Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 3749 2023-05-16 10:27:49 |
2 references update Meta information modification 3749 2023-05-16 10:56:18 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Vanreppelen, G.; Wuyts, J.; Van Dijck, P.; Vandecruys, P. Sources of Antifungal Drugs. Encyclopedia. Available online: https://encyclopedia.pub/entry/44352 (accessed on 16 September 2024).
Vanreppelen G, Wuyts J, Van Dijck P, Vandecruys P. Sources of Antifungal Drugs. Encyclopedia. Available at: https://encyclopedia.pub/entry/44352. Accessed September 16, 2024.
Vanreppelen, Giel, Jurgen Wuyts, Patrick Van Dijck, Paul Vandecruys. "Sources of Antifungal Drugs" Encyclopedia, https://encyclopedia.pub/entry/44352 (accessed September 16, 2024).
Vanreppelen, G., Wuyts, J., Van Dijck, P., & Vandecruys, P. (2023, May 16). Sources of Antifungal Drugs. In Encyclopedia. https://encyclopedia.pub/entry/44352
Vanreppelen, Giel, et al. "Sources of Antifungal Drugs." Encyclopedia. Web. 16 May, 2023.
Sources of Antifungal Drugs
Edit

Due to their eukaryotic heritage, the differences between a fungal pathogen’s molecular makeup and its human host are small. Therefore, the discovery and subsequent development of novel antifungal drugs are extremely challenging. Nevertheless, since the 1940s, researchers have successfully uncovered potent candidates from natural or synthetic sources. Analogs and novel formulations of these drugs enhanced the pharmacological parameters and improved overall drug efficiency. These compounds ultimately became the founding members of novel drug classes and were successfully applied in clinical settings, offering valuable and efficient treatment of mycosis for decades.

antifungal natural products synthetic compound library

1. Introduction

From the foxfire bioluminescent mushrooms to the largest organisms on earth, fungi are diverse, ubiquitous cornerstone members of various ecosystems. Whereas many fungi are beneficial to humans, e.g., for cheese and alcohol production, a number of them also display pathogenic characteristics. Fungal infections pose a continuous global threat to human and animal health, jeopardize entire ecosystems, and place a tremendous burden on food production [1]. Fungi cause a range of infections in humans, from harmless, superficial maladies to life-threatening invasive mycoses. The global acquired immunodeficiency syndrome (AIDS) crisis, the increased use of implants, and the overall improved survival rates of immunocompromised patients have resulted in a steady increase in fungal infections [2][3]. These are associated with relatively high incidence, high mortality rates, and high hospitalization costs [4]. This is particularly the case for tenacious biofilm-associated infections [5]. Biofilms are complex three-dimensional structures with a typical micro-colony architecture characterized by extensive spatial heterogeneity and an extracellular matrix material associated with increased resistance to host immune factors and antifungals [6]. Due to these characteristics, biofilms frequently allow infections to re-establish after treatment. Therefore, it is no surprise that fungal infections are responsible for 1.4 million deaths on a global scale each year [7].
Treatment options for human fungal infections are currently limited to five different classes of antifungals, of which just three are regularly used as standalone treatments for mycosis. Figure 1 provides a chronological overview of their point of origin and both the discovery and introduction to the market of their most established member.
Figure 1. Timeline of the antifungal drug classes. The initial point of discovery of the class itself and both the discovery and introduction to the market of their most established member are depicted. The drug classes and their respective compounds are divided based on their origin, either synthetic (top) or natural (bottom). Created with BioRender.com.

 2. Natural Products

Historically, natural products have been a rich source of antimicrobials [8]. It all started when Alexander Fleming accidentally discovered penicillin [9]. He observed a mold contamination that visibly inhibited the growth of his staphylococci. Selman Waksman (1944) applied the same principle on a larger scale to screen for antimicrobials produced by Actinobacteria. This approach is also referred to as the Waksman platform. As a result, he and his team discovered streptomycin, the first antibiotic active against Gram-negative pathogens that could be used as a drug [10]. Using this same approach, they also identified several antifungal compounds (e.g., candicidin) [11]. Still, it was Hazen and Brown (1951) that discovered nystatin, the first antifungal compound from Actinobacteria, that would be developed as an antifungal drug [12].
It comes as no surprise that a large proportion of the antimicrobials that are currently applied are derived from natural products. Consider how millions of years of evolution shaped the continuous arms race between microorganisms, for which the production of antimicrobials offered a competitive edge to survive or even thrive in a certain niche. The further development of these compounds boomed, a success caused by the effectiveness of the Waksman platform, but also due to the inefficiency of synthetic screening campaigns and target-based approaches. Even though hit rates for antifungals from natural product libraries sometimes exceed synthetic screening campaigns up to 200-fold (see Table 1), they tend to be challenging drug candidates. An overall downside of natural products is that they are often large and complex, making de novo synthesis or production of analogs challenging and, consequently, making it harder to establish them as treatment options in the clinical context [13].
Table 1. An overview of what to expect when setting up a screening to find (novel) antifungals.
It is recognized that microorganisms have a complex life cycle [19]. They often reside in multicellular structures, such as biofilms, and this preferred lifestyle is also reflected in the clinical context. These biofilms are up to 100-fold more resistant to antifungals compared to planktonic cultures and it has been well-estimated that most infections originate from biofilms [20][21][22]. Strikingly, natural product antifungals appear to have higher anti-biofilm activity, compared to synthetic antifungals. Echinocandins and polyenes, both derived from natural products, are associated with strong anti-biofilm activity. In contrast, azoles, allylamines, and pyrimidine analogs are synthetic in origin and exert poor anti-biofilm properties [20][23][24][25]. The reduced efficacy of these antifungals on biofilms is attributed to their sequestration by the extracellular matrix components, which reduces the antifungal concentration able to reach the target cells, resulting in increased tolerance of cells within the biofilm [26][27].
Large-throughput screening campaigns by biotech and pharma companies, but also academia, are well-suited for lead compound discovery [28]. To fill in the gaps, over the last decade, compound libraries containing pure or semi-pure natural products have been composed. A prime example is the compound library of the National Cancer Institute’s Natural Products Branch (NPB). With over 320,000 fractions available for large-scale screening, it holds one of the largest collections of publicly available pre-fractionated natural product libraries [29]. These natural products can be derived from plants, fungi, or bacteria.
