H.; Guan, Z.; Rudolph, J.; Raetz, C.R. A slow, tight-binding inhibitor of the zinc-dependent deacetylase LpxC of lipid A biosynthesis with antibiotic activity comparable to ciprofloxacin. Biochemistry 2005, 44, 16574–16583. [Google Scholar] [CrossRef]
  213. Barb, A.W.; McClerren, A.L.; Snehelatha, K.; Reynolds, C.M.; Zhou, P.; Raetz, C.R. Inhibition of lipid A biosynthesis as the primary mechanism of CHIR-090 antibiotic activity in Escherichia coli. Biochemistry 2007, 46, 3793–3802. [Google Scholar] [CrossRef] [PubMed]
  214. Tomaras, A.P.; McPherson, C.J.; Kuhn, M.; Carifa, A.; Mullins, L.; George, D.; Desbonnet, C.; Eidem, T.M.; Montgomery, J.I.; Brown, M.F.; et al. LpxC inhibitors as new antibacterial agents and tools for studying regulation of lipid A biosynthesis in Gram-negative pathogens. mBio 2014, 5, e01551-14. [Google Scholar] [CrossRef] [PubMed]
  215. Jackman, J.E.; Fierke, C.A.; Tumey, L.N.; Pirrung, M.; Uchiyama, T.; Tahir, S.H.; Hindsgaul, O.; Raetz, C.R. Antibacterial agents that target lipid A biosynthesis in gram-negative bacteria. Inhibition of diverse UDP-3-O-(r-3-hydroxymyristoyl)-n-acetylglucosamine deacetylases by substrate analogs containing zinc binding motifs. J. Biol. Chem. 2000, 275, 11002–11009. [Google Scholar] [CrossRef] [PubMed]
  216. Ma, X.; Prathapam, R.; Wartchow, C.; Chie-Leon, B.; Ho, C.M.; De Vicente, J.; Han, W.; Li, M.; Lu, Y.; Ramurthy, S.; et al. Structural and Biological Basis of Small Molecule Inhibition of Escherichia coli LpxD Acyltransferase Essential for Lipopolysaccharide Biosynthesis. ACS Infect. Dis. 2020, 6, 1480–1489. [Google Scholar] [CrossRef] [PubMed]
  217. Bohl, H.O.; Ieong, P.; Lee, J.K.; Lee, T.; Kankanala, J.; Shi, K.; Demir, Ö.; Kurahashi, K.; Amaro, R.E.; Wang, Z.; et al. The substrate-binding cap of the UDP-diacylglucosamine pyrophosphatase LpxH is highly flexible, enabling facile substrate binding and product release. J. Biol. Chem. 2018, 293, 7969–7981. [Google Scholar] [CrossRef]
  218. Cho, J.; Lee, M.; Cochrane, C.S.; Webster, C.G.; Fenton, B.A.; Zhao, J.; Hong, J.; Zhou, P. Structural basis of the UDP-diacylglucosamine pyrophosphohydrolase LpxH inhibition by sulfonyl piperazine antibiotics. Proc. Natl. Acad. Sci. USA 2020, 117, 4109–4116. [Google Scholar] [CrossRef] [PubMed]
  219. Lee, M.; Zhao, J.; Kwak, S.H.; Cho, J.; Lee, M.; Gillespie, R.A.; Kwon, D.Y.; Lee, H.; Park, H.J.; Wu, Q.; et al. Structure-Activity Relationship of Sulfonyl Piperazine LpxH Inhibitors Analyzed by an LpxE-Coupled Malachite Green Assay. ACS Infect. Dis. 2019, 5, 641–651. [Google Scholar] [CrossRef]
  220. Raetz, C.R.; Reynolds, C.M.; Trent, M.S.; Bishop, R.E. Lipid A modification systems in gram-negative bacteria. Annu. Rev. Biochem. 2007, 76, 295–329. [Google Scholar] [CrossRef]
  221. Williams, A.H.; Raetz, C.R. Structural basis for the acyl chain selectivity and mechanism of UDP-N-acetylglucosamine acyltransferase. Proc. Natl. Acad. Sci. USA 2007, 104, 13543–13550. [Google Scholar] [CrossRef] [PubMed]
  222. Jenkins, R.J.; Heslip, K.A.; Meagher, J.L.; Stuckey, J.A.; Dotson, G.D. Structural basis for the recognition of peptide RJPXD33 by acyltransferases in lipid A biosynthesis. J. Biol. Chem. 2014, 289, 15527–15535. [Google Scholar] [CrossRef] [PubMed]
  223. Dangkulwanich, M.; Raetz, C.R.H.; Williams, A.H. Structure guided design of an antibacterial peptide that targets UDP-N-acetylglucosamine acyltransferase. Sci. Rep. 2019, 9, 3947. [Google Scholar] [CrossRef] [PubMed]
  224. Kroeck, K.G.; Sacco, M.D.; Smith, E.W.; Zhang, X.; Shoun, D.; Akhtar, A.; Darch, S.E.; Cohen, F.; Andrews, L.D.; Knox, J.E.; et al. Discovery of dual-activity small-molecule ligands of Pseudomonas aeruginosa LpxA and LpxD using SPR and X-ray crystallography. Sci. Rep. 2019, 9, 15450. [Google Scholar] [CrossRef]
  225. Bhaskar, B.V.; Babu, T.M.C.; Rammohan, A.; Zheng, G.Y.; Zyryanov, G.V.; Gu, W. Structure-Based Virtual Screening of Pseudomonas aeruginosa LpxA Inhibitors Using Pharmacophore-Based Approach. Biomolecules 2020, 10, 266. [Google Scholar] [CrossRef]
  226. Pratap, S.; Kesari, P.; Yadav, R.; Dev, A.; Narwal, M.; Kumar, P. Acyl chain preference and inhibitor identification of Moraxella catarrhalis LpxA: Insight through crystal structure and computational studies. Int. J. Biol. Macromol. 2017, 96, 759–765. [Google Scholar] [CrossRef]
  227. Shapiro, A.B.; Ross, P.L.; Gao, N.; Livchak, S.; Kern, G.; Yang, W.; Andrews, B.; Thresher, J. A high-throughput-compatible fluorescence anisotropy-based assay for competitive inhibitors of Escherichia coli UDP-N-acetylglucosamine acyltransferase (LpxA). J. Biomol. Screen. 2013, 18, 341–347. [Google Scholar] [CrossRef]
  228. Mochalkin, I.; Knafels, J.D.; Lightle, S. Crystal structure of LpxC from Pseudomonas aeruginosa complexed with the potent BB-78485 inhibitor. Protein Sci. 2008, 17, 450–457. [Google Scholar] [CrossRef]
  229. Onishi, H.R.; Pelak, B.A.; Gerckens, L.S.; Silver, L.L.; Kahan, F.M.; Chen, M.H.; Patchett, A.A.; Galloway, S.M.; Hyland, S.A.; Anderson, M.S.; et al. Antibacterial agents that inhibit lipid A biosynthesis. Science 1996, 274, 980–982. [Google Scholar] [CrossRef]
  230. Coggins, B.E.; McClerren, A.L.; Jiang, L.; Li, X.; Rudolph, J.; Hindsgaul, O.; Raetz, C.R.; Zhou, P. Refined solution structure of the LpxC-TU-514 complex and pKa analysis of an active site histidine: Insights into the mechanism and inhibitor design. Biochemistry 2005, 44, 1114–1126. [Google Scholar] [CrossRef]
  231. Caughlan, R.E.; Jones, A.K.; Delucia, A.M.; Woods, A.L.; Xie, L.; Ma, B.; Barnes, S.W.; Walker, J.R.; Sprague, E.R.; Yang, X.; et al. Mechanisms decreasing In vitro susceptibility to the LpxC inhibitor CHIR-090 in the gram-negative pathogen Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 2012, 56, 17–27. [Google Scholar] [CrossRef] [PubMed]
  232. Lee, C.J.; Liang, X.; Chen, X.; Zeng, D.; Joo, S.H.; Chung, H.S.; Barb, A.W.; Swanson, S.M.; Nicholas, R.A.; Li, Y.; et al. Species-specific and inhibitor-dependent conformations of LpxC: Implications for antibiotic design. Chem. Biol. 2011, 18, 38–47. [Google Scholar] [CrossRef] [PubMed]
  233. Liang, X.; Lee, C.J.; Chen, X.; Chung, H.S.; Zeng, D.; Raetz, C.R.; Li, Y.; Zhou, P.; Toone, E.J. Syntheses, structures and antibiotic activities of LpxC inhibitors based on the diacetylene scaffold. Bioorg. Med. Chem. 2011, 19, 852–860. [Google Scholar] [CrossRef] [PubMed]
  234. Zeng, D.; Zhao, J.; Chung, H.S.; Guan, Z.; Raetz, C.R.; Zhou, P. Mutants resistant to LpxC inhibitors by rebalancing cellular homeostasis. J. Biol. Chem. 2013, 288, 5475–5486. [Google Scholar] [CrossRef]
  235. Jones, A.K.; Caughlan, R.E.; Woods, A.L.; Uehara, K.; Xie, L.; Barnes, S.W.; Walker, J.R.; Thompson, K.V.; Ranjitkar, S.; Lee, P.S.; et al. Mutations Reducing In vitro Susceptibility to Novel LpxC Inhibitors in Pseudomonas aeruginosa and Interplay of Efflux and Nonefflux Mechanisms. Antimicrob. Agents Chemother. 2019, 64, e01490-19. [Google Scholar] [CrossRef] [PubMed]
  236. Niu, Z.; Lei, P.; Wang, Y.; Wang, J.; Yang, J.; Zhang, J. Small molecule LpxC inhibitors against gram-negative bacteria: Advances and future perspectives. Eur. J. Med. Chem. 2023, 253, 115326. [Google Scholar] [CrossRef] [PubMed]
  237. Rath, S.N.; Ray, M.; Pattnaik, A.; Pradhan, S.K. Drug Target Identification and Elucidation of Natural Inhibitors for Bordetella petrii: An In Silico Study. Genom. Inform. 2016, 14, 241–254. [Google Scholar] [CrossRef]
  238. Metzger, L.E., 4th; Lee, J.K.; Finer-Moore, J.S.; Raetz, C.R.; Stroud, R.M. LpxI structures reveal how a lipid A precursor is synthesized. Nat. Struct. Mol. Biol. 2012, 19, 1132–1138. [Google Scholar] [CrossRef]
  239. Metzger, L.E., 4th; Raetz, C.R. An alternative route for UDP-diacylglucosamine hydrolysis in bacterial lipid A biosynthesis. Biochemistry 2010, 49, 6715–6726. [Google Scholar] [CrossRef]
  240. Young, H.E.; Zhao, J.; Barker, J.R.; Guan, Z.; Valdivia, R.H.; Zhou, P. Discovery of the Elusive UDP-Diacylglucosamine Hydrolase in the Lipid A Biosynthetic Pathway in Chlamydia trachomatis. mBio 2016, 7, e00090. [Google Scholar] [CrossRef]
  241. Nayar, A.S.; Dougherty, T.J.; Ferguson, K.E.; Granger, B.A.; McWilliams, L.; Stacey, C.; Leach, L.J.; Narita, S.; Tokuda, H.; Miller, A.A.; et al. Novel antibacterial targets and compounds revealed by a high-throughput cell wall reporter assay. J. Bacteriol. 2015, 197, 1726–1734. [Google Scholar] [CrossRef]
  242. Kwak, S.H.; Cochrane, C.S.; Ennis, A.F.; Lim, W.Y.; Webster, C.G.; Cho, J.; Fenton, B.A.; Zhou, P.; Hong, J. Synthesis and evaluation of sulfonyl piperazine LpxH inhibitors. Bioorg. Chem. 2020, 102, 104055. [Google Scholar] [CrossRef]
  243. Zhou, P.; Hong, J. Structure- and Ligand-Dynamics-Based Design of Novel Antibiotics Targeting Lipid A Enzymes LpxC and LpxH in Gram-Negative Bacteria. Acc. Chem. Res. 2021, 54, 1623–1634. [Google Scholar] [CrossRef]
  244. Martínez-Guitián, M.; Vázquez-Ucha, J.C.; Álvarez-Fraga, L.; Conde-Pérez, K.; Bou, G.; Poza, M.; Beceiro, A. Antisense inhibition of lpxB gene expression in Acinetobacter baumannii by peptide-PNA conjugates and synergy with colistin. J. Antimicrob. Chemother. 2020, 75, 51–59. [Google Scholar] [CrossRef] [PubMed]
  245. Damale, M.G.; Pathan, S.K.; Patil, R.B.; Sangshetti, J.N. Pharmacoinformatics approaches to identify potential hits against tetraacyldisaccharide 4′-kinase (LpxK) of Pseudomonas aeruginosa. RSC Adv. 2020, 10, 32856–32874. [Google Scholar] [CrossRef] [PubMed]
  246. Wang, X.; Quinn, P.J.; Yan, A. Kdo2 -lipid A: Structural diversity and impact on immunopharmacology. Biol. Rev. Camb. Philos. Soc. 2015, 90, 408–427. [Google Scholar] [CrossRef]
  247. Emiola, A.; George, J.; Andrews, S.S. A Complete Pathway Model for Lipid A Biosynthesis in Escherichia coli. PLoS ONE 2015, 10, e0121216. [Google Scholar] [CrossRef] [PubMed]
  248. Hankins, J.V.; Madsen, J.A.; Giles, D.K.; Childers, B.M.; Klose, K.E.; Brodbelt, J.S.; Trent, M.S. Elucidation of a novel Vibrio cholerae lipid A secondary hydroxy-acyltransferase and its role in innate immune recognition. Mol. Microbiol. 2011, 81, 1313–1329. [Google Scholar] [CrossRef] [PubMed]
  249. Shai, Y.; Makovitzky, A.; Avrahami, D. Host defense peptides and lipopeptides: Modes of action and potential candidates for the treatment of bacterial and fungal infections. Curr. Protein Pept. Sci. 2006, 7, 479–486. [Google Scholar] [CrossRef] [PubMed]
  250. Radek, K.; Gallo, R. Antimicrobial peptides: Natural effectors of the innate immune system. Semin. Immunopathol. 2007, 29, 27–43. [Google Scholar] [CrossRef]
  251. Hancock, R.E.; Sahl, H.G. Antimicrobial and host-defense peptides as new anti-infective therapeutic strategies. Nat. Biotechnol. 2006, 24, 1551–1557. [Google Scholar] [CrossRef] [PubMed]
  252. Wang, J.; Dou, X.; Song, J.; Lyu, Y.; Zhu, X.; Xu, L.; Li, W.; Shan, A. Antimicrobial peptides: Promising alternatives in the post feeding antibiotic era. Med. Res. Rev. 2019, 39, 831–859. [Google Scholar] [CrossRef]
  253. Ahmed, T.A.E.; Hammami, R. Recent insights into structure-function relationships of antimicrobial peptides. J. Food Biochem. 2019, 43, e12546. [Google Scholar] [CrossRef] [PubMed]
  254. Antonov, V.F.; Petrov, V.V.; Molnar, A.A.; Predvoditelev, D.A.; Ivanov, A.S. The appearance of single-ion channels in unmodified lipid bilayer membranes at the phase transition temperature. Nature 1980, 283, 585–586. [Google Scholar] [CrossRef]
  255. Finkelstein, A.; Andersen, O.S. The gramicidin A channel: A review of its permeability characteristics with special reference to the single-file aspect of transport. J. Membr. Biol. 1981, 59, 155–171. [Google Scholar] [CrossRef] [PubMed]
  256. Andersen, O.S.; Koeppe, R.E., 2nd. Molecular determinants of channel function. Physiol. Rev. 1992, 72 (Suppl. S4), S89–S158. [Google Scholar] [CrossRef]
  257. Fringeli, U.P.; Fringeli, M. Pore formation in lipid membranes by alamethicin. Proc. Natl. Acad. Sci. USA 1979, 76, 3852–3856. [Google Scholar] [CrossRef]
  258. Shai, Y.; Bach, D.; Yanovsky, A. Channel formation properties of synthetic pardaxin and analogues. J. Biol. Chem. 1990, 265, 20202–20209. [Google Scholar] [CrossRef]
  259. Capone, R.; Mustata, M.; Jang, H.; Arce, F.T.; Nussinov, R.; Lal, R. Antimicrobial protegrin-1 forms ion channels: Molecular dynamic simulation, atomic force microscopy, and electrical conductance studies. Biophys. J. 2010, 98, 2644–2652. [Google Scholar] [CrossRef]
  260. Watanabe, H.; Kawano, R. Channel Current Analysis for Pore-forming Properties of an Antimicrobial Peptide, Magainin 1, Using the Droplet Contact Method. Anal. Sci. 2016, 32, 57–60. [Google Scholar] [CrossRef]
  261. Gallucci, E.; Meleleo, D.; Micelli, S.; Picciarelli, V. Magainin 2 channel formation in planar lipid membranes: The role of lipid polar groups and ergosterol. Eur. Biophys. J. 2003, 32, 22–32. [Google Scholar] [CrossRef]
  262. Mellor, I.R.; Sansom, M.S. Ion-channel properties of mastoparan, a 14-residue peptide from wasp venom, and of MP3, a 12-residue analogue. Proc. R. Soc. Lond. B Biol. Sci. 1990, 239, 383–400. [Google Scholar] [CrossRef]
  263. Arbuzova, A.; Schwarz, G. Pore-forming action of mastoparan peptides on liposomes: A quantitative analysis. Biochim. Biophys. Acta 1999, 1420, 139–152. [Google Scholar] [CrossRef]
  264. Efimova, S.S.; Schagina, L.V.; Ostroumova, O.S. Channel-forming activity of cecropins in lipid bilayers: Effect of agents modifying the membrane dipole potential. Langmuir 2014, 30, 7884–7892. [Google Scholar] [CrossRef] [PubMed]
  265. Efimova, S.S.; Shekunov, E.V.; Chernyshova, D.N.; Zakharova, A.A.; Ostroumova, O.S. Dependence of the channel-forming ability of lantibiotics on the lipid composition of the membranes. Biochem. Suppl. Ser. A Membr. Cell Biol. 2022, 16, 144–150. [Google Scholar] [CrossRef]
  266. Sheth, T.R.; Henderson, R.M.; Hladky, S.B.; Cuthbert, A.W. Ion channel formation by duramycin. Biochim. Biophys. Acta 1992, 1107, 179–185. [Google Scholar] [CrossRef] [PubMed]
  267. Sansom, M.S. Alamethicin and related peptaibols—Model ion channels. Eur. Biophys. J. 1993, 22, 105–124. [Google Scholar] [CrossRef]
  268. Gordon, L.G.; Haydon, D.A. The unit conductance channel of alamethicin. Biochim. Biophys. Acta 1972, 255, 1014–1018. [Google Scholar] [CrossRef] [PubMed]
  269. Christensen, B.; Fink, J.; Merrifield, R.B.; Mauzerall, D. Channel-forming properties of cecropins and related model compounds incorporated into planar lipid membranes. Proc. Natl. Acad. Sci. USA 1988, 85, 5072–5076. [Google Scholar] [CrossRef] [PubMed]
  270. Sokolov, Y.; Mirzabekov, T.; Martin, D.W.; Lehrer, R.I.; Kagan, B.L. Membrane channel formation by antimicrobial protegrins. Biochim. Biophys. Acta 1999, 1420, 23–29. [Google Scholar] [CrossRef]
  271. Saint, N.; Marri, L.; Marchini, D.; Molle, G. The antibacterial peptide ceratotoxin A displays alamethicin-like behavior in lipid bilayers. Peptides 2003, 24, 1779–1784. [Google Scholar] [CrossRef]
  272. Mayer, S.F.; Ducrey, J.; Dupasquier, J.; Haeni, L.; Rothen-Rutishauser, B.; Yang, J.; Fennouri, A.; Mayer, M. Targeting specific membranes with an azide derivative of the pore-forming peptide ceratotoxin A. Biochim. Biophys. Acta Biomembr. 2019, 1861, 183023. [Google Scholar] [CrossRef] [PubMed]
  273. Wang, G. Human antimicrobial peptides and proteins. Pharmaceuticals 2014, 7, 545–594. [Google Scholar] [CrossRef]
  274. Henzler Wildman, K.A.; Lee, D.K.; Ramamoorthy, A. Mechanism of lipid bilayer disruption by the human antimicrobial peptide, LL-37. Biochemistry 2003, 42, 6545–6558. [Google Scholar] [CrossRef]
  275. Khondker, A.; Rheinstädter, M.C. How do bacterial membranes resist polymyxin antibiotics? Commun. Biol. 2020, 3, 77. [Google Scholar] [CrossRef] [PubMed]
  276. Sabnis, A.; Hagart, K.L.; Klöckner, A.; Becce, M.; Evans, L.E.; Furniss, R.C.D.; Mavridou, D.A.; Murphy, R.; Stevens, M.M.; Davies, J.C.; et al. Colistin kills bacteria by targeting lipopolysaccharide in the cytoplasmic membrane. Elife 2021, 10, e65836. [Google Scholar] [CrossRef]
  277. Taylor, R.; Beriashvili, D.; Taylor, S.; Palmer, M. Daptomycin Pore Formation Is Restricted by Lipid Acyl Chain Composition. ACS Infect. Dis. 2017, 3, 797–801. [Google Scholar] [CrossRef] [PubMed]
  278. Beriashvili, D.; Taylor, R.; Kralt, B.; Abu Mazen, N.; Taylor, S.D.; Palmer, M. Mechanistic studies on the effect of membrane lipid acyl chain composition on daptomycin pore formation. Chem. Phys. Lipids 2018, 216, 73–79. [Google Scholar] [CrossRef] [PubMed]
  279. Tyurin, A.P.; Alferova, V.A.; Paramonov, A.S.; Shuvalov, M.V.; Kudryakova, G.K.; Rogozhin, E.A.; Zherebker, A.Y.; Brylev, V.A.; Chistov, A.A.; Baranova, A.A.; et al. Gausemycins-A,B: Cyclic Lipoglycopeptides from Streptomyces sp. Angew. Chem. Int. Ed. Engl. 2021, 60, 18694–18703. [Google Scholar] [CrossRef]
  280. Brogden, K.A. Antimicrobial peptides: Pore formers or metabolic inhibitors in bacteria? Nat. Rev. Microbiol. 2005, 3, 238–250. [Google Scholar] [CrossRef]
  281. Laver, D.R. The barrel-stave model as applied to alamethicin and its analogs reevaluated. Biophys. J. 1994, 66, 355–359. [Google Scholar] [CrossRef] [PubMed]
  282. Porcelli, F.; Buck, B.; Lee, D.K.; Hallock, K.J.; Ramamoorthy, A.; Veglia, G. Structure and orientation of pardaxin determined by NMR experiments in model membranes. J. Biol. Chem. 2004, 279, 45815–45823. [Google Scholar] [CrossRef]
  283. Allende, D.; Simon, S.A.; McIntosh, T.J. Melittin-induced bilayer leakage depends on lipid material properties: Evidence for toroidal pores. Biophys. J. 2005, 88, 1828–1837. [Google Scholar] [CrossRef]
  284. Matsuzaki, K. Magainins as paradigm for the mode of action of pore forming polypeptides. Biochim. Biophys. Acta 1998, 1376, 391–400. [Google Scholar] [CrossRef] [PubMed]
  285. Murzyn, K.; Pasenkiewicz-Gierula, M. Construction of a toroidal model for the magainin pore. J. Mol. Model. 2003, 9, 217–224. [Google Scholar] [CrossRef]
  286. Gazit, E.; Boman, A.; Boman, H.G.; Shai, Y. Interaction of the mammalian antibacterial peptide cecropin P1 with phospholipid vesicles. Biochemistry 1995, 34, 11479–11488. [Google Scholar] [CrossRef]
  287. Fernandez, D.I.; Le Brun, A.P.; Whitwell, T.C.; Sani, M.A.; James, M.; Separovic, F. The antimicrobial peptide aurein 1.2 disrupts model membranes via the carpet mechanism. Phys. Chem. Chem. Phys. 2012, 14, 15739–15751. [Google Scholar] [CrossRef]
  288. Battista, F.; Oliva, R.; Del Vecchio, P.; Winter, R.; Petraccone, L. Insights into the Action Mechanism of the Antimicrobial Peptide Lasioglossin III. Int. J. Mol. Sci. 2021, 22, 2857. [Google Scholar] [CrossRef]
  289. Ma, B.; Fang, C.; Lu, L.; Wang, M.; Xue, X.; Zhou, Y.; Li, M.; Hu, Y.; Luo, X.; Hou, Z. The antimicrobial peptide thanatin disrupts the bacterial outer membrane and inactivates the NDM-1 metallo-β-lactamase. Nat. Commun. 2019, 10, 3517. [Google Scholar] [CrossRef]
  290. Henderson, J.M.; Waring, A.J.; Separovic, F.; Lee, K.Y.C. Antimicrobial Peptides Share a Common Interaction Driven by Membrane Line Tension Reduction. Biophys. J. 2016, 111, 2176–2189. [Google Scholar] [CrossRef]
  291. Efimova, S.S.; Medvedev, R.Y.; Chulkov, E.G.; Ostroumova, O.S. Regulation of the Pore-Forming Activity of Cecropin A by Local Anesthetics. Cell Tiss. Biol. 2018, 12, 331–341. [Google Scholar] [CrossRef]
  292. Lee, C.C.; Sun, Y.; Qian, S.; Huang, H.W. Transmembrane pores formed by human antimicrobial peptide LL-37. Biophys. J. 2011, 100, 1688–1696. [Google Scholar] [CrossRef]
  293. Mak, D.O.; Webb, W.W. Two classes of alamethicin transmembrane channels: Molecular models from single-channel properties. Biophys. J. 1995, 69, 2323–2336. [Google Scholar] [CrossRef]
  294. Fennouri, A.; Mayer, S.F.; Schroeder, T.B.H.; Mayer, M. Single channel planar lipid bilayer recordings of the melittin variant MelP5. Biochim. Biophys. Acta Biomembr. 2017, 1859, 2051–2057. [Google Scholar] [CrossRef]
  295. Wiedemann, I.; Benz, R.; Sahl, H.G. Lipid II-mediated pore formation by the peptide antibiotic nisin: A black lipid membrane study. J. Bacteriol. 2004, 186, 3259–3261. [Google Scholar] [CrossRef]
  296. Kagan, B.L.; Selsted, M.E.; Ganz, T.; Lehrer, R.I. Antimicrobial defensin peptides form voltage-dependent ion-permeable channels in planar lipid bilayer membranes. Proc. Natl. Acad. Sci. USA 1990, 87, 210–214. [Google Scholar] [CrossRef]
  297. Hristova, K.; Selsted, M.E.; White, S.H. Critical role of lipid composition in membrane permeabilization by rabbit neutrophil defensins. J. Biol. Chem. 1997, 272, 24224–24233. [Google Scholar] [CrossRef]
  298. Seydlová, G.; Sokol, A.; Lišková, P.; Konopásek, I.; Fišer, R. Daptomycin Pore Formation and Stoichiometry Depend on Membrane Potential of Target Membrane. Antimicrob. Agents Chemother. 2018, 63, e01589-18. [Google Scholar] [CrossRef]
  299. Peschel, A.; Jack, R.W.; Otto, M.; Collins, L.V.; Staubitz, P.; Nicholson, G.; Kalbacher, H.; Nieuwenhuizen, W.F.; Jung, G.; Tarkowski, A.; et al. Staphylococcus aureus resistance to human defensins and evasion of neutrophil killing via the novel virulence factor MprF is based on modification of membrane lipids with l-lysine. J. Exp. Med. 2001, 193, 1067–1076. [Google Scholar] [CrossRef]
  300. Dorrer, E.; Teuber, M. Induction of polymyxin resistance in Pseudomonas fluorescens by phosphate limitation. Arch. Microbiol. 1977, 114, 87–89. [Google Scholar] [CrossRef]
  301. Breazeale, S.D.; Ribeiro, A.A.; McClerren, A.L.; Raetz, C.R. A formyltransferase required for polymyxin resistance in Escherichia coli and the modification of lipid A with 4-Amino-4-deoxy-L-arabinose. Identification and function oF UDP-4-deoxy-4-formamido-L-arabinose. J. Biol. Chem. 2005, 280, 14154–14167. [Google Scholar] [CrossRef]
  302. Ernst, R.K.; Guina, T.; Miller, S.I. Salmonella typhimurium outer membrane remodeling: Role in resistance to host innate immunity. Microbes Infect. 2001, 3, 1327–1334. [Google Scholar] [CrossRef]
  303. Gunn, J.S.; Lim, K.B.; Krueger, J.; Kim, K.; Guo, L.; Hackett, M.; Miller, S.I. PmrA-PmrB-regulated genes necessary for 4-aminoarabinose lipid A modification and polymyxin resistance. Mol. Microbiol. 1998, 27, 1171–1182. [Google Scholar] [CrossRef] [PubMed]
  304. Moskowitz, S.M.; Ernst, R.K.; Miller, S.I. PmrAB, a two-component regulatory system of Pseudomonas aeruginosa that modulates resistance to cationic antimicrobial peptides and addition of aminoarabinose to lipid A. J. Bacteriol. 2004, 186, 575–579. [Google Scholar] [CrossRef]
  305. Trent, M.S.; Ribeiro, A.A.; Lin, S.; Cotter, R.J.; Raetz, C.R. An inner membrane enzyme in Salmonella and Escherichia coli that transfers 4-amino-4-deoxy-L-arabinose to lipid A: Induction on polymyxin-resistant mutants and role of a novel lipid-linked donor. J. Biol. Chem. 2001, 276, 43122–43131. [Google Scholar] [CrossRef]
  306. Müller, A.; Wenzel, M.; Strahl, H.; Grein, F.; Saaki, T.N.V.; Kohl, B.; Siersma, T.; Bandow, J.E.; Sahl, H.G.; Schneider, T.; et al. Daptomycin inhibits cell envelope synthesis by interfering with fluid membrane microdomains. Proc. Natl. Acad. Sci. USA 2016, 113, E7077–E7086. [Google Scholar] [CrossRef]
  307. Mishra, N.N.; Tran, T.T.; Seepersaud, R.; Garcia-de-la-Maria, C.; Faull, K.; Yoon, A.; Proctor, R.; Miro, J.M.; Rybak, M.J.; Bayer, A.S.; et al. Perturbations of Phosphatidate Cytidylyltransferase (CdsA) Mediate Daptomycin Resistance in Streptococcus mitis/oralis by a Novel Mechanism. Antimicrob. Agents Chemother. 2017, 61, e02435-16. [Google Scholar] [CrossRef]
  308. Tran, T.T.; Panesso, D.; Gao, H.; Roh, J.H.; Munita, J.M.; Reyes, J.; Diaz, L.; Lobos, E.A.; Shamoo, Y.; Mishra, N.N.; et al. Whole-genome analysis of a daptomycin-susceptible enterococcus faecium strain and its daptomycin-resistant variant arising during therapy. Antimicrob. Agents Chemother. 2013, 57, 261–268. [Google Scholar] [CrossRef]
  309. Poshvina, D.V.; Dilbaryan, D.S.; Kasyanov, S.P.; Sadykova, V.S.; Lapchinskaya, O.A.; Rogozhin, E.A.; Vasilchenko, A.S. Staphylococcus aureus is able to generate resistance to novel lipoglycopeptide antibiotic gausemycin A. Front. Microbiol. 2022, 13, 963979. [Google Scholar] [CrossRef]
  310. Boudjemaa, R.; Cabriel, C.; Dubois-Brissonnet, F.; Bourg, N.; Dupuis, G.; Gruss, A.; Lévêque-Fort, S.; Briandet, R.; Fontaine-Aupart, M.P.; Steenkeste, K. Impact of Bacterial Membrane Fatty Acid Composition on the Failure of Daptomycin to Kill Staphylococcus aureus. Antimicrob. Agents Chemother. 2018, 62, e00023-18. [Google Scholar] [CrossRef]
  311. Ming, X.; Daeschel, M.A. Correlation of Cellular Phospholipid Content with Nisin Resistance of Listeria monocytogenes Scott A. J. Food Prot. 1995, 58, 416–420. [Google Scholar] [CrossRef]
  312. Gow, N.A.R.; Latge, J.P.; Munro, C.A. The Fungal Cell Wall: Structure, Biosynthesis, and Function. Microbiol. Spectr. 2017, 5, 10–1128. [Google Scholar] [CrossRef] [PubMed]
  313. Sudoh, M.; Yamazaki, T.; Masubuchi, K.; Taniguchi, M.; Shimma, N.; Arisawa, M.; Yamada-Okabe, H. Identification of a novel inhibitor specific to the fungal chitin synthase. Inhibition of chitin synthase 1 arrests the cell growth, but inhibition of chitin synthase 1 and 2 is lethal in the pathogenic fungus Candida albicans. J. Biol. Chem. 2000, 275, 32901–32905. [Google Scholar] [CrossRef]
  314. Garcia-Effron, G.; Park, S.; Perlin, D.S. Correlating echinocandin MIC and kinetic inhibition of fks1 mutant glucan synthases for Candida albicans: Implications for interpretive breakpoints. Antimicrob. Agents Chemother. 2009, 53, 112–122. [Google Scholar] [CrossRef] [PubMed]
  315. Onishi, J.; Meinz, M.; Thompson, J.; Curotto, J.; Dreikorn, S.; Rosenbach, M.; Douglas, C.; Abruzzo, G.; Flattery, A.; Kong, L.; et al. Discovery of novel antifungal (1,3)-beta-D-glucan synthase inhibitors. Antimicrob. Agents Chemother. 2000, 44, 368–377. [Google Scholar] [CrossRef] [PubMed]
  316. Healey, K.R.; Katiyar, S.K.; Raj, S.; Edlind, T.D. CRS-MIS in Candida glabrata: Sphingolipids modulate echinocandin-Fks interaction. Mol. Microbiol. 2012, 86, 303–313. [Google Scholar] [CrossRef]
  317. Satish, S.; Jiménez-Ortigosa, C.; Zhao, Y.; Lee, M.H.; Dolgov, E.; Krüger, T.; Park, S.; Denning, D.W.; Kniemeyer, O.; Brakhage, A.A.; et al. Stress-Induced Changes in the Lipid Microenvironment of β-(1,3)-d-Glucan Synthase Cause Clinically Important Echinocandin Resistance in Aspergillus fumigatus. mBio 2019, 10, e00779-19. [Google Scholar] [CrossRef]