Plant-derived antifungals. Plants live in timescales that cannot be compared to those of most (micro)organisms. Combined with their sessile lifestyle, plants need defense mechanisms that trigger little to no resistance development, ensuring their usefulness throughout their lifespan. Preferably, these active compounds address various challenges that plants may experience at a given moment, such as predation by rodents or insects, and infection by microorganisms. As a result, plant-derived compounds appear to be mainly toxic with lower specificity, restricting their application scope to anti-cancer or anti-parasitic drugs [30]. Prime examples of aspecific plant-derived natural compounds are curcumin and resveratrol. These compounds have been reported to have antiviral, antibacterial, and antifungal properties (among others) [31][32][33][34]. Despite several hundreds of clinical trials and thousands of publications, these compounds are now regarded as pan-assay interference compounds (PAINs). These molecules are frequent hitters in (phenotypic) screening campaigns and often share structural features that show promiscuous biological activity. Therefore, clinical applications for most of these molecules are unlikely [18]. Notable exceptions here are anti-malaria compounds, such as artemisinin and quinine, which act on malarial mitochondria and purine nucleoside phosphorylase as their specific targets, respectively [35][36]. Interestingly, some anticancer compounds, such as camptothecin and podophyllotoxin, identified in plant extracts, are now assumed to be produced by fungal endosymbionts [37].
Fungal-derived antifungals. Remarkably, fungi are among the best producers of antifungals. As with some bacteria, fungi have multidomain non-ribosomal peptide synthetases (NRPS) that can produce peptides without the aid of ribosomes. Although the principle of NRPS is the same, clear differences between fungal and bacterial non-ribosomal peptides exist, such as peptide size distribution and monomer composition. Aside from the final peptide itself, the enzymatic synthesis methodology can strongly differ [38]. Fungal-derived natural products are often unique to a fungal genus or species, since horizontal gene transfer in fungi is rather rare compared to bacteria [39]. Therefore, the isolation of rare fungi is associated with increased chances of isolating novel natural antifungal products. Several medically useful antifungals derived from natural products produced by fungi include the echinocandins and the novel ibrexafungerp [11][40][41]. Both classes of antifungals target the catalytic subunit of β-glucan synthase. These β-glucan synthase inhibitors are the most frequently isolated compounds from fungal extracts but have never been isolated from bacterial sources [14]. The fungi that produce the natural precursors of these drugs all belong to the family of Trichocomaceae. They are aggressive colonizers and probably produce antifungals to maximize their potential as saprobes. Generally, they themselves are less susceptible to the antifungals they produce. For example, the echinocandins have strong concentration-dependent fungicidal activity against Candida but are only static against Aspergillus, a member of the Trichocomaceae family. Due to the diversity of the fungal kingdom, these family feuds should be considered when the antifungal development program focuses on different pathogenic lineages, for example, on Aspergillus or other members of the Trichocomaceae family.
Bacteria-derived antifungals. Antifungals derived from bacterial sources, in clinical use today, are all derived from Actinobacteria. These aerobic Gram-positive bacteria are highly abundant in soil and marine sediments and constitute one of the largest bacterial phyla [42]. They have a significantly larger genome size compared to other bacteria and a high G/C content. They are self-sustainable, making them easy to isolate and cultivate. Like fungi, they develop a mycelium with spores. During spore formation, the vegetative mycelium undergoes programmed cell death to reallocate nutrients to the spores. To prevent other microbes from using these nutrients, they produce secondary metabolites with antimicrobial activity [7]. Therefore, Actinobacteria and especially the Streptomyces genus are recognized as specialized producers of secondary metabolites [43][44]. It has been estimated that members of the genus of Streptomyces alone could produce up to 100,000 molecules with antimicrobial activity [45]. The potential of these bacteria has been well known for almost a century, resulting in large screening campaigns to exploit the antimicrobial potential of Actinobacteria. Cubist Pharmaceuticals, for example, screened over 107 Actinobacteria every year and estimated that a novel antibiotic could be discovered at frequencies below 10−7 per random Actinobacteria. Moreover, they estimated that the global top 10 cm of soil contains 1025–1026 Actinobacteria, leaving plenty of opportunity for further screening. Because the burden of fungal infections was often less recognized in the past, it is unlikely that as many Actinobacteria have been screened for antifungal activity as for antibiotic properties.
Despite extensive efforts, so far, the only clinically relevant antifungals discovered from Actinobacteria were the polyenes. However, due to nephrotoxicity, their implementation is limited [46][47]. Over 200 polyene compounds have been described, mainly from Streptomyces [48]. They appear to be the most abundant antifungals produced by Actinobacteria, outweighing other antifungals by a factor 20 [11]. A screening by Roemer et al. (2011) confirmed the abundance of polyenes produced by Actinobacteria. Moreover, they also concluded that most antifungals produced by Actinobacteria appear to lack specific targets, with the majority being ionophores [14]. This resulted in a decreased hit rate for target-specific antifungals derived from Actinobacteria (9%), compared to fungi (>50%). One strategy to avoid the rediscovery of polyene antifungals employs the use of a polyene-resistant test strain. However, this resistance is generally associated with a serious fitness cost [49], making it hard to use resistant strains in screening efforts to decrease polyene rediscovery. Fortunately, polyenes can be readily identified in extracts due to their distinct light absorption spectra [50].
Another interesting group of antifungals from actinobacteria are the chitin inhibitors, nikkomycins, and polyoxins. The latter was derived from Streptomyces cacaoi in 1960, while the former was derived from Streptomyces tendae in 1976 [51][52]. These peptidyl nucleoside antibiotics are analogs of the substrate UDP-N-acetylglucosamine and, therefore, act as competitive inhibitors of chitin synthase. Since chitin is a crucial component of a stable fungal cell wall and is absent in mammalian cells, it is generally considered a promising drug target [53]. Polyoxin D showed in vitro activity against Coccidioides immitis, Cryptococcus neoformans, and C. albicans, but failed to remain consistent during in vivo murine assessments [54][55][56]. The compound nikkomycin Z showed potent activity against some infections, such as coccidioidomycosis, but only displayed moderate activity against Histoplasma capsulatum, C. albicans, and C. neoformans. Furthermore, filamentous fungi and non-albicans Candida species were practically resistant. It does, however, work synergistically with glucan synthesis inhibitors and triazoles [57][58][59][60][61]. It underwent clinical trials in the 1990s, but the bankruptcy of the sponsoring pharmaceutical companies resulted in the termination of ongoing trials. Stranded as a research topic, the project was continued by the University of Arizona, which reactivated the clinical studies [59][62][63][64][65][66].
Other notable antifungals that have been discovered more recently from actinobacterial sources include bafilomycins, neomaclafungins, astolides, caniferolides, and azalomycin F [67][68][69][70][71]. However, several of these compounds also inhibit the growth of mammalian cells and bacteria, thereby diminishing their potential for development as medically useful antifungals.
Streptomyces are, historically, the most successful bacterial genus in terms of antifungal drug discovery thanks to the polyenes which have become a cornerstone in mycosis treatment. However, other genera also stood out due to their remarkable antifungal activity. Bacillus and Pseudomonas species have numerous records in the literature reporting their antifungal potency. Pseudomonas aeruginosa is a prominent opportunistic pathogen that displays an antagonistic relationship with fungal pathogens during co-infection. It secretes an array of metabolites to overcome fungal competitors during infection; as such, these metabolites are often characterized as essential virulence factors of the pathogen. These include lactones, alkyl quinolones, rhamnolipids, phenazines, and siderophores such as pyrrolnitrin. Most act as crucial quorum sensing molecules, iron scavengers, and overall virulence factors [72][73][74][75][76][77][78]. Although they exhibit strong antifungal activity, often these metabolites suffer from host toxicity, making further drug development challenging. Bacillus species, especially its most known member bacillus subtilis, have long been known for their biocontrol properties, tackling diseases caused by fungal phytopathogens. Their antifungal activity has been attributed to a multitude of compounds including but not limited to lipopeptides (surfactins, iturins, fengycins), polyketides (bacillaene, macrolactin), enzymes, such as chitinases, and volatile compounds, such as pyrazine [79][80][81][82][83][84][85][86][87][88][89][90][91][92][93][94][95]. Although an increasing number of antifungal agents have been identified and purified from both bacillus and pseudomonas species, none have been able to make it through drug development for clinical adaptation.
Microbial dark matter. Most bacteria and, to a lesser extent, fungi cannot be cultivated in standard laboratory conditions [96]. In natural ecosystems, this so-called “microbial dark matter” makes up roughly 99% of the microorganisms and comprises a diverse set of microorganisms. Undoubtedly, unknown natural compounds with antimicrobial properties stay hidden as this vast potential remains unmined. Soil-derived microorganisms can roughly be divided into three classes.
The first class, the cultured minority, comprises less than 1% of the total amount of microorganisms. Almost all bacteria in this group belong to only four phyla, namely Actinobacteria, Proteobacteria, Bacteroidetes, and Firmicutes. All microbial-derived antimicrobials used today come from this group, but the low-hanging fruits of this group have been picked [97].
The second class is the in situ cultivable group. These microorganisms cannot be immediately cultivated in a laboratory environment because growth factors, such as siderophores, are missing [97]. Cultivating these microorganisms requires more advanced methods, such as, for example, the isolation chip (iChip) developed by Nichols et al. (2010). This device holds miniaturized microbial growth chambers where single cells are confined and separated from the environment by a semi-permeable membrane [98]. This protects slow-growing species from aggressive colonizers that often dominate samples cultivated in the lab. Additionally, growth factors essential for germination or growth produced by other microorganisms or present in the soil can permeate through the membranes, making proliferation possible, which results in pure cultures of potentially novel microorganisms. Although in situ cultivation can be used to isolate a larger proportion of uncultivable microorganisms from soil samples, this approach rarely results in the isolation of microorganisms from uncultured phyla. Instead, the in situ cultivated microorganisms are usually rare or less cultured members of Actinobacteria, Proteobacteria, and Firmicutes [98]. Because these organisms are closely related to microbes that have already been extensively screened, a large proportion will likely produce the same or highly similar antimicrobials. Nevertheless, rare isolates can yield novel antimicrobials. This approach has already proven its success with the discovery of the promising antibiotic teixobactin [99]. Still, it remains to be seen whether antifungals discovered using this platform find their way to the clinic.
The third and final class are microorganisms that cannot be readily cultured in standard laboratory conditions, even when in situ cultivation devices, such as the iChip, are used. In terrestrial habitats, these microorganisms belong to phyla lacking cultured representatives, such as Acidobacteria, Chloroflexi, and Planctomycete. Cultivation is hard, if not impossible, for this group. It has been suggested that some of them are intrinsically slow growers and that cultivation is only possible after growing them for several months in the lab while retaining the correct conditions [100][101]. It remains unclear why some of these uncultured microorganisms are so abundant in the soil [96]. Probably, some necessary factors are still lacking to cultivate these microorganisms in a lab environment. Moreover, it is unknown whether these microorganisms can produce antimicrobials since they generally have relatively small genome sizes, ranging from only 0.148 Mb to 2.4 Mb [102]. It has been estimated that below a genome size of 3 Mb, polyketide synthase (PKS) and non-ribosomal peptide synthase (NRPS) genes are absent or rare [103]. As per the current literature, since these genes are critical components of secondary metabolite pathways, it is unlikely that they are abundant producers of secondary metabolites [44]. Still, it can also not be excluded that species with a small genome size encode antimicrobial molecules that are not encoded by NRPS or PKS operons. In contrast, as mentioned before, Streptomyces coelicolor has a genome size of 7.6 Mb, of which 5–10% of its genomic sequence is dedicated to secondary metabolite production [43][104]. Only a limited number of uncultured microorganisms have been whole-genome sequenced so far. Thus, all we need is the discovery of a novel genus from the microbial dark matter with a similar coding potential to Streptomyces, enabling us to unlock another era of highly successful natural product discovery.
Although beneficial, strain cultivation is not necessary for antimicrobial discovery. A recent study discovered the malacidin class of antibiotics using a culture-independent approach [105]. A polymerase chain reaction-based approach was used to amplify calcium-dependent antibiotic gene clusters directly from soil samples. After heterologous expression in the model host Streptomyces albus, secretion extracts were screened for antibiotic activity resulting in the isolation and purification of malacidins.

3. Synthetic Compounds

Whereas most antibiotics are of natural origin, the most frequently used antifungals, the azoles, are of synthetic origin. The reason for this may be the relatively late interest in antifungal drug discovery. After several decades of steady mortality rates due to candidiasis, in 1970, mortality rates increased substantially. This rise can be attributed to the use of immunosuppressive therapies, the increase in immunodeficient patients, such as those suffering from human immunodeficiency virus (HIV) infections, the increased use of antibacterial agents with a broad spectrum, and the frequent use of indwelling intravenous devices. Only in the 1980s were invasive mycoses recognized as a health threat [106]. Consequently, when large-scale screening platforms emerged, the focus resided on bacteria rather than fungi. So far, only one class of antifungals approved for standalone systemic use, the azoles, are derived from synthetic compound libraries [107][108].
Synthetic compounds are the result of available techniques and a chemist’s imagination. Consequently, they occupy a more limited chemical space than natural products [30]. Since the outcomes of these screening efforts are restricted by the envisioned goal and pharmaceutical and chemical parameters of the included compounds, compound libraries are generally biased [109][110]. One way to resolve this is by using a synthetic compound library that is comprised of a diverse set of compounds. For example, the Community for Open Antimicrobial Drug Discovery (CO-ADD) has composed a library of chemical compounds from academic sources and is continuously using crowdsourcing to increase its library size [111]. A proof-of-concept screening resulted in 20–30 times higher hit rates for bacteria (compared to commercially available libraries) and a hit rate of 0.98% for fungi [112]. This library will also be used to screen against the fungal targets C. albicans and C. neoformans.