  318. Ren, Z.; Chhetri, A.; Guan, Z.; Suo, Y.; Yokoyama, K.; Lee, S.Y. Structural basis for inhibition and regulation of a chitin synthase from Candida albicans. Nat. Struct. Mol. Biol. 2022, 29, 653–664. [Google Scholar] [CrossRef]
  319. Hu, X.; Yang, P.; Chai, C.; Liu, J.; Sun, H.; Wu, Y.; Zhang, M.; Zhang, M.; Liu, X.; Yu, H. Structural and mechanistic insights into fungal β-1,3-glucan synthase FKS1. Nature 2023, 616, 190–198. [Google Scholar] [CrossRef]
  320. Leibundgut, M.; Maier, T.; Jenni, S.; Ban, N. The multienzyme architecture of eukaryotic fatty acid synthases. Curr. Opin. Struct. Biol. 2008, 18, 714–725. [Google Scholar] [CrossRef]
  321. Maier, T.; Jenni, S.; Ban, N. Architecture of mammalian fatty acid synthase at 4.5 A resolution. Science 2006, 311, 1258–1262. [Google Scholar] [CrossRef]
  322. Jenni, S.; Leibundgut, M.; Maier, T.; Ban, N. Architecture of a fungal fatty acid synthase at 5 A resolution. Science 2006, 311, 1263–1267. [Google Scholar] [CrossRef]
  323. Chayakulkeeree, M.; Rude, T.H.; Toffaletti, D.L.; Perfect, J.R. Fatty acid synthesis is essential for survival of Cryptococcus neoformans and a potential fungicidal target. Antimicrob. Agents Chemother. 2007, 51, 3537–3545. [Google Scholar] [CrossRef]
  324. Zhao, X.J.; McElhaney-Feser, G.E.; Bowen, W.H.; Cole, M.F.; Broedel, S.E., Jr.; Cihlar, R.L. Requirement for the Candida albicans FAS2 gene for infection in a rat model of oropharyngeal candidiasis. Microbiology 1996, 142, 2509–2514. [Google Scholar] [CrossRef]
  325. Zhao, X.J.; McElhaney-Feser, G.E.; Sheridan, M.J.; Broedel, S.E., Jr.; Cihlar, R.L. Avirulence of Candida albicans FAS2 mutants in a mouse model of systemic candidiasis. Infect. Immun. 1997, 65, 829–832. [Google Scholar] [CrossRef]
  326. Faergeman, N.J.; Black, P.N.; Zhao, X.D.; Knudsen, J.; DiRusso, C.C. The Acyl-CoA synthetases encoded within FAA1 and FAA4 in Saccharomyces cerevisiae function as components of the fatty acid transport system linking import, activation, and intracellular Utilization. J. Biol. Chem. 2001, 276, 37051–37059. [Google Scholar] [CrossRef] [PubMed]
  327. Johansson, P.; Wiltschi, B.; Kumari, P.; Kessler, B.; Vonrhein, C.; Vonck, J.; Oesterhelt, D.; Grininger, M. Inhibition of the fungal fatty acid synthase type I multienzyme complex. Proc. Natl. Acad. Sci. USA 2008, 105, 12803–12808. [Google Scholar] [CrossRef]
  328. DeJarnette, C.; Meyer, C.J.; Jenner, A.R.; Butts, A.; Peters, T.; Cheramie, M.N.; Phelps, G.A.; Vita, N.A.; Loudon-Hossler, V.C.; Lee, R.E.; et al. Identification of Inhibitors of Fungal Fatty Acid Biosynthesis. ACS Infect. Dis. 2021, 7, 3210–3223. [Google Scholar] [CrossRef] [PubMed]
  329. Xu, D.; Sillaots, S.; Davison, J.; Hu, W.; Jiang, B.; Kauffman, S.; Martel, N.; Ocampo, P.; Oh, C.; Trosok, S.; et al. Chemical genetic profiling and characterization of small-molecule compounds that affect the biosynthesis of unsaturated fatty acids in Candida albicans. J. Biol. Chem. 2009, 284, 19754–19764. [Google Scholar] [CrossRef] [PubMed]
  330. Sant, D.G.; Tupe, S.G.; Ramana, C.V.; Deshpande, M.V. Fungal cell membrane-promising drug target for antifungal therapy. J. Appl. Microbiol. 2016, 121, 1498–1510. [Google Scholar] [CrossRef] [PubMed]
  331. Carman, G.M.; Han, G.S. Regulation of phospholipid synthesis in the yeast Saccharomyces cerevisiae. Annu. Rev. Biochem. 2011, 80, 859–883. [Google Scholar] [CrossRef]
  332. Pan, J.; Hu, C.; Yu, J.H. Lipid Biosynthesis as an Antifungal Target. J. Fungi 2018, 4, 50. [Google Scholar] [CrossRef] [PubMed]
  333. McDonough, V.M.; Buxeda, R.J.; Bruno, M.E.; Ozier-Kalogeropoulos, O.; Adeline, M.T.; McMaster, C.R.; Bell, R.M.; Carman, G.M. Regulation of phospholipid biosynthesis in Saccharomyces cerevisiae by CTP. J. Biol. Chem. 1995, 270, 18774–18780. [Google Scholar] [CrossRef] [PubMed]
  334. Tams, R.N.; Cassilly, C.D.; Anaokar, S.; Brewer, W.T.; Dinsmore, J.T.; Chen, Y.L.; Patton-Vogt, J.; Reynolds, T.B. Overproduction of Phospholipids by the Kennedy Pathway Leads to Hypervirulence in Candida albicans. Front. Microbiol. 2019, 10, 86. [Google Scholar] [CrossRef] [PubMed]
  335. Kim, Y.S.; Oh, J.Y.; Hwang, B.K.; Kim, K.D. Variation in sensitivity of Magnaporthe oryzae isolates from Korea to edifenphos and iprobenfos. Crop Prot. 2008, 27, 1464–1470. [Google Scholar] [CrossRef]
  336. Kennedy, E.P.; Weiss, S.B. The function of cytidine coenzymes in the biosynthesis of phospholipides. J. Biol. Chem. 1956, 222, 193–214. [Google Scholar] [CrossRef]
  337. Kennedy, E.P. Metabolism of lipides. Annu. Rev. Biochem. 1957, 26, 119–148. [Google Scholar] [CrossRef]
  338. Nagiec, M.M.; Nagiec, E.E.; Baltisberger, J.A.; Wells, G.B.; Lester, R.L.; Dickson, R.C. Sphingolipid synthesis as a target for antifungal drugs. Complementation of the inositol phosphorylceramide synthase defect in a mutant strain of Saccharomyces cerevisiae by the AUR1 gene. J. Biol. Chem. 1997, 272, 9809–9817. [Google Scholar] [CrossRef]
  339. Dickson, R.C. Roles for sphingolipids in Saccharomyces cerevisiae. Adv. Exp. Med. Biol. 2010, 688, 217–231. [Google Scholar] [CrossRef]
  340. Obeid, L.M.; Okamoto, Y.; Mao, C. Yeast sphingolipids: Metabolism and biology. Biochim. Biophys. Acta 2002, 1585, 163–171. [Google Scholar] [CrossRef]
  341. Ren, J.; Snider, J.; Airola, M.V.; Zhong, A.; Rana, N.A.; Obeid, L.M.; Hannun, Y.A. Quantification of 3-ketodihydrosphingosine using HPLC-ESI-MS/MS to study SPT activity in yeast Saccharomyces cerevisiae. J. Lipid Res. 2018, 59, 162–170. [Google Scholar] [CrossRef]
  342. Dickson, R.C.; Lester, R.L. Sphingolipid functions in Saccharomyces cerevisiae. Biochim. Biophys. Acta 2002, 1583, 13–25. [Google Scholar] [CrossRef]
  343. Dickson, R.C.; Sumanasekera, C.; Lester, R.L. Functions and metabolism of sphingolipids in Saccharomyces cerevisiae. Prog. Lipid Res. 2006, 45, 447–465. [Google Scholar] [CrossRef]
  344. Whaley, H.A. The structure of lipoxamycin, a novel antifungal antibiotic. J. Am. Chem. Soc. 1971, 93, 3767–3769. [Google Scholar] [CrossRef]
  345. Mandala, S.M.; Frommer, B.R.; Thornton, R.A.; Kurtz, M.B.; Young, N.M.; Cabello, M.A.; Genilloud, O.; Liesch, J.M.; Smith, J.L.; Horn, W.S. Inhibition of serine palmitoyl-transferase activity by lipoxamycin. J. Antibiot. 1994, 47, 376–379. [Google Scholar] [CrossRef]
  346. Kluepfel, D.; Bagli, J.; Baker, H.; Charest, M.P.; Kudelski, A. Myriocin, a new antifungal antibiotic from Myriococcum albomyces. J. Antibiot. 1972, 25, 109–115. [Google Scholar] [CrossRef] [PubMed]
  347. VanMiddlesworth, F.; Giacobbe, R.A.; Lopez, M.; Garrity, G.; Bland, J.A.; Bartizal, K.; Fromtling, R.A.; Polishook, J.; Zweerink, M.; Edison, A.M.; et al. Sphingofungins A, B, C, and D; a new family of antifungal agents. I. Fermentation, isolation, and biological activity. J. Antibiot. 1992, 45, 861–867. [Google Scholar] [CrossRef]
  348. Mandala, S.M.; Thornton, R.A.; Frommer, B.R.; Curotto, J.E.; Rozdilsky, W.; Kurtz, M.B.; Giacobbe, R.A.; Bills, G.F.; Cabello, M.A.; Martín, I.; et al. The discovery of australifungin, a novel inhibitor of sphinganine N-acyltransferase from Sporormiella australis. Producing organism, fermentation, isolation, and biological activity. J. Antibiot. 1995, 48, 349–356. [Google Scholar] [CrossRef] [PubMed]
  349. Raguž, L.; Peng, C.C.; Kaiser, M.; Görls, H.; Beemelmanns, C. A Modular Approach to the Antifungal Sphingofungin Family: Concise Total Synthesis of Sphingofungin A and C. Angew. Chem. Int. Ed. Engl. 2022, 61, e202112616. [Google Scholar] [CrossRef] [PubMed]
  350. Mandala, S.M.; Thornton, R.A.; Frommer, B.R.; Dreikorn, S.; Kurtz, M.B. Viridiofungins, novel inhibitors of sphingolipid synthesis. J. Antibiot. 1997, 50, 339–343. [Google Scholar] [CrossRef]
  351. Delgado, A.; Casas, J.; Llebaria, A.; Abad, J.L.; Fabrias, G. Inhibitors of sphingolipid metabolism enzymes. Biochim. Biophys. Acta 2006, 1758, 1957–1977. [Google Scholar] [CrossRef]
  352. Gable, K.; Slife, H.; Bacikova, D.; Monaghan, E.; Dunn, T.M. Tsc3p is an 80-amino acid protein associated with serine palmitoyltransferase and required for optimal enzyme activity. J. Biol. Chem. 2000, 275, 7597–7603. [Google Scholar] [CrossRef]
  353. Yoo, H.S.; Norred, W.P.; Wang, E.; Merrill, A.H., Jr.; Riley, R.T. Fumonisin inhibition of de novo sphingolipid biosynthesis and cytotoxicity are correlated in LLC-PK1 cells. Toxicol. Appl. Pharmacol. 1992, 114, 9–15. [Google Scholar] [CrossRef] [PubMed]
  354. McEvoy, K.; Normile, T.G.; Del Poeta, M. Antifungal Drug Development: Targeting the Fungal Sphingolipid Pathway. J. Fungi 2020, 6, 142. [Google Scholar] [CrossRef] [PubMed]
  355. Heidler, S.A.; Radding, J.A. The AUR1 gene in Saccharomyces cerevisiae encodes dominant resistance to the antifungal agent aureobasidin A (LY295337). Antimicrob. Agents Chemother. 1995, 39, 2765–2769. [Google Scholar] [CrossRef]
  356. Mandala, S.M.; Thornton, R.A.; Rosenbach, M.; Milligan, J.; Garcia-Calvo, M.; Bull, H.G.; Kurtz, M.B. Khafrefungin, a novel inhibitor of sphingolipid synthesis. J. Biol. Chem. 1997, 272, 32709–32714. [Google Scholar] [CrossRef]
  357. Ohnuki, T.; Yano, T.; Ono, Y.; Kozuma, S.; Suzuki, T.; Ogawa, Y.; Takatsu, T. Haplofungins, novel inositol phosphorylceramide synthase inhibitors, from Lauriomyces bellulus SANK 26899 I. Taxonomy, fermentation, isolation and biological activities. J. Antibiot. 2009, 62, 545–549. [Google Scholar] [CrossRef] [PubMed]
  358. Yano, T.; Aoyagi, A.; Kozuma, S.; Kawamura, Y.; Tanaka, I.; Suzuki, Y.; Takamatsu, Y.; Takatsu, T.; Inukai, M. Pleofungins, novel inositol phosphorylceramide synthase inhibitors, from Phoma sp. SANK 13899. I. Taxonomy, fermentation, isolation, and biological activities. J. Antibiot. 2007, 60, 136–142. [Google Scholar] [CrossRef]
  359. Mandala, S.M.; Thornton, R.A.; Milligan, J.; Rosenbach, M.; Garcia-Calvo, M.; Bull, H.G.; Harris, G.; Abruzzo, G.K.; Flattery, A.M.; Gill, C.J.; et al. Rustmicin, a potent antifungal agent, inhibits sphingolipid synthesis at inositol phosphoceramide synthase. J. Biol. Chem. 1998, 273, 14942–14949. [Google Scholar] [CrossRef]
  360. Harris, G.H.; Shafiee, A.; Cabello, M.A.; Curotto, J.E.; Genilloud, O.; Göklen, K.E.; Kurtz, M.B.; Rosenbach, M.; Salmon, P.M.; Thornton, R.A.; et al. Inhibition of fungal sphingolipid biosynthesis by rustmicin, galbonolide B and their new 21-hydroxy analogs. J. Antibiot. 1998, 51, 837–844. [Google Scholar] [CrossRef]
  361. Fauth, U.; Zähner, H.; Mühlenfeld, A.; Achenbach, H. Galbonolides A and B--two non-glycosidic antifungal macrolides. J. Antibiot. 1986, 39, 1760–1764. [Google Scholar] [CrossRef] [PubMed]
  362. Nakamura, M.; Mori, Y.; Okuyama, K.; Tanikawa, K.; Yasuda, S.; Hanada, K.; Kobayashi, S. Chemistry and biology of khafrefungin. Large-scale synthesis, design, and structure-activity relationship of khafrefungin, an antifungal agent. Org. Biomol. Chem. 2003, 1, 3362–3376. [Google Scholar] [CrossRef] [PubMed]
  363. Zhong, W.; Jeffries, M.W.; Georgopapadakou, N.H. Inhibition of inositol phosphorylceramide synthase by aureobasidin A in Candida and Aspergillus species. Antimicrob. Agents Chemother. 2000, 44, 651–653. [Google Scholar] [CrossRef] [PubMed]
  364. Dupont, S.; Lemetais, G.; Ferreira, T.; Cayot, P.; Gervais, P.; Beney, L. Ergosterol biosynthesis: A fungal pathway for life on land? Evolution 2012, 66, 2961–2968. [Google Scholar] [CrossRef]
  365. Jordá, T.; Puig, S. Regulation of Ergosterol Biosynthesis in Saccharomyces cerevisiae. Genes 2020, 11, 795. [Google Scholar] [CrossRef]
  366. Hu, Z.; He, B.; Ma, L.; Sun, Y.; Niu, Y.; Zeng, B. Recent Advances in Ergosterol Biosynthesis and Regulation Mechanisms in Saccharomyces cerevisiae. Indian J. Microbiol. 2017, 57, 270–277. [Google Scholar] [CrossRef]
  367. Miziorko, H.M. Enzymes of the mevalonate pathway of isoprenoid biosynthesis. Arch. Biochem. Biophys. 2011, 505, 131–143. [Google Scholar] [CrossRef]
  368. Tavakkoli, A.; Johnston, T.P.; Sahebkar, A. Antifungal effects of statins. Pharmacol. Ther. 2020, 208, 107483. [Google Scholar] [CrossRef]
  369. Ting, M.; Whitaker, E.J.; Albandar, J.M. Systematic review of the In vitro effects of statins on oral and perioral microorganisms. Eur. J. Oral Sci. 2016, 124, 4–10. [Google Scholar] [CrossRef]
  370. Westermeyer, C.; Macreadie, I.G. Simvastatin reduces ergosterol levels, inhibits growth and causes loss of mtDNA in Candida glabrata. FEMS Yeast Res. 2007, 7, 436–441. [Google Scholar] [CrossRef]
  371. Gyetvai, A.; Emri, T.; Takács, K.; Dergez, T.; Fekete, A.; Pesti, M.; Pócsi, I.; Lenkey, B. Lovastatin possesses a fungistatic effect against Candida albicans, but does not trigger apoptosis in this opportunistic human pathogen. FEMS Yeast Res. 2006, 6, 1140–1148. [Google Scholar] [CrossRef]
  372. Liu, G.; Vellucci, V.F.; Kyc, S.; Hostetter, M.K. Simvastatin inhibits Candida albicans biofilm In vitro. Pediatr. Res. 2009, 66, 600–604. [Google Scholar] [CrossRef]