Imidazole and triazole pharmacophores are relatively abundant in these libraries and are estimated to constitute around 15% of the hits when screening these libraries against the opportunistic fungal pathogen Candida albicans [14]. Compounds that enter these synthetic libraries need to pass through filters to make them a “good drug” later in the development process. One such rule is the Lipinski rule of five (RO5) which states that molecules should have a limited size and a lipophilic nature to ensure a good oral bioavailability [113]. Therefore, most compound libraries are biased toward compounds that follow these rules but do not necessarily have good antifungal properties. These rules were defined by comparing the properties of compounds that made it through the first phase of clinical trials. However, recently it was disproven that molecular weight can be used to predict oral bioavailability [114]. An additional advantage of these low molecular weight (<500 Dalton) compounds was their relative ease of synthesis. Therefore, when RO5 was established in 1997, synthetic compound libraries contained relatively smaller molecules resulting in a bias towards smaller molecules that were used as drugs. The molecular weight of approved drugs has been steadily increasing over the past years, and it has been estimated that good absorption drops sharply above 975 Dalton [115], almost twice the size originally described in RO5. It is expected that higher molecular weight molecules will be added to synthetic compound libraries in the future, and this concomitant increase in complexity could also yield higher hit rates against fungi.
Synthetic compound libraries are often used for target-based drug discovery, while natural products are more often screened in whole-cell assays. Unfortunately, target-based antifungal drug discovery faces identical issues to antibiotic-based drug discovery, and has so far failed to yield a clinically applied antimicrobial [14][100][116]. One study exemplifies the difficulties that an in vitro target-based screening can encounter during translation to in vivo viability screens [15]. During this study, the activator–mediator interaction responsible for Candida glabrata azole resistance (Pdr1 activation domain and the Gal11A KIX domain) was targeted. In this screen, small molecules that could inhibit this interaction would re-sensitize drug-resistant C. glabrata to azole antifungals. A dozen synthetic compound libraries were screened, totaling over 143,000 compounds. This resulted in 352 potential inhibitors in an in vitro screening. Due to the presence of the fungal cell wall, a large proportion of these active compounds lacked the ability to penetrate the fungal cell wall envelope. Only five compounds showed activity on live cells, with iKIX1 as the most promising lead, corresponding to less than 2% of all in vitro hits.
Despite the challenges associated with the screens using synthetic compound libraries, they can still be a very successful approach, as shown by the recent discovery of F901318 (olorofim) [117]. The F2G company screened 340,000 compounds against the airborne pathogenic mold Aspergillus fumigatus and discovered a novel chemical series with potent activity against Aspergillus species, but with no activity against C. albicans. This might explain why these compounds went unnoticed in previous screening campaigns, because Candida was typically the target pathogen. This indicates that using a panel of different fungi as targets can reveal novel compounds with a novel mode of action.

References

  1. Fisher, M.C.; Henk, D.A.; Briggs, C.J.; Brownstein, J.S.; Madoff, L.C.; McCraw, S.L.; Gurr, S.J. Emerging fungal threats to animal, plant and ecosystem health. Nature 2012, 484, 186–194.
  2. Alcazar-Fuoli, L.; Mellado, E. Current status of antifungal resistance and its impact on clinical practice. Br. J. Haematol. 2014, 166, 471–484.
  3. Armstrong-James, D.; Meintjes, G.; Brown, G.D. A neglected epidemic: Fungal infections in HIV/AIDS. Trends Microbiol. 2014, 22, 120–127.
  4. Benedict, K.; Jackson, B.R.; Chiller, T.; Beer, K.D. Estimation of Direct Healthcare Costs of Fungal Diseases in the United States. Clin. Infect. Dis. 2019, 68, 1791–1797.
  5. Douglas, L.J. Candida biofilms and their role in infection. Trends Microbiol. 2003, 11, 30–36.
  6. Ramage, G.; Martinez, J.P.; Lopez-Ribot, J.L. Candida biofilms on implanted biomaterials: A clinically significant problem. FEMS Yeast Res. 2006, 6, 979–986.
  7. Sanglard, D. Emerging Threats in Antifungal-Resistant Fungal Pathogens. Front. Med. 2016, 3, 11.
  8. Newman, D.J.; Cragg, G.M. Natural Products as Sources of New Drugs from 1981 to 2014. J. Nat. Prod. 2016, 79, 629–661.
  9. Fleming, A.G. Responsibilities and Opportunities of the Private Practitioner in Preventive Medicine. Can. Med. Assoc. J. 1929, 20, 11–13.
  10. Schatz, A.; Bugle, E.; Waksman, S.A. Streptomycin, a Substance Exhibiting Antibiotic Activity Against Gram-Positive and Gram-Negative Bacteria. Exp. Biol. Med. 1944, 55, 66–69.
  11. Waksman, S.A.; Lechevalier, H.A.; Schaffner, C.P. Candicidin and other polyenic antifungal antibiotics. Bull. World Health Organ. 1965, 33, 219–226.
  12. Hazen, E.L.; Brown, R. Fungicidin, an antibiotic produced by a soil actinomycete. Proc. Soc. Exp. Biol. Med. 1951, 76, 93–97.
  13. Miller, S.J.; Clardy, J. Natural products: Beyond grind and find. Nat. Chem. 2009, 1, 261–263.
  14. Roemer, T.; Xu, D.; Singh, S.B.; Parish, C.A.; Harris, G.; Wang, H.; Davies, J.E.; Bills, G.F. Confronting the challenges of natural product-based antifungal discovery. Chem. Biol. 2011, 18, 148–164.
  15. Nishikawa, J.L.; Boeszoermenyi, A.; Vale-Silva, L.A.; Torelli, R.; Posteraro, B.; Sohn, Y.J.; Ji, F.; Gelev, V.; Sanglard, D.; Sanguinetti, M.; et al. Inhibiting fungal multidrug resistance by disrupting an activator-Mediator interaction. Nature 2016, 530, 485–489.
  16. Pierce, C.G.; Chaturvedi, A.K.; Lazzell, A.L.; Powell, A.T.; Saville, S.P.; McHardy, S.F.; Lopez-Ribot, J.L. A Novel Small Molecule Inhibitor of Candida albicans Biofilm Formation, Filamentation and Virulence with Low Potential for the Development of Resistance. NPJ Biofilms Microbiomes 2015, 1, 15012.
  17. Vila, T.; Lopez-Ribot, J.L. Screening the Pathogen Box for Identification of Candida albicans Biofilm Inhibitors. Antimicrob. Agents Chemother. 2017, 61, e02006–e02016.