  373. Rosales-Acosta, B.; Mendieta, A.; Zúñiga, C.; Tamariz, J.; Hernández Rodríguez, C.; Ibarra-García, J.A.; Villa-Tanaca, L. Simvastatin and other inhibitors of the enzyme 3-hydroxy-3-methylglutaryl coenzyme A reductase of Ustilago maydis (Um-Hmgr) affect the viability of the fungus, its synthesis of sterols and mating. Rev. Iberoam. Micol. 2019, 36, 1–8. [Google Scholar] [CrossRef] [PubMed]
  374. Florin-Christensen, M.; Florin-Christensen, J.; Garin, C.; Isola, E.; Brenner, R.R.; Rasmussen, L. Inhibition of Trypanosoma cruzi growth and sterol biosynthesis by lovastatin. Biochem. Biophys. Res. Commun. 1990, 166, 1441–1445. [Google Scholar] [CrossRef] [PubMed]
  375. Callegari, S.; McKinnon, R.A.; Andrews, S.; de Barros Lopes, M.A. Atorvastatin-induced cell toxicity in yeast is linked to disruption of protein isoprenylation. FEMS Yeast Res. 2010, 10, 1881–1898. [Google Scholar] [CrossRef] [PubMed]
  376. Qiao, J.; Kontoyiannis, D.P.; Wan, Z.; Li, R.; Liu, W. Antifungal activity of statins against Aspergillus species. Med. Mycol. 2007, 45, 589–593. [Google Scholar] [CrossRef] [PubMed]
  377. Macreadie, I.G.; Johnson, G.; Schlosser, T.; Macreadie, P.I. Growth inhibition of Candida species and Aspergillus fumigatus by statins. FEMS Microbiol. Lett. 2006, 262, 9–13. [Google Scholar] [CrossRef]
  378. Chin, N.X.; Weitzman, I.; Della-Latta, P. In vitro activity of fluvastatin, a cholesterol-lowering agent, and synergy with flucanazole and itraconazole against Candida species and Cryptococcus neoformans. Antimicrob. Agents Chemother. 1997, 41, 850–852. [Google Scholar] [CrossRef] [PubMed]
  379. Chamilos, G.; Lewis, R.E.; Kontoyiannis, D.P. Lovastatin has significant activity against zygomycetes and interacts synergistically with voriconazole. Antimicrob. Agents Chemother. 2006, 50, 96–103. [Google Scholar] [CrossRef]
  380. Madrigal-Aguilar, D.A.; Gonzalez-Silva, A.; Rosales-Acosta, B.; Bautista-Crescencio, C.; Ortiz-Álvarez, J.; Escalante, C.H.; Sánchez-Navarrete, J.; Hernández-Rodríguez, C.; Chamorro-Cevallos, G.; Tamariz, J.; et al. Antifungal Activity of Fibrate-Based Compounds and Substituted Pyrroles That Inhibit the Enzyme 3-Hydroxy-methyl-glutaryl-CoA Reductase of Candida glabrata (CgHMGR), Thus Decreasing Yeast Viability and Ergosterol Synthesis. Microbiol. Spectr. 2022, 10, e0164221. [Google Scholar] [CrossRef]
  381. Brilhante, R.S.N.; Fonseca, X.M.Q.C.; Pereira, V.S.; Araújo, G.D.S.; Oliveira, J.S.; Garcia, L.G.S.; Rodrigues, A.M.; Camargo, Z.P.; Pereira-Neto, W.A.; Castelo-Branco, D.S.C.M.; et al. In vitro inhibitory effect of statins on planktonic cells and biofilms of the Sporothrix schenckii species complex. J. Med. Microbiol. 2020, 69, 838–843. [Google Scholar] [CrossRef]
  382. Darwazeh, A.; Al-Shorman, H.; Mrayan, B. Effect of statin therapy on oral Candida carriage in hyperlipidemia patients: A pioneer study. Dent. Med. Probl. 2022, 59, 93–97. [Google Scholar] [CrossRef] [PubMed]
  383. Bellanger, A.P.; Tatara, A.M.; Shirazi, F.; Gebremariam, T.; Albert, N.D.; Lewis, R.E.; Ibrahim, A.S.; Kontoyiannis, D.P. Statin Concentrations Below the Minimum Inhibitory Concentration Attenuate the Virulence of Rhizopus oryzae. J. Infect. Dis. 2016, 214, 114–121. [Google Scholar] [CrossRef] [PubMed]
  384. Tavakkoli, A.; Johnston, T.P.; Sahebkar, A. Fluvastatin: A choice for COVID-19 associated mucormycosis management. Curr. Med. Chem. 2023, 6. E-pub Ahead of Print. [Google Scholar] [CrossRef]
  385. Hussain, M.K.; Ahmed, S.; Khan, A.; Siddiqui, A.J.; Khatoon, S.; Jahan, S. Mucormycosis: A hidden mystery of fungal infection, possible diagnosis, treatment and development of new therapeutic agents. Eur. J. Med. Chem. 2023, 246, 115010. [Google Scholar] [CrossRef] [PubMed]
  386. Klug, L.; Daum, G. Yeast lipid metabolism at a glance. FEMS Yeast Res. 2014, 14, 369–388. [Google Scholar] [CrossRef]
  387. Plochocka, D.; Karst, F.; Swiezewska, E.; Szkopińska, A. The role of ERG20 gene (encoding yeast farnesyl diphosphate synthase) mutation in long dolichol formation. Molecular modeling of FPP synthase. Biochimie 2000, 82, 733–738. [Google Scholar] [CrossRef]
  388. Bergstrom, J.D.; Dufresne, C.; Bills, G.F.; Nallin-Omstead, M.; Byrne, K. Discovery, biosynthesis, and mechanism of action of the zaragozic acids: Potent inhibitors of squalene synthase. Annu. Rev. Microbiol. 1995, 49, 607–639. [Google Scholar] [CrossRef]
  389. Pospiech, M.; Owens, S.E.; Miller, D.J.; Austin-Muttitt, K.; Mullins, J.G.L.; Cronin, J.G.; Allemann, R.K.; Sheldon, I.M. Bisphosphonate inhibitors of squalene synthase protect cells against cholesterol-dependent cytolysins. FASEB J. 2021, 35, e21640. [Google Scholar] [CrossRef]
  390. Ryder, N.S. Inhibition of squalene epoxidase and sterol side-chain methylation by allylamines. Biochem. Soc. Trans. 1990, 18, 45–46. [Google Scholar] [CrossRef]
  391. Birnbaum, J.E. Pharmacology of the allylamines. J. Am. Acad. Dermatol. 1990, 23, 782–785. [Google Scholar] [CrossRef]
  392. Favre, B.; Ryder, N.S. Characterization of squalene epoxidase activity from the dermatophyte Trichophyton rubrum and its inhibition by terbinafine and other antimycotic agents. Antimicrob. Agents Chemother. 1996, 40, 443–447. [Google Scholar] [CrossRef]
  393. Astvad, K.M.T.; Hare, R.K.; Jørgensen, K.M.; Saunte, D.M.L.; Thomsen, P.K.; Arendrup, M.C. Increasing Terbinafine Resistance in Danish Trichophyton Isolates 2019–2020. J. Fungi 2022, 8, 150. [Google Scholar] [CrossRef] [PubMed]
  394. Moreno-Sabater, A.; Normand, A.C.; Bidaud, A.L.; Cremer, G.; Foulet, F.; Brun, S.; Bonnal, C.; Aït-Ammar, N.; Jabet, A.; Ayachi, A.; et al. Terbinafine Resistance in Dermatophytes: A French Multicenter Prospective Study. J. Fungi 2022, 8, 220. [Google Scholar] [CrossRef]
  395. Gaurav, V.; Bhattacharya, S.N.; Sharma, N.; Datt, S.; Kumar, P.; Rai, G.; Singh, P.K.; Taneja, B.; Das, S. Terbinafine resistance in dermatophytes: Time to revisit alternate antifungal therapy. J. Mycol. Med. 2021, 31, 101087. [Google Scholar] [CrossRef] [PubMed]
  396. Ajit, C.; Suvannasankha, A.; Zaeri, N.; Munoz, S.J. Terbinafine-associated hepatotoxicity. Am. J. Med. Sci. 2003, 325, 292–295. [Google Scholar] [CrossRef] [PubMed]
  397. Trösken, E.R.; Adamska, M.; Arand, M.; Zarn, J.A.; Patten, C.; Völkel, W.; Lutz, W.K. Comparison of lanosterol-14 alpha-demethylase (CYP51) of human and Candida albicans for inhibition by different antifungal azoles. Toxicology 2006, 228, 24–32. [Google Scholar] [CrossRef] [PubMed]
  398. Singh, A.; Singh, K.; Sharma, A.; Kaur, K.; Chadha, R.; Bedi, P.M.S. Recent advances in antifungal drug development targeting lanosterol 14α-demethylase (CYP51): A comprehensive review with structural and molecular insights. Chem. Biol. Drug Des. 2023, 102, 606–639. [Google Scholar] [CrossRef]
  399. Sabatelli, F.; Patel, R.; Mann, P.A.; Mendrick, C.A.; Norris, C.C.; Hare, R.; Loebenberg, D.; Black, T.A.; McNicholas, P.M. In vitro activities of posaconazole, fluconazole, itraconazole, voriconazole, and amphotericin B against a large collection of clinically important molds and yeasts. Antimicrob. Agents Chemother. 2006, 50, 2009–2015. [Google Scholar] [CrossRef]
  400. Shafiei, M.; Peyton, L.; Hashemzadeh, M.; Foroumadi, A. History of the development of antifungal azoles: A review on structures, SAR, and mechanism of action. Bioorg. Chem. 2020, 104, 104240. [Google Scholar] [CrossRef]
  401. Khan, A.; Iqbal, A.; Ahmedi, S.; Manzoor, N.; Siddiqui, T. Synthesis, Anti-Fungal Potency and In silico Studies of Novel Steroidal 1,4-Dihydropyridines. Chem. Biodivers. 2023, 20, e202300096. [Google Scholar] [CrossRef]
  402. Pristov, K.E.; Ghannoum, M.A. Resistance of Candida to azoles and echinocandins worldwide. Clin. Microbiol. Infect. 2019, 25, 792–798. [Google Scholar] [CrossRef] [PubMed]
  403. Paul, S.; Shaw, D.; Joshi, H.; Singh, S.; Chakrabarti, A.; Rudramurthy, S.M.; Ghosh, A.K. Mechanisms of azole antifungal resistance in clinical isolates of Candida tropicalis. PLoS ONE 2022, 17, e0269721. [Google Scholar] [CrossRef]
  404. Husselstein, T.; Schaller, H.; Gachotte, D.; Benveniste, P. Delta7-sterol-C5-desaturase: Molecular characterization and functional expression of wild-type and mutant alleles. Plant Mol. Biol. 1999, 39, 891–906. [Google Scholar] [CrossRef] [PubMed]
  405. Ganapathy, K.; Jones, C.W.; Stephens, C.M.; Vatsyayan, R.; Marshall, J.A.; Nes, W.D. Molecular probing of the Saccharomyces cerevisiae sterol 24-C methyltransferase reveals multiple amino acid residues involved with C2-transfer activity. Biochim. Biophys. Acta 2008, 1781, 344–351. [Google Scholar] [CrossRef] [PubMed]
  406. Kaneshiro, E.S.; Johnston, L.Q.; Nkinin, S.W.; Romero, B.I.; Giner, J.L. Sterols of Saccharomyces cerevisiae erg6 Knockout Mutant Expressing the Pneumocystis carinii S-Adenosylmethionine:Sterol C-24 Methyltransferase. J. Eukaryot. Microbiol. 2015, 62, 298–306. [Google Scholar] [CrossRef] [PubMed]
  407. Krauß, J.; Müller, C.; Klimt, M.; Valero, L.J.; Martínez, J.F.; Müller, M.; Bartel, K.; Binder, U.; Bracher, F. Synthesis, Biological Evaluation, and Structure-Activity Relationships of 4-Aminopiperidines as Novel Antifungal Agents Targeting Ergosterol Biosynthesis. Molecules 2021, 26, 7208. [Google Scholar] [CrossRef] [PubMed]
  408. Jachak, G.R.; Ramesh, R.; Sant, D.G.; Jorwekar, S.U.; Jadhav, M.R.; Tupe, S.G.; Deshpande, M.V.; Reddy, D.S. Silicon Incorporated Morpholine Antifungals: Design, Synthesis, and Biological Evaluation. ACS Med. Chem. Lett. 2015, 6, 1111–1116. [Google Scholar] [CrossRef]
  409. Hata, M.; Yoshida, K.; Ishii, C.; Otani, T.; Ando, A. In vitro and in vivo antifungal activities of aminopiperidine derivatives, novel ergosterol synthesis inhibitors. Biol. Pharm. Bull. 2010, 33, 473–476. [Google Scholar] [CrossRef]
  410. Mitsche, M.A.; McDonald, J.G.; Hobbs, H.H.; Cohen, J.C. Flux analysis of cholesterol biosynthesis in vivo reveals multiple tissue and cell-type specific pathways. Elife 2015, 4, e07999. [Google Scholar] [CrossRef]
  411. Benveniste, P. Sterol metabolism. Arab. Book 2002, 1, e0004. [Google Scholar] [CrossRef] [PubMed]
  412. Warrilow, A.G.; Parker, J.E.; Kelly, D.E.; Kelly, S.L. Azole affinity of sterol 14α-demethylase (CYP51) enzymes from Candida albicans and Homo sapiens. Antimicrob. Agents Chemother. 2013, 57, 1352–1360. [Google Scholar] [CrossRef] [PubMed]
  413. Warrilow, A.G.; Parker, J.E.; Price, C.L.; Nes, W.D.; Garvey, E.P.; Hoekstra, W.J.; Schotzinger, R.J.; Kelly, D.E.; Kelly, S.L. The Investigational Drug VT-1129 Is a Highly Potent Inhibitor of Cryptococcus Species CYP51 but Only Weakly Inhibits the Human Enzyme. Antimicrob. Agents Chemother. 2016, 60, 4530–4538. [Google Scholar] [CrossRef] [PubMed]
  414. Warrilow, A.G.; Price, C.L.; Parker, J.E.; Rolley, N.J.; Smyrniotis, C.J.; Hughes, D.D.; Thoss, V.; Nes, W.D.; Kelly, D.E.; Holman, T.R.; et al. Azole Antifungal Sensitivity of Sterol 14α-Demethylase (CYP51) and CYP5218 from Malassezia globosa. Sci. Rep. 2016, 6, 27690. [Google Scholar] [CrossRef]
  415. Hata, M.; Ishii, Y.; Watanabe, E.; Uoto, K.; Kobayashi, S.; Yoshida, K.; Otani, T.; Ando, A. Inhibition of ergosterol synthesis by novel antifungal compounds targeting C-14 reductase. Med. Mycol. 2010, 48, 613–621. [Google Scholar] [CrossRef] [PubMed]
  416. Agner, G.; Kaulin, Y.A.; Gurnev, P.A.; Szabo, Z.; Schagina, L.V.; Takemoto, J.Y.; Blasko, K. Membrane-permeabilizing activities of cyclic lipodepsipeptides, syringopeptin 22A and syringomycin E from Pseudomonas syringae pv. syringae in human red blood cells and in bilayer lipid membranes. Bioelectrochemistry 2000, 52, 161–167. [Google Scholar]
  417. Hutchison, M.L.; Gross, D.C. Lipopeptide phytotoxins produced by Pseudomonas syringae pv. syringae: Comparison of the biosurfactant and ion channel-forming activities of syringopeptin and syringomycin. Mol. Plant Microbe Interact. 1997, 10, 347–354. [Google Scholar] [CrossRef]
  418. Bensaci, M.F.; Gurnev, P.A.; Bezrukov, S.M.; Takemoto, J.Y. Fungicidal Activities and Mechanisms of Action of Pseudomonas syringae pv. syringae Lipodepsipeptide Syringopeptins 22A and 25A. Front. Microbiol. 2011, 2, 216. [Google Scholar] [CrossRef]
  419. Falardeau, J.; Wise, C.; Novitsky, L.; Avis, T.J. Ecological and mechanistic insights into the direct and indirect antimicrobial properties of Bacillus subtilis lipopeptides on plant pathogens. J. Chem. Ecol. 2013, 39, 869–878. [Google Scholar] [CrossRef]
  420. Hutchison, M.L.; Tester, M.A.; Gross, D.C. Role of biosurfactant and ion channel-forming activities of syringomycin in transmembrane ion flux: A model for the mechanism of action in the plant-pathogen interaction. Mol. Plant Microbe Interact. 1995, 8, 610–620. [Google Scholar] [CrossRef]
  421. Maget-Dana, R.; Heitz, F.; Ptak, M.; Peypoux, F.; Guinand, M. Bacterial lipopeptides induce ion-conducting pores in planar bilayers. Biochem. Biophys. Res. Commun. 1985, 129, 965–971. [Google Scholar] [CrossRef]
  422. Malev, V.V.; Schagina, L.V.; Gurnev, P.A.; Takemoto, J.Y.; Nestorovich, E.M.; Bezrukov, S.M. Syringomycin E channel: A lipidic pore stabilized by lipopeptide? Biophys. J. 2002, 82, 1985–1994. [Google Scholar] [CrossRef]
  423. Ostroumova, O.S.; Gurnev, P.A.; Schagina, L.V.; Bezrukov, S.M. Asymmetry of syringomycin E channel studied by polymer partitioning. FEBS Lett. 2007, 581, 804–808. [Google Scholar] [CrossRef] [PubMed]
  424. Ostroumova, O.S.; Malev, V.V.; Ilin, M.G.; Schagina, L.V. Surfactin activity depends on the membrane dipole potential. Langmuir 2010, 26, 15092–15097. [Google Scholar] [CrossRef] [PubMed]
  425. Zakharova, A.A.; Efimova, S.S.; Malev, V.V.; Ostroumova, O.S. Fengycin induces ion channels in lipid bilayers mimicking target fungal cell membranes. Sci. Rep. 2019, 9, 16034. [Google Scholar] [CrossRef] [PubMed]
  426. Sheppard, J.D.; Jumarie, C.; Cooper, D.G.; Laprade, R. Ionic channels induced by surfactin in planar lipid bilayer membranes. Biochim. Biophys. Acta 1991, 1064, 13–23. [Google Scholar] [CrossRef] [PubMed]
  427. Maget-Dana, R.; Peypoux, F. Iturins, a special class of pore-forming lipopeptides: Biological and physicochemical properties. Toxicology 1994, 87, 151–174. [Google Scholar] [CrossRef]
  428. Dalla Serra, M.; Bernhart, I.; Nordera, P.; Di Giorgio, D.; Ballio, A.; Menestrina, G. Conductive properties and gating of channels formed by syringopeptin 25A, a bioactive lipodepsipeptide from Pseudomonas syringae pv. syringae, in planar lipid membranes. Mol. Plant Microbe Interact. 1999, 12, 401–409. [Google Scholar] [CrossRef]
  429. Maget-Dana, R.; Ptak, M.; Peypoux, F.; Michel, G. Pore-forming properties of iturin A, a lipopeptide antibiotic. Biochim. Biophys. Acta 1985, 815, 405–409. [Google Scholar] [CrossRef]
  430. Maget-Dana, R.; Ptak, M. Iturin lipopeptides: Interactions of mycosubtilin with lipids in planar membranes and mixed monolayers. Biochim. Biophys. Acta 1990, 1023, 34–40. [Google Scholar] [CrossRef] [PubMed]
  431. Ermishkin, L.N.; Kasumov, K.M.; Potzeluyev, V.M. Single ionic channels induced in lipid bilayers by polyene antibiotics amphotericin B and nystatine. Nature 1976, 262, 698–699. [Google Scholar] [CrossRef]
  432. Ostroumova, O.S.; Efimova, S.S.; Schagina, L.V. Probing amphotericin B single channel activity by membrane dipole modifiers. PLoS ONE 2012, 7, e30261. [Google Scholar] [CrossRef] [PubMed]
  433. Tevyashova, A.; Efimova, S.; Alexandrov, A.; Omelchuk, O.; Ghazy, E.; Bychkova, E.; Zatonsky, G.; Grammatikova, N.; Dezhenkova, L.; Solovieva, S.; et al. Semisynthetic Amides of Amphotericin B and Nystatin A1: A Comparative Study of In vitro Activity/Toxicity Ratio in Relation to Selectivity to Ergosterol Membranes. Antibiotics 2023, 12, 151. [Google Scholar] [CrossRef] [PubMed]
  434. Samedova, A.A.; Kasumov, K.M. Mechanism of action of macrolide antibiotic filipin on cell and lipid membranes. Antibiot. Khimioter. 2009, 54, 44–52. [Google Scholar] [PubMed]
  435. Ostroumova, O.S.; Efimova, S.S.; Chulkov, E.G.; Schagina, L.V. The interaction of dipole modifiers with polyene-sterol complexes. PLoS ONE 2012, 7, e45135. [Google Scholar] [CrossRef]
  436. Campagna, S.; Saint, N.; Molle, G.; Aumelas, A. Structure and mechanism of action of the antimicrobial peptide piscidin. Biochemistry 2007, 46, 1771–1778. [Google Scholar] [CrossRef]
  437. Tevyashova, A.N.; Efimova, S.S.; Alexandrov, A.I.; Ghazy, E.S.M.O.; Bychkova, E.N.; Solovieva, S.E.; Zatonsky, G.B.; Grammatikova, N.E.; Dezhenkova, L.G.; Pereverzeva, E.R.; et al. Semisynthetic Amides of Polyene Antibiotic Natamycin. ACS Infect. Dis. 2023, 9, 42–55. [Google Scholar] [CrossRef]
  438. Andreoli, T.E. The structure and function of amphotericin B-cholesterol pores in lipid bilayer membranes. Ann. N. Y. Acad. Sci. 1974, 235, 448–468. [Google Scholar] [CrossRef]
  439. De Kruijff, B.; Gerritsen, W.J.; Oerlemans, A.; Demel, R.A.; van Deenen, L.L. Polyene antibiotic-sterol interactions in membranes of Acholeplasma laidlawii cells and lecithin liposomes. I. Specificity of the membrane permeability changes induced by the polyene antibiotics. Biochim. Biophys. Acta 1974, 339, 30–43. [Google Scholar] [CrossRef]
  440. Cohen, B.E. Amphotericin B membrane action: Role for two types of ion channels in eliciting cell survival and lethal effects. J. Membr. Biol. 2010, 238, 1–20. [Google Scholar] [CrossRef]
  441. Dos Santos, A.G.; Marquês, J.T.; Carreira, A.C.; Castro, I.R.; Viana, A.S.; Mingeot-Leclercq, M.P.; de Almeida, R.F.M.; Silva, L.C. The molecular mechanism of Nystatin action is dependent on the membrane biophysical properties and lipid composition. Phys. Chem. Chem. Phys. 2017, 19, 30078–30088. [Google Scholar] [CrossRef] [PubMed]
  442. Baghirova, A.A.; Kasumov, K.M. Antifungal Macrocycle Antibiotic Amphotericin B-Its Present and Future. Multidisciplinary Perspective for the Use in the Medical Practice. Biochem. Mosc. Suppl. B Biomed. Chem. 2022, 16, 1–12. [Google Scholar] [CrossRef] [PubMed]
  443. Chulkov, E.G.; Schagina, L.V.; Ostroumova, O.S. Membrane dipole modifiers modulate single-length nystatin channels via reducing elastic stress in the vicinity of the lipid mouth of a pore. Biochim. Biophys. Acta 2015, 1848, 192–199. [Google Scholar] [CrossRef] [PubMed]
  444. Umegawa, Y.; Yamamoto, T.; Dixit, M.; Funahashi, K.; Seo, S.; Nakagawa, Y.; Suzuki, T.; Matsuoka, S.; Tsuchikawa, H.; Hanashima, S.; et al. Amphotericin B assembles into seven-molecule ion channels: An NMR and molecular dynamics study. Sci. Adv. 2022, 8, eabo2658. [Google Scholar] [CrossRef]
  445. Akkerman, V.; Scheidt, H.A.; Reinholdt, P.; Bashawat, M.; Szomek, M.; Lehmann, M.; Wessig, P.; Covey, D.F.; Kongsted, J.; Müller, P.; et al. Natamycin interferes with ergosterol-dependent lipid phases in model membranes. BBA Adv. 2023, 4, 100102. [Google Scholar] [CrossRef] [PubMed]
  446. Iwamoto, M.; Sumino, A.; Shimada, E.; Kinoshita, M.; Matsumori, N.; Oiki, S. Channel Formation and Membrane Deformation via Sterol-Aided Polymorphism of Amphidinol 3. Sci. Rep. 2017, 7, 10782. [Google Scholar] [CrossRef] [PubMed]
  447. Sung, W.S.; Lee, J.; Lee, D.G. Fungicidal effect of piscidin on Candida albicans: Pore formation in lipid vesicles and activity in fungal membranes. Biol. Pharm. Bull. 2008, 31, 1906–1910. [Google Scholar] [CrossRef]
  448. Gomes, I.P.; Santos, T.L.; de Souza, A.N.; Nunes, L.O.; Cardoso, G.A.; Matos, C.O.; Costa, L.M.F.; Lião, L.M.; Resende, J.M.; Verly, R.M. Membrane interactions of the anuran antimicrobial peptide HSP1-NH2: Different aspects of the association to anionic and zwitterionic biomimetic systems. Biochim. Biophys. Acta Biomembr. 2021, 1863, 183449. [Google Scholar] [CrossRef]
  449. Stock, S.D.; Hama, H.; Radding, J.A.; Young, D.A.; Takemoto, J.Y. Syringomycin E inhibition of Saccharomyces cerevisiae: Requirement for biosynthesis of sphingolipids with very-long-chain fatty acids and mannose- and phosphoinositol-containing head groups. Antimicrob. Agents Chemother. 2000, 44, 1174–1180. [Google Scholar] [CrossRef]
  450. Grilley, M.M.; Stock, S.D.; Dickson, R.C.; Lester, R.L.; Takemoto, J.Y. Syringomycin action gene SYR2 is essential for sphingolipid 4-hydroxylation in Saccharomyces cerevisiae. J. Biol. Chem. 1998, 273, 11062–11068. [Google Scholar] [CrossRef]
  451. Bento-Oliveira, A.; Santos, F.C.; Marquês, J.T.; Paulo, P.M.R.; Korte, T.; Herrmann, A.; Marinho, H.S.; de Almeida, R.F.M. Yeast Sphingolipid-Enriched Domains and Membrane Compartments in the Absence of Mannosyldiinositolphosphorylceramide. Biomolecules 2020, 10, 871. [Google Scholar] [CrossRef] [PubMed]
  452. Idkowiak-Baldys, J.; Grilley, M.M.; Takemoto, J.Y. Sphingolipid C4 hydroxylation influences properties of yeast detergent-insoluble glycolipid-enriched membranes. FEBS Lett. 2004, 569, 272–276. [Google Scholar] [CrossRef]
  453. Kaulin, Y.A.; Takemoto, J.Y.; Schagina, L.V.; Ostroumova, O.S.; Wangspa, R.; Teeter, J.H.; Brand, J.G. Sphingolipids influence the sensitivity of lipid bilayers to fungicide, syringomycin E. J. Bioenerg. Biomembr. 2005, 37, 339–348. [Google Scholar] [CrossRef]
  454. Efimova, S.S.; Zakharova, A.A.; Schagina, L.V.; Ostroumova, O.S. Two types of syringomycin E channels in sphingomyelin-containing bilayers. Eur. Biophys. J. 2016, 45, 91–98. [Google Scholar] [CrossRef]
  455. Mbongo, N.; Loiseau, P.M.; Billion, M.A.; Robert-Gero, M. Mechanism of amphotericin B resistance in Leishmania donovani promastigotes. Antimicrob. Agents Chemother. 1998, 42, 352–357. [Google Scholar] [CrossRef]
  456. Young, L.Y.; Hull, C.M.; Heitman, J. Disruption of ergosterol biosynthesis confers resistance to amphotericin B in Candida lusitaniae. Antimicrob. Agents Chemother. 2003, 47, 2717–2724. [Google Scholar] [CrossRef]
  457. Aloia, R.C.; Jensen, F.C.; Curtain, C.C.; Mobley, P.W.; Gordon, L.M. Lipid composition and fluidity of the human immunodeficiency virus. Proc. Natl. Acad. Sci. USA 1988, 85, 900–904. [Google Scholar] [CrossRef]
  458. Kalvodova, L.; Sampaio, J.L.; Cordo, S.; Ejsing, C.S.; Shevchenko, A.; Simons, K. The lipidomes of vesicular stomatitis virus, semliki forest virus, and the host plasma membrane analyzed by quantitative shotgun mass spectrometry. J. Virol. 2009, 83, 7996–8003. [Google Scholar] [CrossRef]
  459. Merz, A.; Long, G.; Hiet, M.S.; Brügger, B.; Chlanda, P.; Andre, P.; Wieland, F.; Krijnse-Locker, J.; Bartenschlager, R. Biochemical and morphological properties of hepatitis C virus particles and determination of their lipidome. J. Biol. Chem. 2011, 286, 3018–3032. [Google Scholar] [CrossRef]
  460. Gerl, M.J.; Sampaio, J.L.; Urban, S.; Kalvodova, L.; Verbavatz, J.M.; Binnington, B.; Lindemann, D.; Lingwood, C.A.; Shevchenko, A.; Schroeder, C.; et al. Quantitative analysis of the lipidomes of the influenza virus envelope and MDCK cell apical membrane. J. Cell Biol. 2012, 196, 213–221. [Google Scholar] [CrossRef]
  461. Martín-Acebes, M.A.; Merino-Ramos, T.; Blázquez, A.B.; Casas, J.; Escribano-Romero, E.; Sobrino, F.; Saiz, J.C. The composition of West Nile virus lipid envelope unveils a role of sphingolipid metabolism in flavivirus biogenesis. J. Virol. 2014, 88, 12041–12054. [Google Scholar] [CrossRef]
  462. Hofmann, S.; Krajewski, M.; Scherer, C.; Scholz, V.; Mordhorst, V.; Truschow, P.; Schöbel, A.; Reimer, R.; Schwudke, D.; Herker, E. Complex lipid metabolic remodeling is required for efficient hepatitis C virus replication. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 2018, 1863, 1041–1056. [Google Scholar] [CrossRef] [PubMed]
  463. Martín-Acebes, M.A.; Vázquez-Calvo, Á.; Saiz, J.C. Lipids and flaviviruses, present and future perspectives for the control of dengue, Zika, and West Nile viruses. Prog. Lipid Res. 2016, 64, 123–137. [Google Scholar] [CrossRef] [PubMed]
  464. Mariewskaya, K.A.; Tyurin, A.P.; Chistov, A.A.; Korshun, V.A.; Alferova, V.A.; Ustinov, A.V. Photosensitizing Antivirals. Molecules 2021, 26, 3971. [Google Scholar] [CrossRef]
  465. Mariewskaya, K.A.; Krasilnikov, M.S.; Korshun, V.A.; Ustinov, A.V.; Alferova, V.A. Near-Infrared Dyes: Towards Broad-Spectrum Antivirals. Int. J. Mol. Sci. 2022, 24, 188. [Google Scholar] [CrossRef] [PubMed]
  466. Schinazi, R.F.; Chu, C.K.; Babu, J.R.; Oswald, B.J.; Saalmann, V.; Cannon, D.L.; Eriksson, B.F.; Nasr, M. Anthraquinones as a new class of antiviral agents against human immunodeficiency virus. Antivir. Res. 1990, 13, 265–272. [Google Scholar] [CrossRef]
  467. Tang, J.; Colacino, J.M.; Larsen, S.H.; Spitzer, W. Virucidal activity of hypericin against enveloped and non-enveloped DNA and RNA viruses. Antivir. Res. 1990, 13, 313–325. [Google Scholar] [CrossRef]
  468. Kraus, G.A.; Pratt, D.; Tossberg, J.; Carpenter, S. Antiretroviral activity of synthetic hypericin and related analogs. Biochem. Biophys. Res. Commun. 1990, 172, 149–153. [Google Scholar] [CrossRef]
  469. Hudson, J.B.; Imperial, V.; Haugland, R.P.; Diwu, Z. Antiviral activities of photoactive perylenequinones. Photochem. Photobiol. 1997, 65, 352–354. [Google Scholar] [CrossRef]
  470. Andersen, D.O.; Weber, N.D.; Wood, S.G.; Hughes, B.G.; Murray, B.K.; North, J.A. In vitro virucidal activity of selected anthraquinones and anthraquinone derivatives. Antivir. Res. 1991, 16, 185–196. [Google Scholar] [CrossRef]
  471. Hudson, J.B.; Lopez-Bazzocchi, I.; Towers, G.H. Antiviral activities of hypericin. Antivir. Res. 1991, 15, 101–112. [Google Scholar] [CrossRef]
  472. Cohen, P.A.; Hudson, J.B.; Towers, G.H. Antiviral activities of anthraquinones, bianthrones and hypericin derivatives from lichens. Experientia 1996, 52, 180–183. [Google Scholar] [CrossRef]
  473. Hudson, J.B.; Delaey, E.; de Witte, P.A. Bromohypericins Are Potent Photoactive Antiviral Agents. Photochem. Photobiol. 1999, 70, 820–822. [Google Scholar] [CrossRef]
  474. Laille, M.; Gerald, F.; Debitus, C. In vitro antiviral activity on dengue virus of marine natural products. Cell. Mol. Life Sci. 1998, 54, 167–170. [Google Scholar] [CrossRef]
  475. Laurent, D.; Baumann, F.; Benoit, A.G.; Mortelecqe, A.; Nitatpattana, N.; Desvignes, I.; Debitus, C.; Laille, M.; Gonzalez, J.P.; Chungue, E. Structure-activity relationships of dengue antiviral polycyclic quinones. Southeast Asian J. Trop. Med. Public Health 2005, 36, 901–905. [Google Scholar]
  476. Hudson, J.B.; Zhou, J.; Chen, J.; Harris, L.; Yip, L.; Towers, G.H. Hypocrellin, from Hypocrella bambuase, is phototoxic to human immunodeficiency virus. Photochem. Photobiol. 1994, 60, 253–255. [Google Scholar] [CrossRef]