  18. Pouliot, M.; Jeanmart, S. Pan Assay Interference Compounds (PAINS) and Other Promiscuous Compounds in Antifungal Research. J. Med. Chem. 2016, 59, 497–503.
  19. Claessen, D.; Rozen, D.E.; Kuipers, O.P.; Sogaard-Andersen, L.; van Wezel, G.P. Bacterial solutions to multicellularity: A tale of biofilms, filaments and fruiting bodies. Nat. Rev. Microbiol. 2014, 12, 115–124.
  20. Chandra, J.; Kuhn, D.M.; Mukherjee, P.K.; Hoyer, L.L.; McCormick, T.; Ghannoum, M.A. Biofilm formation by the fungal pathogen Candida albicans: Development, architecture, and drug resistance. J. Bacteriol. 2001, 183, 5385–5394.
  21. Desai, J.V.; Mitchell, A.P.; Andes, D.R. Fungal biofilms, drug resistance, and recurrent infection. Cold Spring Harb. Perspect. Med. 2014, 4, a019729.
  22. Mathe, L.; Van Dijck, P. Recent insights into Candida albicans biofilm resistance mechanisms. Curr. Genet. 2013, 59, 251–264.
  23. Chandra, J.; Mukherjee, P.K.; Leidich, S.D.; Faddoul, F.F.; Hoyer, L.L.; Douglas, L.J.; Ghannoum, M.A. Antifungal resistance of candidal biofilms formed on denture acrylic in vitro. J. Dent. Res. 2001, 80, 903–908.
  24. Hawser, S.P.; Douglas, L.J. Resistance of Candida albicans biofilms to antifungal agents in vitro. Antimicrob. Agents Chemother. 1995, 39, 2128–2131.
  25. Ramage, G.; Wickes, B.L.; Lopez-Ribot, J.L. Biofilms of Candida albicans and their associated resistance to antifungal agents. Am. Clin. Lab. 2001, 20, 42–44.
  26. Nett, J.; Lincoln, L.; Marchillo, K.; Massey, R.; Holoyda, K.; Hoff, B.; VanHandel, M.; Andes, D. Putative role of beta-1,3 glucans in Candida albicans biofilm resistance. Antimicrob. Agents Chemother. 2007, 51, 510–520.
  27. Taff, H.T.; Mitchell, K.F.; Edward, J.A.; Andes, D.R. Mechanisms of Candida biofilm drug resistance. Future Microbiol. 2013, 8, 1325–1337.
  28. Mayr, L.M.; Fuerst, P. The future of high-throughput screening. J. Biomol. Screen. 2008, 13, 443–448.
  29. Thornburg, C.C.; Britt, J.R.; Evans, J.R.; Akee, R.K.; Whitt, J.A.; Trinh, S.K.; Harris, M.J.; Thompson, J.R.; Ewing, T.L.; Shipley, S.M.; et al. NCI Program for Natural Product Discovery: A Publicly-Accessible Library of Natural Product Fractions for High-Throughput Screening. ACS Chem. Biol. 2018, 13, 2484–2497.
  30. Lewis, K. Platforms for antibiotic discovery. Nat. Rev. Drug Discov. 2013, 12, 371–387.
  31. Abba, Y.; Hassim, H.; Hamzah, H.; Noordin, M.M. Antiviral Activity of Resveratrol against Human and Animal Viruses. Adv. Virol. 2015, 2015, 184241.
  32. Jennings, M.R.; Parks, R.J. Curcumin as an Antiviral Agent. Viruses 2020, 12, 1242.
  33. Nelson, K.M.; Dahlin, J.L.; Bisson, J.; Graham, J.; Pauli, G.F.; Walters, M.A. The Essential Medicinal Chemistry of Curcumin. J. Med. Chem. 2017, 60, 1620–1637.
  34. Vestergaard, M.; Ingmer, H. Antibacterial and antifungal properties of resveratrol. Int. J. Antimicrob. Agents 2019, 53, 716–723.
  35. Dziekan, J.M.; Yu, H.; Chen, D.; Dai, L.; Wirjanata, G.; Larsson, A.; Prabhu, N.; Sobota, R.M.; Bozdech, Z.; Nordlund, P. Identifying purine nucleoside phosphorylase as the target of quinine using cellular thermal shift assay. Sci. Transl. Med. 2019, 11, eaau3174.
  36. Wang, J.; Huang, L.; Li, J.; Fan, Q.; Long, Y.; Li, Y.; Zhou, B. Artemisinin directly targets malarial mitochondria through its specific mitochondrial activation. PLoS ONE 2010, 5, e9582.
  37. Schmidt, E.W. Trading molecules and tracking targets in symbiotic interactions. Nat. Chem. Biol. 2008, 4, 466–473.
  38. Bills, G.; Li, Y.; Chen, L.; Yue, Q.; Niu, X.M.; An, Z. New insights into the echinocandins and other fungal non-ribosomal peptides and peptaibiotics. Nat. Prod. Rep. 2014, 31, 1348–1375.
  39. Fitzpatrick, D.A. Horizontal gene transfer in fungi. FEMS Microbiol. Lett. 2012, 329, 1–8.
  40. Balkovec, J.M.; Hughes, D.L.; Masurekar, P.S.; Sable, C.A.; Schwartz, R.E.; Singh, S.B. Discovery and development of first in class antifungal caspofungin (CANCIDAS(R))—A case study. Nat. Prod. Rep. 2014, 31, 15–34.
  41. Pfaller, M.A.; Messer, S.A.; Rhomberg, P.R.; Borroto-Esoda, K.; Castanheira, M. Differential Activity of the Oral Glucan Synthase Inhibitor SCY-078 against Wild-Type and Echinocandin-Resistant Strains of Candida Species. Antimicrob. Agents Chemother. 2017, 61, e00161-17.
  42. Barka, E.A.; Vatsa, P.; Sanchez, L.; Gaveau-Vaillant, N.; Jacquard, C.; Meier-Kolthoff, J.P.; Klenk, H.P.; Clement, C.; Ouhdouch, Y.; van Wezel, G.P. Taxonomy, Physiology, and Natural Products of Actinobacteria. Microbiol. Mol. Biol. Rev. 2016, 80, 1–43.
  43. Bentley, S.D.; Chater, K.F.; Cerdeno-Tarraga, A.M.; Challis, G.L.; Thomson, N.R.; James, K.D.; Harris, D.E.; Quail, M.A.; Kieser, H.; Harper, D.; et al. Complete genome sequence of the model actinomycete Streptomyces coelicolor A3(2). Nature 2002, 417, 141–147.
  44. Liu, G.; Chater, K.F.; Chandra, G.; Niu, G.; Tan, H. Molecular regulation of antibiotic biosynthesis in streptomyces. Microbiol. Mol. Biol. Rev. 2013, 77, 112–143.