  477. Hirayama, J.; Ikebuchi, K.; Abe, H.; Kwon, K.W.; Ohnishi, Y.; Horiuchi, M.; Shinagawa, M.; Ikuta, K.; Kamo, N.; Sekiguchi, S. Photoinactivation of virus infectivity by hypocrellin A. Photochem. Photobiol. 1997, 66, 697–700. [Google Scholar] [CrossRef]
  478. Sun, Y.; Chen, Y.L.; Xu, C.P.; Gao, J.; Feng, Y.; Wu, Q.F. Disinfection of influenza a viruses by Hypocrellin a-mediated photodynamic inactivation. Photodiagnosis Photodyn. Ther. 2023, 43, 103674. [Google Scholar] [CrossRef]
  479. St Vincent, M.R.; Colpitts, C.C.; Ustinov, A.V.; Muqadas, M.; Joyce, M.A.; Barsby, N.L.; Epand, R.F.; Epand, R.M.; Khramyshev, S.A.; Valueva, O.A.; et al. Rigid amphipathic fusion inhibitors, small molecule antiviral compounds against enveloped viruses. Proc. Natl. Acad. Sci. USA 2010, 107, 17339–17344. [Google Scholar] [CrossRef]
  480. Colpitts, C.C.; Ustinov, A.V.; Epand, R.F.; Epand, R.M.; Korshun, V.A.; Schang, L.M. 5-(Perylen-3-yl)ethynyl-arabino-uridine (aUY11), an arabino-based rigid amphipathic fusion inhibitor, targets virion envelope lipids to inhibit fusion of influenza virus, hepatitis C virus, and other enveloped viruses. J. Virol. 2013, 87, 3640–3654. [Google Scholar] [CrossRef]
  481. Speerstra, S.; Chistov, A.A.; Proskurin, G.V.; Aralov, A.V.; Ulashchik, E.A.; Streshnev, P.P.; Shmanai, V.V.; Korshun, V.A.; Schang, L.M. Antivirals acting on viral envelopes via biophysical mechanisms of action. Antivir. Res. 2018, 149, 164–173. [Google Scholar] [CrossRef] [PubMed]
  482. Vigant, F.; Hollmann, A.; Lee, J.; Santos, N.C.; Jung, M.E.; Lee, B. The rigid amphipathic fusion inhibitor dUY11 acts through photosensitization of viruses. J. Virol. 2014, 88, 1849–1853. [Google Scholar] [CrossRef]
  483. Orlov, A.A.; Chistov, A.A.; Kozlovskaya, L.I.; Ustinov, A.V.; Korshun, V.A.; Karganova, G.G.; Osolodkin, D.I. Rigid amphipathic nucleosides suppress reproduction of the tick-borne encephalitis virus. Med. Chem. Commun. 2016, 7, 495–499. [Google Scholar] [CrossRef]
  484. Chistov, A.A.; Chumakov, S.P.; Mikhnovets, I.E.; Nikitin, T.D.; Slesarchuk, N.A.; Uvarova, V.I.; Rubekina, A.A.; Nikolaeva, Y.V.; Radchenko, E.V.; Khvatov, E.V.; et al. 5-(Perylen-3-ylethynyl)uracil as an antiviral scaffold: Potent suppression of enveloped virus reproduction by 3-methyl derivatives In vitro. Antivir. Res. 2023, 209, 105508. [Google Scholar] [CrossRef]
  485. Hakobyan, A.; Galindo, I.; Nañez, A.; Arabyan, E.; Karalyan, Z.; Chistov, A.A.; Streshnev, P.P.; Korshun, V.A.; Alonso, C.; Zakaryan, H. Rigid amphipathic fusion inhibitors demonstrate antiviral activity against African swine fever virus. J. Gen. Virol. 2018, 99, 148–156. [Google Scholar] [CrossRef] [PubMed]
  486. Wiehe, A.; O’Brien, J.M.; Senge, M.O. Trends and targets in antiviral phototherapy. Photochem. Photobiol. Sci. 2019, 18, 2565–2612. [Google Scholar] [CrossRef] [PubMed]
  487. Straková, P.; Bednář, P.; Kotouček, J.; Holoubek, J.; Fořtová, A.; Svoboda, P.; Štefánik, M.; Huvarová, I.; Šimečková, P.; Mašek, J.; et al. Antiviral activity of singlet oxygen-photogenerating perylene compounds against SARS-CoV-2: Interaction with the viral envelope and photodynamic virion inactivation. Virus Res. 2023, 334, 199158. [Google Scholar] [CrossRef]
  488. Chistov, A.A.; Orlov, A.A.; Streshnev, P.P.; Slesarchuk, N.A.; Aparin, I.O.; Rathi, B.; Brylev, V.A.; Kutyakov, S.V.; Mikhura, I.V.; Ustinov, A.V.; et al. Compounds based on 5-(perylen-3-ylethynyl)uracil scaffold: High activity against tick-borne encephalitis virus and non-specific activity against enterovirus A. Eur. J. Med. Chem. 2019, 171, 93–103. [Google Scholar] [CrossRef]
  489. Mariewskaya, K.A.; Gvozdev, D.A.; Chistov, A.A.; Straková, P.; Huvarová, I.; Svoboda, P.; Kotouček, J.; Ivanov, N.M.; Krasilnikov, M.S.; Zhitlov, M.Y.; et al. Membrane-Targeting Perylenylethynylphenols Inactivate Medically Important Coronaviruses via the Singlet Oxygen Photogeneration Mechanism. Molecules 2023, 28, 6278. [Google Scholar] [CrossRef]
  490. Carpenter, B.L.; Situ, X.; Scholle, F.; Bartelmess, J.; Weare, W.W.; Ghiladi, R.A. Antiviral, Antifungal and Antibacterial Activities of a BODIPY-Based Photosensitizer. Molecules 2015, 20, 10604–10621. [Google Scholar] [CrossRef]
  491. Wolf, M.C.; Freiberg, A.N.; Zhang, T.; Akyol-Ataman, Z.; Grock, A.; Hong, P.W.; Li, J.; Watson, N.F.; Fang, A.Q.; Aguilar, H.C.; et al. A broad-spectrum antiviral targeting entry of enveloped viruses. Proc. Natl. Acad. Sci. USA 2010, 107, 3157–3162. [Google Scholar] [CrossRef]
  492. Vigant, F.; Lee, J.; Hollmann, A.; Tanner, L.B.; Akyol Ataman, Z.; Yun, T.; Shui, G.; Aguilar, H.C.; Zhang, D.; Meriwether, D.; et al. A mechanistic paradigm for broad-spectrum antivirals that target virus-cell fusion. PLoS Pathog. 2013, 9, e1003297. [Google Scholar] [CrossRef]
  493. Hollmann, A.; Gonçalves, S.; Augusto, M.T.; Castanho, M.A.; Lee, B.; Santos, N.C. Effects of singlet oxygen generated by a broad-spectrum viral fusion inhibitor on membrane nanoarchitecture. Nanomedicine 2015, 11, 1163–1167. [Google Scholar] [CrossRef]
  494. Arnaut, Z.A.; Pinto, S.M.A.; Aroso, R.T.; Amorim, A.S.; Lobo, C.S.; Schaberle, F.A.; Pereira, D.; Núñez, J.; Nunes, S.C.C.; Pais, A.A.C.C.; et al. Selective, broad-spectrum antiviral photodynamic disinfection with dicationic imidazolyl chlorin photosensitizers. Photochem. Photobiol. Sci. 2023, 22, 2607–2620. [Google Scholar] [CrossRef] [PubMed]
  495. Jurak, I.; Cokarić Brdovčak, M.; Djaković, L.; Bertović, I.; Knežević, K.; Lončarić, M.; Jurak Begonja, A.; Malatesti, N. Photodynamic Inhibition of Herpes Simplex Virus 1 Infection by Tricationic Amphiphilic Porphyrin with a Long Alkyl Chain. Pharmaceutics 2023, 15, 956. [Google Scholar] [CrossRef]
  496. Monjo, A.L.; Pringle, E.S.; Thornbury, M.; Duguay, B.A.; Monro, S.M.A.; Hetu, M.; Knight, D.; Cameron, C.G.; McFarland, S.A.; McCormick, C. Photodynamic Inactivation of Herpes Simplex Viruses. Viruses 2018, 10, 532. [Google Scholar] [CrossRef] [PubMed]
  497. Meunier, T.; Desmarets, L.; Bordage, S.; Bamba, M.; Hervouet, K.; Rouillé, Y.; François, N.; Decossas, M.; Sencio, V.; Trottein, F.; et al. A Photoactivable Natural Product with Broad Antiviral Activity against Enveloped Viruses, Including Highly Pathogenic Coronaviruses. Antimicrob. Agents Chemother. 2022, 66, e0158121. [Google Scholar] [CrossRef] [PubMed]
  498. Nikolayeva, Y.V.; Ulashchik, E.A.; Chekerda, E.V.; Galochkina, A.V.; Slesarchuk, N.A.; Chistov, A.A.; Nikitin, T.D.; Korshun, V.A.; Shmanai, V.V.; Ustinov, A.V.; et al. 5-(Perylen-3-ylethynyl)uracil Derivatives Inhibit Reproduction of Respiratory Viruses. Russ. J. Bioorg. Chem. 2020, 46, 315–320. [Google Scholar] [CrossRef] [PubMed]
  499. Proskurin, G.V.; Orlov, A.A.; Brylev, V.A.; Kozlovskaya, L.I.; Chistov, A.A.; Karganova, G.G.; Palyulin, V.A.; Osolodkin, D.I.; Korshun, V.A.; Aralov, A.V. 3′-O-Substituted 5-(perylen-3-ylethynyl)-2′-deoxyuridines as tick-borne encephalitis virus reproduction inhibitors. Eur. J. Med. Chem. 2018, 155, 77–83. [Google Scholar] [CrossRef]
  500. Weil, T.; Groß, R.; Röcker, A.; Bravo-Rodriguez, K.; Heid, C.; Sowislok, A.; Le, M.H.; Erwin, N.; Dwivedi, M.; Bart, S.M.; et al. Supramolecular Mechanism of Viral Envelope Disruption by Molecular Tweezers. J. Am. Chem. Soc. 2020, 142, 17024–17038. [Google Scholar] [CrossRef]
  501. Lump, E.; Castellano, L.M.; Meier, C.; Seeliger, J.; Erwin, N.; Sperlich, B.; Stürzel, C.M.; Usmani, S.; Hammond, R.M.; von Einem, J.; et al. A molecular tweezer antagonizes seminal amyloids and HIV infection. Elife 2015, 4, e05397. [Google Scholar] [CrossRef]
  502. Röcker, A.E.; Müller, J.A.; Dietzel, E.; Harms, M.; Krüger, F.; Heid, C.; Sowislok, A.; Riber, C.F.; Kupke, A.; Lippold, S.; et al. The molecular tweezer CLR01 inhibits Ebola and Zika virus infection. Antivir. Res. 2018, 152, 26–35. [Google Scholar] [CrossRef] [PubMed]
  503. Weil, T.; Kirupakaran, A.; Le, M.H.; Rebmann, P.; Mieres-Perez, J.; Issmail, L.; Conzelmann, C.; Müller, J.A.; Rauch, L.; Gilg, A.; et al. Advanced Molecular Tweezers with Lipid Anchors against SARS-CoV-2 and Other Respiratory Viruses. JACS Au 2022, 2, 2187–2202. [Google Scholar] [CrossRef]
  504. Wang, G.; Watson, K.M.; Buckheit, R.W., Jr. Anti-human immunodeficiency virus type 1 activities of antimicrobial peptides derived from human and bovine cathelicidins. Antimicrob. Agents Chemother. 2008, 52, 3438–3440. [Google Scholar] [CrossRef]
  505. Kagan, B.L.; Ganz, T.; Lehrer, R.I. Defensins: A family of antimicrobial and cytotoxic peptides. Toxicology 1994, 87, 131–149. [Google Scholar] [CrossRef] [PubMed]
  506. Liu, R.; Liu, Z.; Peng, H.; Lv, Y.; Feng, Y.; Kang, J.; Lu, N.; Ma, R.; Hou, S.; Sun, W.; et al. Bomidin: An Optimized Antimicrobial Peptide with Broad Antiviral Activity Against Enveloped Viruses. Front. Immunol. 2022, 13, 851642. [Google Scholar] [CrossRef] [PubMed]
  507. Omer, A.A.M.; Hinkula, J.; Tran, P.T.; Melik, W.; Zattarin, E.; Aili, D.; Selegård, R.; Bengtsson, T.; Khalaf, H. Plantaricin NC8 αβ rapidly and efficiently inhibits flaviviruses and SARS-CoV-2 by disrupting their envelopes. PLoS ONE 2022, 17, e0278419. [Google Scholar] [CrossRef] [PubMed]
  508. Henriques, S.T.; Huang, Y.H.; Rosengren, K.J.; Franquelim, H.G.; Carvalho, F.A.; Johnson, A.; Sonza, S.; Tachedjian, G.; Castanho, M.A.; Daly, N.L.; et al. Decoding the membrane activity of the cyclotide kalata B1: The importance of phosphatidylethanolamine phospholipids and lipid organization on hemolytic and anti-HIV activities. J. Biol. Chem. 2011, 286, 24231–24241. [Google Scholar] [CrossRef]
  509. Kozlov, M.M.; Leikin, S.L.; Chernomordik, L.V.; Markin, V.S.; Chizmadzhev, Y.A. Stalk mechanism of vesicle fusion. Intermixing of aqueous contents. Eur. Biophys. J. 1989, 17, 121–129. [Google Scholar] [CrossRef]
  510. Cooke, I.R.; Deserno, M. Coupling between lipid shape and membrane curvature. Biophys. J. 2006, 91, 487–495. [Google Scholar] [CrossRef]
  511. Miao, L.; Stafford, A.; Nir, S.; Turco, S.J.; Flanagan, T.D.; Epand, R.M. Potent inhibition of viral fusion by the lipophosphoglycan of Leishmania donovani. Biochemistry 1995, 34, 4676–4683. [Google Scholar] [CrossRef] [PubMed]
  512. Rasmusson, B.J.; Flanagan, T.D.; Turco, S.J.; Epand, R.M.; Petersen, N.O. Fusion of Sendai virus and individual host cells and inhibition of fusion by lipophosphoglycan measured with image correlation spectroscopy. Biochim. Biophys. Acta 1998, 1404, 338–352. [Google Scholar] [CrossRef] [PubMed]
  513. Sardar, A.; Lahiri, A.; Kamble, M.; Mallick, A.I.; Tarafdar, P.K. Translation of Mycobacterium Survival Strategy to Develop a Lipo-peptide based Fusion Inhibitor. Angew. Chem. Int. Ed. Engl. 2021, 60, 6101–6106. [Google Scholar] [CrossRef] [PubMed]
  514. Yuan, L.; Zhang, S.; Wang, Y.; Li, Y.; Wang, X.; Yang, Q. Surfactin Inhibits Membrane Fusion during Invasion of Epithelial Cells by Enveloped Viruses. J. Virol. 2018, 92, e00809-18. [Google Scholar] [CrossRef]
  515. Kracht, M.; Rokos, H.; Ozel, M.; Kowall, M.; Pauli, G.; Vater, J. Antiviral and hemolytic activities of surfactin isoforms and their methyl ester derivatives. J. Antibiot. 1999, 52, 613–619. [Google Scholar] [CrossRef]
  516. Shekunov, E.V.; Zlodeeva, P.D.; Efimova, S.S.; Muryleva, A.A.; Zarubaev, V.V.; Slita, A.V.; Ostroumova, O.S. Cyclic lipopeptides as membrane fusion inhibitors against SARS-CoV-2: New tricks for old dogs. Antivir. Res. 2023, 212, 105575. [Google Scholar] [CrossRef]
  517. Shekunov, E.V.; Efimova, S.S.; Yudintceva, N.M.; Muryleva, A.A.; Zarubaev, V.V.; Slita, A.V.; Ostroumova, O.S. Plant Alkaloids Inhibit Membrane Fusion Mediated by Calcium and Fragments of MERS-CoV and SARS-CoV/SARS-CoV-2 Fusion Peptides. Biomedicines 2021, 9, 1434. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the synthesis of fatty acids of bacterial membrane lipids. The red, blue, and purple ellipses indicate that the enzyme is produced by Gram-positive, Gram-negative, or both Gram-positive and Gram-negative bacteria, respectively. Some examples of enzyme inhibitors are shown in the black box. Abbreviations: AccC—biotin carboxylase; AccB—biotin carboxyl carrier protein; AccAD—biotin carboxyl transferase; FabD—malonyl-CoA:acyl carrier protein (ACP) transacylase; FabH, FabF, and FabB—β-ketoacyl-ACP synthase III (KAS III), II (KAS II), and I (KAS I), respectively; mtFabH—FabH homolog of Mycobacterium tuberculosis; FabG—NADPH-dependent β-ketoacyl-ACP reductase; FabZ—β-hydroxyacyl-ACP dehydratase; FabA/FabQ/FabN—bifunctional β-hydroxyacyl-ACP dehydratase/trans-2-,cis-3-decenoyl-ACP isomerase; FabMtrans-2-,cis-3-decenoyl-ACP isomerase; FabI/FabK/FabL/FabVtrans-2-enoyl-ACP reductase; PlsX—phosphate acyltransferase; PlsY—acyl-phosphate:glycerol-3-phosphate acyltransferase; PlsB—glycerol-3-phosphate acyltransferase; PlsC—1-acyl-sn-glycerol-3-phosphate acyltransferase; BCFA—branched-chain fatty acids; LCFA—long-chain fatty acids; LPA—lysophosphatidic acid; PA—phosphatidic acid; UFA—unsaturated fatty acids; SFA—saturated fatty acids; CFA—cyclopropane fatty acids; R, R′, R1, and R2—fatty acid hydrocarbon radicals.