  45. Watve, M.G.; Tickoo, R.; Jog, M.M.; Bhole, B.D. How many antibiotics are produced by the genus Streptomyces? Arch. Microbiol. 2001, 176, 386–390.
  46. Bagnis, C.I.; Deray, G. Amphotericin B nephrotoxicity. Saudi J. Kidney Dis. Transpl. 2002, 13, 481–491.
  47. Ullmann, A.J. Nephrotoxicity in the setting of invasive fungal diseases. Mycoses 2008, 51 (Suppl. S1), 25–30.
  48. Lemke, A.; Kiderlen, A.F.; Kayser, O. Amphotericin B. Appl. Microbiol. Biotechnol. 2005, 68, 151–162.
  49. Vincent, B.M.; Lancaster, A.K.; Scherz-Shouval, R.; Whitesell, L.; Lindquist, S. Fitness trade-offs restrict the evolution of resistance to amphotericin B. PLoS Biol. 2013, 11, e1001692.
  50. Hargreaves, P.L.; Nguyen, T.S.; Ryan, R.O. Spectroscopic studies of amphotericin B solubilized in nanoscale bilayer membranes. Biochim. Biophys. Acta 2006, 1758, 38–44.
  51. Dahn, U.; Hagenmaier, H.; Hohne, H.; Konig, W.A.; Wolf, G.; Zahner, H. Stoffwechselprodukte von mikroorganismen. 154. Mitteilung. Nikkomycin, ein neuer hemmstoff der chitinsynthese bei pilzen. Arch. Microbiol. 1976, 107, 143–160.
  52. Suzuki, S.; Isono, K.; Nagatsu, J.; Mizutani, T.; Kawashima, Y.; Mizuno, T. A New Antibiotic, Polyoxin A. J. Antibiot. 1965, 18, 131.
  53. Ruiz-Herrera, J.; San-Blas, G. Chitin synthesis as target for antifungal drugs. Curr. Drug Targets Infect. Disord. 2003, 3, 77–91.
  54. Khare, R.K.; Becker, J.M.; Naider, F.R. Synthesis and anticandidal properties of polyoxin L analogues containing alpha-amino fatty acids. J. Med. Chem. 1988, 31, 650–656.
  55. Hector, R.F.; Pappagianis, D. Inhibition of chitin synthesis in the cell wall of Coccidioides immitis by polyoxin D. J. Bacteriol. 1983, 154, 488–498.
  56. Becker, J.M.; Covert, N.L.; Shenbagamurthi, P.; Steinfeld, A.S.; Naider, F. Polyoxin D inhibits growth of zoopathogenic fungi. Antimicrob. Agents Chemother. 1983, 23, 926–929.
  57. Chiou, C.C.; Mavrogiorgos, N.; Tillem, E.; Hector, R.; Walsh, T.J. Synergy, pharmacodynamics, and time-sequenced ultrastructural changes of the interaction between nikkomycin Z and the echinocandin FK463 against Aspergillus fumigatus. Antimicrob. Agents Chemother. 2001, 45, 3310–3321.
  58. Goldberg, J.; Connolly, P.; Schnizlein-Bick, C.; Durkin, M.; Kohler, S.; Smedema, M.; Brizendine, E.; Hector, R.; Wheat, J. Comparison of nikkomycin Z with amphotericin B and itraconazole for treatment of histoplasmosis in a murine model. Antimicrob. Agents Chemother. 2000, 44, 1624–1629.
  59. Hector, R.F.; Zimmer, B.L.; Pappagianis, D. Evaluation of nikkomycins X and Z in murine models of coccidioidomycosis, histoplasmosis, and blastomycosis. Antimicrob. Agents Chemother. 1990, 34, 587–593.
  60. Hector, R.F.; Braun, P.C. Synergistic action of nikkomycins X and Z with papulacandin B on whole cells and regenerating protoplasts of Candida albicans. Antimicrob. Agents Chemother. 1986, 29, 389–394.
  61. Hector, R.F.; Schaller, K. Positive interaction of nikkomycins and azoles against Candida albicans in vitro and in vivo. Antimicrob. Agents Chemother. 1992, 36, 1284–1289.
  62. Larwood, D.J. Nikkomycin Z-Ready to Meet the Promise? J. Fungi 2020, 6, 261.
  63. Holden, W.M.; Fites, J.S.; Reinert, L.K.; Rollins-Smith, L.A. Nikkomycin Z is an effective inhibitor of the chytrid fungus linked to global amphibian declines. Fungal Biol. 2014, 118, 48–60.
  64. Nix, D.E.; Swezey, R.R.; Hector, R.; Galgiani, J.N. Pharmacokinetics of nikkomycin Z after single rising oral doses. Antimicrob. Agents Chemother. 2009, 53, 2517–2521.
  65. Galgiani, J.N. Coccidioidomycosis: Changing perceptions and creating opportunities for its control. Ann. N. Y. Acad. Sci. 2007, 1111, 1–18.
  66. Shubitz, L.F.; Trinh, H.T.; Perrill, R.H.; Thompson, C.M.; Hanan, N.J.; Galgiani, J.N.; Nix, D.E. Modeling nikkomycin Z dosing and pharmacology in murine pulmonary coccidioidomycosis preparatory to phase 2 clinical trials. J. Infect. Dis. 2014, 209, 1949–1954.
  67. Aldholmi, M.; Marchand, P.; Ourliac-Garnier, I.; Le Pape, P.; Ganesan, A. A Decade of Antifungal Leads from Natural Products: 2010–2019. Pharmaceuticals 2019, 12, 182.
  68. Alferova, V.A.; Novikov, R.A.; Bychkova, O.P.; Rogozhin, E.A.; Shuvalov, M.V.; Prokhorenko, I.A.; Sadykova, V.S.; Kulko, A.B.; Dezhenkova, L.G.; Stepashkina, E.A.; et al. Astolides A and B, antifungal and cytotoxic naphthoquinone-derived polyol macrolactones from Streptomyces hygroscopicus. Tetrahedron 2018, 74, 7442–7449.
  69. Ding, N.; Jiang, Y.; Han, L.; Chen, X.; Ma, J.; Qu, X.; Mu, Y.; Liu, J.; Li, L.; Jiang, C.; et al. Bafilomycins and Odoriferous Sesquiterpenoids from Streptomyces albolongus Isolated from Elephas maximus Feces. J. Nat. Prod. 2016, 79, 799–805.
  70. Perez-Victoria, I.; Oves-Costales, D.; Lacret, R.; Martin, J.; Sanchez-Hidalgo, M.; Diaz, C.; Cautain, B.; Vicente, F.; Genilloud, O.; Reyes, F. Structure elucidation and biosynthetic gene cluster analysis of caniferolides A-D, new bioactive 36-membered macrolides from the marine-derived Streptomyces caniferus CA-271066. Org. Biomol. Chem. 2019, 17, 2954–2971.
  71. Sato, S.; Iwata, F.; Yamada, S.; Katayama, M. Neomaclafungins A-I: Oligomycin-class macrolides from a marine-derived actinomycete. J. Nat. Prod. 2012, 75, 1974–1982.
  72. Hammer, P.E.; Hill, D.S.; Lam, S.T.; Van Pee, K.H.; Ligon, J.M. Four genes from Pseudomonas fluorescens that encode the biosynthesis of pyrrolnitrin. Appl. Environ. Microbiol. 1997, 63, 2147–2154.
  73. Sulochana, M.B.; Jayachandra, S.Y.; Kumar, S.K.; Dayanand, A. Antifungal attributes of siderophore produced by the Pseudomonas aeruginosa JAS-25. J. Basic Microbiol. 2014, 54, 418–424.
  74. Nielsen, T.H.; Thrane, C.; Christophersen, C.; Anthoni, U.; Sorensen, J. Structure, production characteristics and fungal antagonism of tensin—A new antifungal cyclic lipopeptide from Pseudomonas fluorescens strain 96.578. J. Appl. Microbiol. 2000, 89, 992–1001.
  75. Morales, D.K.; Jacobs, N.J.; Rajamani, S.; Krishnamurthy, M.; Cubillos-Ruiz, J.R.; Hogan, D.A. Antifungal mechanisms by which a novel Pseudomonas aeruginosa phenazine toxin kills Candida albicans in biofilms. Mol. Microbiol. 2010, 78, 1379–1392.
  76. Chatterjee, P.; Sass, G.; Swietnicki, W.; Stevens, D.A. Review of Potential Pseudomonas Weaponry, Relevant to the Pseudomonas-Aspergillus Interplay, for the Mycology Community. J. Fungi 2020, 6, 81.
  77. Castaldi, S.; Masi, M.; Sautua, F.; Cimmino, A.; Isticato, R.; Carmona, M.; Tuzi, A.; Evidente, A. Pseudomonas fluorescens Showing Antifungal Activity against Macrophomina phaseolina, a Severe Pathogenic Fungus of Soybean, Produces Phenazine as the Main Active Metabolite. Biomolecules 2021, 11, 1728.
  78. Li, D.; Tao, W.; Yu, D.; Li, S. Emulsifying Properties of Rhamnolipids and Their In Vitro Antifungal Activity against Plant Pathogenic Fungi. Molecules 2022, 27, 7746.
  79. Li, X.; Zhang, Y.; Wei, Z.; Guan, Z.; Cai, Y.; Liao, X. Antifungal Activity of Isolated Bacillus amyloliquefaciens SYBC H47 for the Biocontrol of Peach Gummosis. PLoS ONE 2016, 11, e0162125.
  80. Zhao, H.; Shao, D.; Jiang, C.; Shi, J.; Li, Q.; Huang, Q.; Rajoka, M.S.R.; Yang, H.; Jin, M. Biological activity of lipopeptides from Bacillus. Appl. Microbiol. Biotechnol. 2017, 101, 5951–5960.
  81. Wu, J.J.; Chou, H.P.; Huang, J.W.; Deng, W.L. Genomic and biochemical characterization of antifungal compounds produced by Bacillus subtilis PMB102 against Alternaria brassicicola. Microbiol. Res. 2021, 251, 126815.
  82. Sarwar, A.; Brader, G.; Corretto, E.; Aleti, G.; Ullah, M.A.; Sessitsch, A.; Hafeez, F.Y. Qualitative analysis of biosurfactants from Bacillus species exhibiting antifungal activity. PLoS ONE 2018, 13, e0198107.
  83. Rafiee, F.; Reza Fazeli, M.; Akhavan Sepahi, A.; Noormohammadi, Z. Isolation, Screening and Identification of Native and New Bacillus subtilis with Strong Antifungal Compound against Fusarium oxysporum. Biocontrol Sci. 2022, 27, 201–208.
  84. Salazar, F.; Ortiz, A.; Sansinenea, E. A Strong Antifungal Activity of 7-O-Succinyl Macrolactin A vs Macrolactin A from Bacillus amyloliquefaciens ELI149. Curr. Microbiol. 2020, 77, 3409–3413.
  85. Wang, K.; Qin, Z.; Wu, S.; Zhao, P.; Zhen, C.; Gao, H. Antifungal Mechanism of Volatile Organic Compounds Produced by Bacillus subtilis CF-3 on Colletotrichum gloeosporioides Assessed Using Omics Technology. J. Agric. Food Chem. 2021, 69, 5267–5278.
  86. Um, S.; Fraimout, A.; Sapountzis, P.; Oh, D.C.; Poulsen, M. The fungus-growing termite Macrotermes natalensis harbors bacillaene-producing Bacillus sp. that inhibit potentially antagonistic fungi. Sci. Rep. 2013, 3, 3250.
  87. Kang, B.R.; Park, J.S.; Jung, W.J. Antifungal evaluation of fengycin isoforms isolated from Bacillus amyloliquefaciens PPL against Fusarium oxysporum f. sp. lycopersici. Microb. Pathog. 2020, 149, 104509.
  88. Wang, J.; Qiu, J.; Yang, X.; Yang, J.; Zhao, S.; Zhou, Q.; Chen, L. Identification of Lipopeptide Iturin A Produced by Bacillus amyloliquefaciens NCPSJ7 and Its Antifungal Activities against Fusarium oxysporum f. sp. niveum. Foods 2022, 11, 2996.
  89. Ajesh, K.; Sudarslal, S.; Arunan, C.; Sreejith, K. Kannurin, a novel lipopeptide from Bacillus cereus strain AK1: Isolation, structural evaluation and antifungal activities. J. Appl. Microbiol. 2013, 115, 1287–1296.
  90. Chen, K.; Tian, Z.; Luo, Y.; Cheng, Y.; Long, C.A. Antagonistic Activity and the Mechanism of Bacillus amyloliquefaciens DH-4 Against Citrus Green Mold. Phytopathology 2018, 108, 1253–1262.
  91. Dehghanifar, S.; Keyhanfar, M.; Emtiazi, G. Production and partial purification of thermostable bacteriocins from Bacillus pumilus ZED17 and DFAR8 strains with antifungal activity. Mol. Biol. Res. Commun. 2019, 8, 41–49.
  92. Schneider, K.; Chen, X.H.; Vater, J.; Franke, P.; Nicholson, G.; Borriss, R.; Sussmuth, R.D. Macrolactin is the polyketide biosynthesis product of the pks2 cluster of Bacillus amyloliquefaciens FZB42. J. Nat. Prod. 2007, 70, 1417–1423.