Figure 1. Schematic representation of the synthesis of fatty acids of bacterial membrane lipids. The red, blue, and purple ellipses indicate that the enzyme is produced by Gram-positive, Gram-negative, or both Gram-positive and Gram-negative bacteria, respectively. Some examples of enzyme inhibitors are shown in the black box. Abbreviations: AccC—biotin carboxylase; AccB—biotin carboxyl carrier protein; AccAD—biotin carboxyl transferase; FabD—malonyl-CoA:acyl carrier protein (ACP) transacylase; FabH, FabF, and FabB—β-ketoacyl-ACP synthase III (KAS III), II (KAS II), and I (KAS I), respectively; mtFabH—FabH homolog of Mycobacterium tuberculosis; FabG—NADPH-dependent β-ketoacyl-ACP reductase; FabZ—β-hydroxyacyl-ACP dehydratase; FabA/FabQ/FabN—bifunctional β-hydroxyacyl-ACP dehydratase/trans-2-,cis-3-decenoyl-ACP isomerase; FabMtrans-2-,cis-3-decenoyl-ACP isomerase; FabI/FabK/FabL/FabVtrans-2-enoyl-ACP reductase; PlsX—phosphate acyltransferase; PlsY—acyl-phosphate:glycerol-3-phosphate acyltransferase; PlsB—glycerol-3-phosphate acyltransferase; PlsC—1-acyl-sn-glycerol-3-phosphate acyltransferase; BCFA—branched-chain fatty acids; LCFA—long-chain fatty acids; LPA—lysophosphatidic acid; PA—phosphatidic acid; UFA—unsaturated fatty acids; SFA—saturated fatty acids; CFA—cyclopropane fatty acids; R, R′, R1, and R2—fatty acid hydrocarbon radicals.
Antibiotics 12 01716 g001
Figure 2. Schematic representation of the synthesis of “heads” of bacterial membrane lipids. The red, blue, and purple ellipses indicate that the enzyme is produced by Gram-positive, Gram-negative, or both Gram-positive and Gram-negative bacteria, respectively. Some examples of enzyme inhibitors are shown in a black box. Abbreviations: CdsA—cytidine diphosphate-diacylglycerol synthase; PgsA—phosphatidylglycerophosphate synthase; PgpA, PgpB, and PgpC—phosphatidylglycerolphosphate phosphatases; ClsA, ClsB, ClsC, Cls1, and Cls2—cardiolipin synthases; PssA—phosphatidylserine synthase; Psd—phosphatidylserine decarboxylase; Pcs—phosphatidylcholine synthase; PIS—phosphatidylinositol synthase; MprF—lysil phosphatidylglycerol synthase and flippase (multiple peptide resistance factor); MdoB—phosphoglycerol transferase; DgkA and DgkB—diacylglycerol kinases; YpfP—diacylglycerol β-glucosyltransferase; PA—phosphatidic acid; CDP-DG—CDP-diacylglycerol; PS—phosphatidylserine; PE—phosphatidylethanolamine; PC—phosphatidylcholine; PI—phosphatidylinositol; PGP—phosphatidylglycerol phosphate; PG—phosphatidylglycerol; CL—cardiolipin; LPG—lysil phosphatidylglycerol; DG—diacylglycerol; MGDG—monoglycosyl-DG; DGDG—diglycosyl-DG; R1 and R2—fatty acid hydrocarbon radicals.
Figure 2. Schematic representation of the synthesis of “heads” of bacterial membrane lipids. The red, blue, and purple ellipses indicate that the enzyme is produced by Gram-positive, Gram-negative, or both Gram-positive and Gram-negative bacteria, respectively. Some examples of enzyme inhibitors are shown in a black box. Abbreviations: CdsA—cytidine diphosphate-diacylglycerol synthase; PgsA—phosphatidylglycerophosphate synthase; PgpA, PgpB, and PgpC—phosphatidylglycerolphosphate phosphatases; ClsA, ClsB, ClsC, Cls1, and Cls2—cardiolipin synthases; PssA—phosphatidylserine synthase; Psd—phosphatidylserine decarboxylase; Pcs—phosphatidylcholine synthase; PIS—phosphatidylinositol synthase; MprF—lysil phosphatidylglycerol synthase and flippase (multiple peptide resistance factor); MdoB—phosphoglycerol transferase; DgkA and DgkB—diacylglycerol kinases; YpfP—diacylglycerol β-glucosyltransferase; PA—phosphatidic acid; CDP-DG—CDP-diacylglycerol; PS—phosphatidylserine; PE—phosphatidylethanolamine; PC—phosphatidylcholine; PI—phosphatidylinositol; PGP—phosphatidylglycerol phosphate; PG—phosphatidylglycerol; CL—cardiolipin; LPG—lysil phosphatidylglycerol; DG—diacylglycerol; MGDG—monoglycosyl-DG; DGDG—diglycosyl-DG; R1 and R2—fatty acid hydrocarbon radicals.
Antibiotics 12 01716 g002
Figure 3. Schematic representation of the biosynthesis of lipid A. The blue ellipses indicate that all enzymes are only produced by Gram-negative bacteria. Some examples of enzyme inhibitors are shown in a black box. Abbreviations: LpxA—UDP-N-acetylglucosamine acyltransferase; LpxC—UDP-3-O-(R-3-hydroxyacyl)-N-acetylglucosamine deacetylase; LpxD—UDP-3-O-(R-3-hydroxyacyl)glucosamine N-acyltransferase; LpxH, LpxI, and LpxG—UDP-diacylglucosamine pyrophosphohydrolases; LpxB—lipid-A-disaccharide synthase; LpxK—tetraacyldisaccharide-1-phosphate 4′-kinase; WaaA—3-deoxy-D-manno-oct-2-ulosonic acid (Kdo) transferase; LpxL, LpxM, and LpxP—lysophospholipid acyltransferases.
Figure 3. Schematic representation of the biosynthesis of lipid A. The blue ellipses indicate that all enzymes are only produced by Gram-negative bacteria. Some examples of enzyme inhibitors are shown in a black box. Abbreviations: LpxA—UDP-N-acetylglucosamine acyltransferase; LpxC—UDP-3-O-(R-3-hydroxyacyl)-N-acetylglucosamine deacetylase; LpxD—UDP-3-O-(R-3-hydroxyacyl)glucosamine N-acyltransferase; LpxH, LpxI, and LpxG—UDP-diacylglucosamine pyrophosphohydrolases; LpxB—lipid-A-disaccharide synthase; LpxK—tetraacyldisaccharide-1-phosphate 4′-kinase; WaaA—3-deoxy-D-manno-oct-2-ulosonic acid (Kdo) transferase; LpxL, LpxM, and LpxP—lysophospholipid acyltransferases.
Antibiotics 12 01716 g003
Figure 4. Schematic representation of major mechanisms of antimicrobial action via pore formation (described by two distinct models: barrel-stave channel and toroidal pore-containing lipids) and membrane disruption. Some examples of antibacterials and antifungals directly targeting lipid membranes by two different mechanisms are shown in black boxes.
Figure 4. Schematic representation of major mechanisms of antimicrobial action via pore formation (described by two distinct models: barrel-stave channel and toroidal pore-containing lipids) and membrane disruption. Some examples of antibacterials and antifungals directly targeting lipid membranes by two different mechanisms are shown in black boxes.
Antibiotics 12 01716 g004
Figure 5. The PC biosynthesis in S. cerevisiae. The de novo and Kennedy pathways are represented by the grey and violet lines, respectively. The enzymes of the indicated pathways are highlighted with green and light-green ellipses, respectively. Abbreviations: Pems—phosphatidylethanolamine N-methyltransferase; EK—ethanolamine kinase; CK—choline kinase; ECT—phosphoethanolamine cytidylyltransferase; CCT—phosphocholine cytidylyltransferase; EPT—ethanolaminephosphotransferase; CPT—cholinephosphotransferase; PMME—phosphatidyl-N-monomethylethanolamine; PDME—phosphatidyl-N,N-dimethylethanolamine.
Figure 5. The PC biosynthesis in S. cerevisiae. The de novo and Kennedy pathways are represented by the grey and violet lines, respectively. The enzymes of the indicated pathways are highlighted with green and light-green ellipses, respectively. Abbreviations: Pems—phosphatidylethanolamine N-methyltransferase; EK—ethanolamine kinase; CK—choline kinase; ECT—phosphoethanolamine cytidylyltransferase; CCT—phosphocholine cytidylyltransferase; EPT—ethanolaminephosphotransferase; CPT—cholinephosphotransferase; PMME—phosphatidyl-N-monomethylethanolamine; PDME—phosphatidyl-N,N-dimethylethanolamine.
Antibiotics 12 01716 g005
Figure 6. Schematic representation of the sphingolipid biosynthetic pathways in S. cerevisiae. The enzymes are highlighted with orange ellipses. Some examples of enzyme inhibitors are shown in black boxes. Abbreviations: SPT—serine palmitoyltransferase; KDSR—3-ketodihydrosphingosine reductase; CerS—ceramide synthase; SCH—sphingosine C4-hydroxylase; IPCS—inositol-phosphoceramide synthase; MIPCS—mannosylinositol phosphorylceramide synthase; IPS—inositolphosphotransferase; IPC—inositol-phosphoceramide; MIPC—mannose-inositol-phosphoceramide; M(IP)2C—mannose-(inositol-P)2-ceramide.
Figure 6. Schematic representation of the sphingolipid biosynthetic pathways in S. cerevisiae. The enzymes are highlighted with orange ellipses. Some examples of enzyme inhibitors are shown in black boxes. Abbreviations: SPT—serine palmitoyltransferase; KDSR—3-ketodihydrosphingosine reductase; CerS—ceramide synthase; SCH—sphingosine C4-hydroxylase; IPCS—inositol-phosphoceramide synthase; MIPCS—mannosylinositol phosphorylceramide synthase; IPS—inositolphosphotransferase; IPC—inositol-phosphoceramide; MIPC—mannose-inositol-phosphoceramide; M(IP)2C—mannose-(inositol-P)2-ceramide.
Antibiotics 12 01716 g006
Figure 7. Schematic representation of the ergosterol biosynthetic pathway in S. cerevisiae. The enzymes are highlighted with pink ellipses. The branch for the cholesterol biosynthesis in mammalian cells is marked with an orange color. The branch for the biosynthesis of phytosterols (campesterol, β-sitosterol, and stigmasterol) is marked with a green color. Some examples of enzyme inhibitors are shown in the black box. Abbreviations: ERG10—acetyl-CoA C-acetyltransferase; ERG13—hydroxymethylglutaryl-CoA synthase; HMG1/2—hydroxymethylglutaryl-CoA reductase; ERG12—mevalonate kinase; ERG8—phosphomevalonate kinase; ERG19—diphosphomevalonate decarboxylase; ERG20—farnesyl diphosphate synthase; ERG9—squalene synthase; ERG1—squalene epoxidase; ERG7—2,3-oxidosqualene cyclase; ERG11—lanosterol 14α-demethylase; ERG24—sterol C14-reductase; ERG25/26—sterol C4-methyloxidase; ERG6—sterol C24-methyltransferase; ERG2—sterol C8,7-isomerase; ERG3—sterol C5(6)-desaturase; ERG5—sterol C22-desaturase; ERG4—sterol C24-reductase.
Figure 7. Schematic representation of the ergosterol biosynthetic pathway in S. cerevisiae. The enzymes are highlighted with pink ellipses. The branch for the cholesterol biosynthesis in mammalian cells is marked with an orange color. The branch for the biosynthesis of phytosterols (campesterol, β-sitosterol, and stigmasterol) is marked with a green color. Some examples of enzyme inhibitors are shown in the black box. Abbreviations: ERG10—acetyl-CoA C-acetyltransferase; ERG13—hydroxymethylglutaryl-CoA synthase; HMG1/2—hydroxymethylglutaryl-CoA reductase; ERG12—mevalonate kinase; ERG8—phosphomevalonate kinase; ERG19—diphosphomevalonate decarboxylase; ERG20—farnesyl diphosphate synthase; ERG9—squalene synthase; ERG1—squalene epoxidase; ERG7—2,3-oxidosqualene cyclase; ERG11—lanosterol 14α-demethylase; ERG24—sterol C14-reductase; ERG25/26—sterol C4-methyloxidase; ERG6—sterol C24-methyltransferase; ERG2—sterol C8,7-isomerase; ERG3—sterol C5(6)-desaturase; ERG5—sterol C22-desaturase; ERG4—sterol C24-reductase.
Antibiotics 12 01716 g007
Figure 8. Modes of action of lipid-targeting antivirals: photosensitizers (A); molecular tweezers (B); curvature stress-induced fusion inhibitors (C).
Figure 8. Modes of action of lipid-targeting antivirals: photosensitizers (A); molecular tweezers (B); curvature stress-induced fusion inhibitors (C).
Antibiotics 12 01716 g008
Table 1. Major inhibitors of bacterial FASII.
Table 1. Major inhibitors of bacterial FASII.
InhibitorStructureEnzymeOriginIC50, μMReferences
amino-oxazole dibenzylamideAntibiotics 12 01716 i001AccCE. coli0.125[3]
(R)-2-(2-chlorobenzylamino)-1-(2,3-dihydro-1H-inden-1-yl)-1H-imidazo[4,5-b]pyridine-5-carboxamideAntibiotics 12 01716 i002AccCE. coli0.02[4]
moiramide BAntibiotics 12 01716 i003AccADS. aureus0.096[5]
E. coli0.006[5]
andrimidAntibiotics 12 01716 i004AccADS. aureus0.091[5]
E. coli0.004[5]
thiolactomycinAntibiotics 12 01716 i005FabHS. pneumoniae7.9 ± 1.1[6]
H. influenzae5.8 ± 1.6[6]
M. tuberculosis24[7]
E. coli32–110[6,8]
FabFE. coli6[8]
FabBE. coli2–25[8,9]
SB418011Antibiotics 12 01716 i006FabHS. pneumoniae0.016 ± 0.003[6]
H. influenzae0.59 ± 0.05[6]
E. coli1.20 ± 0.40[6]
ceruleninAntibiotics 12 01716 i007FabFE. coli20[8]
FabBE. coli3[8]
platensimycinAntibiotics 12 01716 i008FabFS. aureus0.02–0.29[10,11]
E. coli0.02[12]
platencinAntibiotics 12 01716 i009FabHS. aureus9.2–16.2[10,11]
FabFS. aureus0.1–4.6[10,11]
(-)-epigallocatechin gallateAntibiotics 12 01716 i010FabGE. coli5[13]
P. falciparum0.3[14,15]
FabIE. coli15[13]
P. falciparum0.2[14]
FabZP. falciparum0.03–0.4[14,15]
(-)-gallocatechin gallateAntibiotics 12 01716 i011FabGE. coli10[13]
P. falciparum1.1[14]
FabIE. coli5[13]
P. falciparum0.5[14]
FabZP. falciparum0.6[14]
(-)-epicatechin gallateAntibiotics 12 01716 i012FabGE. coli15[13]
P. falciparum1[14]
FabIE. coli10[13]
P. falciparum0.2[14]
FabZP. falciparum0.4[14]
(-)-catechin gallateAntibiotics 12 01716 i013FabGE. coli10[13]
P. falciparum1[14]
FabIE. coli5[13]
P. falciparum0.3[14]
FabZP. falciparum0.4[14]
buteinAntibiotics 12 01716 i014FabGE. coli10[13]
FabIE. coli30[13]
isoliquiritigeninAntibiotics 12 01716 i015FabGE. coli20[13]
FabIE. coli40[13]
2,2′,4′-trihydroxychalconeAntibiotics 12 01716 i016FabGE. coli25[13]
FabIE. coli40[13]
fisetinAntibiotics 12 01716 i017FabGE. coli30[13]
P. falciparum4.1[14]
FabIE. coli50[13]
P. falciparum1[14]
FabZP. falciparum2[14]
quercetinAntibiotics 12 01716 i018FabGE. coli20[13]
P. falciparum5.4[14]
FabIE. coli20[13]
P. falciparum1.5[14]
FabZP. falciparum1.5[14]
resveratrolAntibiotics 12 01716 i019FabGE. coli65[13]
FabIE. coli30[13]
piceatannolAntibiotics 12 01716 i020FabGE. coli35[13]
FabIE. coli15[13]
fustinAntibiotics 12 01716 i021FabGE. coli25[13]
FabIE. coli40[13]
taxifolinAntibiotics 12 01716 i022FabGE. coli20[13]
FabIE. coli30[13]
kaempferolAntibiotics 12 01716 i023FabGP. falciparum4[14]
FabIP. falciparum20[14]
luteolinAntibiotics 12 01716 i024FabGP. falciparum4[14]
FabIP. falciparum2[14]
FabZP. falciparum5[14]
luteolin 7-O-β-D-glucopyranosideAntibiotics 12 01716 i025FabIP. falciparum22[16]
myricetinAntibiotics 12 01716 i026FabGP. falciparum14[14]
FabIP. falciparum0.4[14]
FabZP. falciparum2[14]
isorhamnetinAntibiotics 12 01716 i027FabGP. falciparum8.3[14]
FabIP. falciparum5[14]
7,3′,4′-trihydroxyisoflavoneAntibiotics 12 01716 i028FabGE. coli35[13]
FabIE. coli25[13]
morinAntibiotics 12 01716 i029FabGP. falciparum2.3[14]
FabIP. falciparum5[14]
FabZP. falciparum8[14]
macrolactin SAntibiotics 12 01716 i030FabGS. aureus130[17]
macrolactin BAntibiotics 12 01716 i031FabGS. aureus100[17]
NAS-21Antibiotics 12 01716 i032FabZM. smegmatis360[18]
NAS-91Antibiotics 12 01716 i033FabZM. smegmatis498[18]
emodinAntibiotics 12 01716 i034FabZF. tularensis43.1 ± 9.2[19]
Y. pestis29.7 ± 6.0[19]
H. pylori9.70 ± 1.0[20]
mangostinAntibiotics 12 01716 i035FabZF. tularensis7.7 ± 2.0[19]
Y. pestis6.1 ± 1.4[19]
stictic acidAntibiotics 12 01716 i036FabZF. tularensis27.8 ± 6.1[19]
Y. pestis13.0 ± 1.4[19]
1,4-naphthoquinoneAntibiotics 12 01716 i037FabDM. catarrhalis23.18 ± 2.48[21]
FabZM. catarrhalis26.67 ± 3.34[21]
jugloneAntibiotics 12 01716 i038FabDH. pylori20 ± 1[22]
FabZF. tularensis5.4 ± 1.4[19]
Y. pestis5.3 ± 1.0[19]
H. pylori30 ± 4[22]
triclosanAntibiotics 12 01716 i039FabIE. coli0.98[23]
S. aureus0.44–0.66[23,24,25]
P. falciparum0.05–2[15,26]
C. trachomatis0.32 ± 0.08[27]
AFN-1252Antibiotics 12 01716 i040FabIC. trachomatis0.95 ± 0.21[27]
xanthorrhizolAntibiotics 12 01716 i041FabIE. coli17.1 ± 1.8[28]
complestatinAntibiotics 12 01716 i042FabIS. aureus0.5[25]
FabKS. pneumoniae10[25]
neuroprotectin AAntibiotics 12 01716 i043FabIS. aureus0.3[25]
chloropeptin IAntibiotics 12 01716 i044FabIS. aureus0.6[25]
meleagrinAntibiotics 12 01716 i045FabIS. aureus40.1[23]
E. coli33.2[23]
phellinstatinAntibiotics 12 01716 i046FabIS. aureus6[29]
chalcomoracinAntibiotics 12 01716 i047FabIS. aureus5.5[30]
moracin CAntibiotics 12 01716 i048FabIS. aureus83.8[30]
aquastatin AAntibiotics 12 01716 i049FabIS. aureus3.2[31]
FabKS. pneumoniae9.2[31]
atromentinAntibiotics 12 01716 i050FabKS. pneumoniae0.24[32]
leucomeloneAntibiotics 12 01716 i051FabKS. pneumoniae1.57[32]
(Z)-1-oxooctadec-11-enylphosphoramidic acidAntibiotics 12 01716 i052PlsYS. pneumoniae11[33]
1,1-difluoro-2-oxotridecylphosphonic acidAntibiotics 12 01716 i053PlsYB. anthracis25[33]
phenyl (8-phenyloctanoyl) sulfamateAntibiotics 12 01716 i054PlsYS. aureus25[34]
dioctylamineAntibiotics 12 01716 i055CfaSH. pylori63.81[35]
IC50 is determined as concentration required for 50% inhibition of activity of appropriate enzyme.