  93. Wang, D.; Li, A.; Han, H.; Liu, T.; Yang, Q. A potent chitinase from Bacillus subtilis for the efficient bioconversion of chitin-containing wastes. Int. J. Biol. Macromol. 2018, 116, 863–868.
  94. Prakash, J.; Arora, N.K. Novel metabolites from Bacillus safensis and their antifungal property against Alternaria alternata. Antonie Van Leeuwenhoek 2021, 114, 1245–1258.
  95. Zhou, M.; Li, P.; Wu, S.; Zhao, P.; Gao, H. Bacillus subtilis CF-3 Volatile Organic Compounds Inhibit Monilinia fructicola Growth in Peach Fruit. Front. Microbiol. 2019, 10, 1804.
  96. Lloyd, K.G.; Steen, A.D.; Ladau, J.; Yin, J.; Crosby, L. Phylogenetically Novel Uncultured Microbial Cells Dominate Earth Microbiomes. mSystems 2018, 3, e00055-18.
  97. Lewis, K.; Epstein, S.; D’Onofrio, A.; Ling, L.L. Uncultured microorganisms as a source of secondary metabolites. J. Antibiot. 2010, 63, 468–476.
  98. Nichols, D.; Cahoon, N.; Trakhtenberg, E.M.; Pham, L.; Mehta, A.; Belanger, A.; Kanigan, T.; Lewis, K.; Epstein, S.S. Use of ichip for high-throughput in situ cultivation of “uncultivable” microbial species. Appl. Environ. Microbiol. 2010, 76, 2445–2450.
  99. Ling, L.L.; Schneider, T.; Peoples, A.J.; Spoering, A.L.; Engels, I.; Conlon, B.P.; Mueller, A.; Schaberle, T.F.; Hughes, D.E.; Epstein, S.; et al. A new antibiotic kills pathogens without detectable resistance. Nature 2015, 517, 455–459.
  100. Davis, K.E.; Sangwan, P.; Janssen, P.H. Acidobacteria, Rubrobacteridae and Chloroflexi are abundant among very slow-growing and mini-colony-forming soil bacteria. Environ. Microbiol. 2011, 13, 798–805.
  101. George, I.F.; Hartmann, M.; Liles, M.R.; Agathos, S.N. Recovery of as-yet-uncultured soil acidobacteria on dilute solid media. Appl. Environ. Microbiol. 2011, 77, 8184–8188.
  102. Rinke, C.; Schwientek, P.; Sczyrba, A.; Ivanova, N.N.; Anderson, I.J.; Cheng, J.F.; Darling, A.; Malfatti, S.; Swan, B.K.; Gies, E.A.; et al. Insights into the phylogeny and coding potential of microbial dark matter. Nature 2013, 499, 431–437.
  103. Donadio, S.; Monciardini, P.; Sosio, M. Polyketide synthases and nonribosomal peptide synthetases: The emerging view from bacterial genomics. Nat. Prod. Rep. 2007, 24, 1073–1109.
  104. Baltz, R.H. Renaissance in antibacterial discovery from actinomycetes. Curr. Opin. Pharmacol. 2008, 8, 557–563.
  105. Hover, B.M.; Kim, S.H.; Katz, M.; Charlop-Powers, Z.; Owen, J.G.; Ternei, M.A.; Maniko, J.; Estrela, A.B.; Molina, H.; Park, S.; et al. Culture-independent discovery of the malacidins as calcium-dependent antibiotics with activity against multidrug-resistant Gram-positive pathogens. Nat. Microbiol. 2018, 3, 415–422.
  106. Nucci, M.; Marr, K.A. Emerging fungal diseases. Clin. Infect. Dis. 2005, 41, 521–526.
  107. Heeres, J.; Backx, L.J.; Mostmans, J.H.; Van Cutsem, J. Antimycotic imidazoles. part 4. Synthesis and antifungal activity of ketoconazole, a new potent orally active broad-spectrum antifungal agent. J. Med. Chem. 1979, 22, 1003–1005.
  108. Richardson, K.; Cooper, K.; Marriott, M.S.; Tarbit, M.H.; Troke, P.F.; Whittle, P.J. Discovery of fluconazole, a novel antifungal agent. Rev. Infect. Dis. 1990, 12 (Suppl. S3), S267–S271.
  109. Bohacek, R.S.; McMartin, C.; Guida, W.C. The art and practice of structure-based drug design: A molecular modeling perspective. Med. Res. Rev. 1996, 16, 3–50.
  110. Hert, J.; Irwin, J.J.; Laggner, C.; Keiser, M.J.; Shoichet, B.K. Quantifying biogenic bias in screening libraries. Nat. Chem. Biol. 2009, 5, 479–483.
  111. McGilvray, A. Compound screening: Fresh hunting ground. Nature 2016, 533, S65–S67.
  112. Desselle, M.R.; Neale, R.; Hansford, K.A.; Zuegg, J.; Elliott, A.G.; Cooper, M.A.; Blaskovich, M.A. Institutional profile: Community for Open Antimicrobial Drug Discovery—Crowdsourcing new antibiotics and antifungals. Future Sci. OA 2017, 3, FSO171.
  113. Lipinski, C.A.; Lombardo, F.; Dominy, B.W.; Feeney, P.J. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Adv. Drug Deliver. Rev. 1997, 23, 3–25.
  114. Shultz, M.D. Two Decades under the Influence of the Rule of Five and the Changing Properties of Approved Oral Drugs. J. Med. Chem. 2019, 62, 1701–1714.
  115. Pye, C.R.; Hewitt, W.M.; Schwochert, J.; Haddad, T.D.; Townsend, C.E.; Etienne, L.; Lao, Y.; Limberakis, C.; Furukawa, A.; Mathiowetz, A.M.; et al. Nonclassical Size Dependence of Permeation Defines Bounds for Passive Adsorption of Large Drug Molecules. J. Med. Chem. 2017, 60, 1665–1672.
  116. Payne, D.J.; Gwynn, M.N.; Holmes, D.J.; Pompliano, D.L. Drugs for bad bugs: Confronting the challenges of antibacterial discovery. Nat. Rev. Drug Discov. 2007, 6, 29–40.
  117. Oliver, J.D.; Sibley, G.E.M.; Beckmann, N.; Dobb, K.S.; Slater, M.J.; McEntee, L.; du Pre, S.; Livermore, J.; Bromley, M.J.; Wiederhold, N.P.; et al. F901318 represents a novel class of antifungal drug that inhibits dihydroorotate dehydrogenase. Proc. Natl. Acad. Sci. USA 2016, 113, 12809–12814.
More
Information
Subjects: Microbiology
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , ,
View Times: 351
Revisions: 2 times (View History)
Update Date: 16 May 2023
1000/1000
ScholarVision Creations