Table 2. Major inhibitors of lipid A biosynthetic pathway.
Table 2. Major inhibitors of lipid A biosynthetic pathway.
InhibitorStructureEnzymeOriginIC50, μMReferences
peptide 920NH2-SSGWMLDPIAGKWSR-COOHLpxAE. coli0.06 ± 0.01[206]
RJPXD33TNLYMLPKWDIP-NH2LpxAE. coli19.0 ± 1.2[207]
LpxDE. coli3.5 ± 0.1[207]
(R)-(3-(2-chloro-6-methoxybenzyl)morpholino)(3-(4-methylpyridin-2-yl)-1H-pyrazol-5-yl)methanoneAntibiotics 12 01716 i056LpxAE. coli0.6[208]
BB-78485Antibiotics 12 01716 i057LpxCE. coli0.16 ± 0.07[209]
L-161,240Antibiotics 12 01716 i058LpxCE. coli0.023 ± 0.003[210]
P. aeruginosa0.22 ± 0.003[210]
L-573,655Antibiotics 12 01716 i059LpxCE. coli8.5[211]
CHIR-090Antibiotics 12 01716 i060LpxCA. aeolicus~0.003[212]
E. coli0.009[213]
R. leguminosarum0.69[213]
LpxC-4Antibiotics 12 01716 i061LpxCP. aeruginosa0.001[214]
K. pneumoniae0.00007[214]
A. baumannii0.183[214]
TU-514Antibiotics 12 01716 i062LpxCA. aeolicus7.0 ± 0.5[215]
E. coli7.2 ± 1.9[215]
4-(2-Chlorophenyl)-3-hydroxy-7,7-dimethyl-2-phenyl-6,7,8,9-tetrahydro-2H-pyrazolo[3,4-b]quinolin-5(4H)-oneAntibiotics 12 01716 i063LpxDE. coli3.2[216]
1-(5-((4-(3-(trifluoromethyl)phenyl)piperazin-1-yl)sulfonyl)indolin-1-yl)ethan-1-oneAntibiotics 12 01716 i064LpxHE. coli1.2 ± 0.2[217]
AZ1Antibiotics 12 01716 i065LpxHK. pneumoniae0.36[218]
E. coli0.14[218,219]
JH-LPH-28Antibiotics 12 01716 i066LpxHK. pneumoniae0.11[218]
E. coli0.083[218]
JH-LPH-33Antibiotics 12 01716 i067LpxHK. pneumoniae0.026[218]
E. coli0.046[218]
Table 3. The effects of antibacterial agents on the model lipid membrane’s permeability.
Table 3. The effects of antibacterial agents on the model lipid membrane’s permeability.
AgentStructureCmin, μMCtr, μMLipid
Composition
References
Pore formation
gramicidin AAntibiotics 12 01716 i0680.001– *DSPC[254]
alamethicinAntibiotics 12 01716 i0690.1– *DOPS:DOPE 1:1 (m/m)[293]
pardaxinAntibiotics 12 01716 i0700.006– *soybean lecithin[258]
melittinAntibiotics 12 01716 i0710.23– *POPC:cholesterol 3:1 (m/m)[294]
magainin IAntibiotics 12 01716 i07210– *DOPC;
DOPE:DOPG 3:1 (m/m)
[260]
magainin IIAntibiotics 12 01716 i0730.08– *POPC:DOPG 6:1 (m/m);
DOPS:ergosterol 3:1 (m/m);
POPC:ergosterol 3:1 (m/m)
[261]
mastoparanAntibiotics 12 01716 i0740.68– *DPhPC[262]
ceratotoxin AAntibiotics 12 01716 i0750.02– *POPC:DOPE 7:3 (w/w);
POPC:DOPE:POPS 7:3:1 (w/w)
[271,272]
protegrin-1Antibiotics 12 01716 i0760.25–10– *DOPC:DOPE 1:1 (m/m)[259]
nisinAntibiotics 12 01716 i0770.1 lipid II[295]
~40>500TOCL[265]
cinnamycinAntibiotics 12 01716 i078~1.5>10DOPE;
TOCL
[265]
duramycinAntibiotics 12 01716 i0794.5–10– *GMO[266]
~2>12DOPE;
TOCL
[265]
rabbit α-defensins (NP-1/2)Antibiotics 12 01716 i080~1>16PE/PC/PS 2:2:1 (w/w);
PE/CL
[296,297]
daptomycinAntibiotics 12 01716 i0816.2– *DPhPG[298]
polymyxin BAntibiotics 12 01716 i0822.5>100DOPG[108]
1>20Kdo2-Lipid A
gaysemycinAntibiotics 12 01716 i083~26– *DOPG[279]
Pore formation and detergent action
cecropin AAntibiotics 12 01716 i0841>5DOPS:DOPE 1:1 (m/m)[264]
cecropin BAntibiotics 12 01716 i0851>5DOPS:DOPE 1:1 (m/m)[264]
Detergent action
cecropin P1Antibiotics 12 01716 i086– *>50DOPS:DOPE 1:1 (m/m)[264]
aurein 1.2Antibiotics 12 01716 i087– *>10DMPG[287]
Cmin—the antibiotic threshold concentration required to observe single pores; Ctr—the antibiotic threshold concentration required to disintegrate the lipid bilayers; *—data are absent. Abbreviations: DSPC—1,2-distearoyl-sn-glycero-3-phosphocholine; DPPC—1,2-dipalmitoyl-sn-glycero-3-phosphocholine; POPC—1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine; DOPC—1,2-dioleoyl-sn-glycero-3-phosphocholine; DOPE—1,2-dioleoyl-sn-glycero-3-phosphoethanolamine; DOPG—1,2-dioleoyl-sn-glycero-3-phospho-(1′-rac-glycerol); DOPS—1,2-dioleyl-sn-glycero-3-phosphoserine; DPhPC—1,2-diphytanoyl-sn-glycero-3-phosphocholine; DPhPG—1,2-diphytanoyl-sn-glycero-3-phospho-(1′-rac-glycerol); DMPG—1,2-dimyristoyl-sn-glycero-3-phospho-(1′-rac-glycerol); TOCL—1′,3′-bis-[1,2-dioleoyl-sn-glycero-3-phospho]-1′,3′-glycerol; GMO—glyceryl monooleate; Kdo2-Lipid A—di [3-deoxy-D-manno-octulosonyl]-lipid A. Abbreviations for nonproteinogenic amino acids: α-Me-Ala—α-methyl-alanyl; Phe-ol—phenylalaninol; Dab—2,4-diaminobutyric acid; Dhb—2,3-dehydroaminobutyric acid; Dhb—2,3-didehydrobutyrine, Abu—α-aminobutyric acid; MeOGlu—3-methyl-glutamic acid; Kyn—kynurenine; Orn—ornithine. Only D-enantiomers of amino acids are indicated.
Table 4. Major inhibitors of fungal sphingolipid biosynthesis.
Table 4. Major inhibitors of fungal sphingolipid biosynthesis.
InhibitorStructureEnzyme IC50, μMReferences
sphingofungin BAntibiotics 12 01716 i088SPTC. albicans0.049[350]
S. cerevisiae0.051
viridiofungin AAntibiotics 12 01716 i089SPTC. albicans0.022[350]
S. cerevisiae4.7
viridiofungin BAntibiotics 12 01716 i090SPTC. albicans0.017[350]
S. cerevisiae1.84
viridiofungin CAntibiotics 12 01716 i091SPTC. albicans0.025[350]
S. cerevisiae1.68
aureobasidin AAntibiotics 12 01716 i092IPCSC. albicans0.002[363]
C. glabrata0.002
Candida tropicalis0.003
Candida parapsilosis0.003
Candida krusei0.003
A. fumigatus0.005
Aspergillus flavus0.002
Aspergillus terreus0.004
Aspergillus niger0.004
S. cerevisiae0.0009[358]
khafrefunginAntibiotics 12 01716 i093IPCSC. albicans0.0006[356]
C. neoformans0.031
S. cerevisiae0.007
haplofungin AAntibiotics 12 01716 i094IPCSS. cerevisiae0.0025[357]
A. fumigatus0.41
haplofungin BAntibiotics 12 01716 i095IPCSS. cerevisiae0.042[357]
A. fumigatus1.33
pleofungin AAntibiotics 12 01716 i096IPCSS. cerevisiae0.007[358]
A. fumigatus0.0009
galbonolide A (rustmicin)Antibiotics 12 01716 i097IPCSC. albicans0.0038[359]
C. neoformans0.00007
S. cerevisiae0.0198
Table 6. The effects of antifungal agents on the model lipid membrane’s permeability.
Table 6. The effects of antifungal agents on the model lipid membrane’s permeability.
AgentStructureCmin, μMCtr, μMTarget LipidReferences
Pore formation
syringomycin EAntibiotics 12 01716 i1141–5– *DPhPC;
DOPS:DOPE 1:1 (m/m)
[422]
syringopeptin 22AAntibiotics 12 01716 i1150.003– *DPhPC;
DOPS:DOPE 1:1 (m/m)
[416]
syringopeptin 25AAntibiotics 12 01716 i1160.004– *PC:PE:PS 2:2:1 (m/m/m)[428]
fengycinsAntibiotics 12 01716 i1170.1–0.5>10POPC:POPE:POPG:ergosterol 2:2:5:1 (m/m)[425]
surfactinAntibiotics 12 01716 i1180.2–0.4– *DPhPC[424]
1.4– *glyceryl
monooleate
[426]
iturin AAntibiotics 12 01716 i1190.001– *egg-PC;
egg-PC:DMPE 8:2 (v/v)
[429]
– *– *glyceromonoolein[421]
mycosubtilinAntibiotics 12 01716 i120– *–*DPhPC[430]
bacillomycinsAntibiotics 12 01716 i121– *– *glyceromonoolein[421]
amphotericin BAntibiotics 12 01716 i1220.02–0.03– *phospholipid:cholesterol 20:1 (m/m)[431]
0.01>20DPhPC:ergosterol 2:1 (m/m)[432]
nystatinAntibiotics 12 01716 i1230.1– *phospholipid:cholesterol 20:1 (m/m)[431]
0.01>100DPhPC:ergosterol 2:1 (m/m)[433]
filipinAntibiotics 12 01716 i1240.02– *phospholipid:cholesterol 2:1 (v/v)[434]
0.01>100DPhPC:ergosterol 2:1 (m/m)[435]
piscidinAntibiotics 12 01716 i1250.005– *azolectin[436]
Detergent action
natamycinAntibiotics 12 01716 i126– *110DPhPC:ergosterol 2:1 (m/m)[437]
Cmin—the antibiotic threshold concentration required to observe single pores; Ctr—the antibiotic threshold concentration required to disintegrate the lipid bilayers; *—data are absent. Abbreviations: DPhPC—1,2-diphytanoyl-sn-glycero-3-phosphocholine; DOPS—1,2-dioleoyl-sn-glycero-3- phosphoserine; DOPE—1,2-dioleoyl-sn-glycero-3-phosphoethanolamine; POPC—1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine; POPE—1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoethanolamine; POPG—1-palmitoyl-2-oleoyl-sn-glycero-3-phospho-(1′-rac-glycerol); DOPG—1,2-dioleoyl-sn-glycero-3-phospho-(1′-rac-glycerol); DMPE—1,2-dimyristoyl-sn-glycero-3-phosphoethanolamine. Abbreviations for nonproteinogenic amino acids: α-Me-Ala—α-methyl-alanyl; Phe-ol—phenylalaninol; Dab—2,4-diaminobutyric acid; Dhb—2,3-dehydroaminobutyric acid; Dhb—2,3-didehydrobutyrine; Abu—α-aminobutyric acid; MeOGlu—3-methyl-glutamic acid; Kyn—kynurenine; Orn—ornithine. Only D-enantiomers of amino acids are indicated.
Table 8. Molecular tweezers and their antiviral activity.
Table 8. Molecular tweezers and their antiviral activity.
Molecular TweezerStructureVirusIC50, µMReference
CLR01Antibiotics 12 01716 i137HIV-113.7–20.1[501]
Ebola25.8[502]
Zika8.2[502]
HSV-119.3[500]
HSV-2
measles
IVA
SARS-CoV-276.7[503]
CLR05Antibiotics 12 01716 i138Zika38.1[500]
HSV-1
HSV-2
measles
influenza
SARS-CoV-2167.3[503]
CLR01eAntibiotics 12 01716 i139HIV-1~10[500]
CLR01fAntibiotics 12 01716 i140HIV-1~7[500]
CP006Antibiotics 12 01716 i141SARS-CoV-20.3[503]
CP020Antibiotics 12 01716 i142SARS-CoV-20.4[503]
CP025Antibiotics 12 01716 i143SARS-CoV-20.6[503]
CP036Antibiotics 12 01716 i144SARS-CoV-20.2[503]
Table 9. Antimicrobial peptides with direct antiviral action through lipid envelope disruption.
Table 9. Antimicrobial peptides with direct antiviral action through lipid envelope disruption.
Antimicrobial PeptideStructureVirusIC50, µMReference
D-plantaricin NC8 αAntibiotics 12 01716 i145SARS-CoV-2~0.001[507]
IVA~0.1[507]
kalata B1Antibiotics 12 01716 i146HIV~2–5[508]
Table 10. Fusion inhibitors affecting membrane curvature stress and their antiviral activity.
Table 10. Fusion inhibitors affecting membrane curvature stress and their antiviral activity.
InhibitorStructureVirusIC50, µMReference
aculeacin AAntibiotics 12 01716 i147SARS-CoV-21.3 ± 0.3[517]
anidulafunginAntibiotics 12 01716 i148SARS-CoV-24.7 ± 0.9[517]
iturin AAntibiotics 12 01716 i149SARS-CoV-210.9 ± 2.0[517]
mycosubtilinAntibiotics 12 01716 i150SARS-CoV-22.3 ± 0.3[517]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ostroumova, O.S.; Efimova, S.S. Lipid-Centric Approaches in Combating Infectious Diseases: Antibacterials, Antifungals and Antivirals with Lipid-Associated Mechanisms of Action. Antibiotics 2023, 12, 1716. https://doi.org/10.3390/antibiotics12121716

AMA Style

Ostroumova OS, Efimova SS. Lipid-Centric Approaches in Combating Infectious Diseases: Antibacterials, Antifungals and Antivirals with Lipid-Associated Mechanisms of Action. Antibiotics. 2023; 12(12):1716. https://doi.org/10.3390/antibiotics12121716

Chicago/Turabian Style

Ostroumova, Olga S., and Svetlana S. Efimova. 2023. "Lipid-Centric Approaches in Combating Infectious Diseases: Antibacterials, Antifungals and Antivirals with Lipid-Associated Mechanisms of Action" Antibiotics 12, no. 12: 1716. https://doi.org/10.3390/antibiotics12121716

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop