Acessibilidade / Reportar erro

Antifungal chemicals promising function in disease prevention, method of action and mechanism

Função promissora de produtos químicos antifúngicos na prevenção de doenças, método de ação e mecanismo

Abstract

The increasing use of antimicrobial drugs has been linked to the rise of drug-resistant fungus in recent years. Antimicrobial resistance is being studied from a variety of perspectives due to the important clinical implication of resistance. The processes underlying this resistance, enhanced methods for identifying resistance when it emerges, alternate treatment options for infections caused by resistant organisms, and so on are reviewed, along with strategies to prevent and regulate the formation and spread of resistance. This overview will focus on the action mechanism of antifungals and the resistance mechanisms against them. The link between antibacterial and antifungal resistance is also briefly discussed. Based on their mechanism action, antifungals are divided into three distinct categories: azoles, which target the ergosterol synthesis; 5-fluorocytosine, which targets macromolecular synthesis and polyenes, which interact physiochemically with fungal membrane sterols. Antifungal resistance can arise through a wide variety of ways. Overexpression of the target of the antifungal drug, changes to the drug target, changes to sterol biosynthesis, decreased intercellular concentration of the target enzyme, and other processes. A correlation exists between the mechanisms of resistance to antibacterial and antifungals, despite the fact that the comparison between the two is inevitably constrained by various parameters mentioned in the review. Drug extrusion via membrane pumps has been thoroughly documented in both prokaryotic and eukaryotic cells, and development of new antifungal compounds and strategies has also been well characterized.

Keywords:
antifungal; disease prevention; mechanism; strategies; application

Resumo

O uso crescente de medicamentos antimicrobianos tem sido associado ao aumento de fungos resistentes aos medicamentos nos últimos anos. A resistência antimicrobiana está sendo estudada sob diversas perspectivas devido à importante implicação clínica da resistência. Os processos subjacentes a esta resistência, os métodos melhorados para identificar a resistência quando esta surge, as opções alternativas de tratamento para infecções causadas por organismos resistentes, etc., são revistos, juntamente com estratégias para prevenir e regular a formação e propagação da resistência. Esta visão geral focará no mecanismo de ação dos antifúngicos e nos mecanismos de resistência contra eles. A ligação entre resistência antibacteriana e antifúngica também é brevemente discutida. Com base no seu mecanismo de ação, os antifúngicos são divididos em três categorias distintas: azóis, que têm como alvo a síntese do ergosterol; 5-fluorocitosina, que tem como alvo a síntese macromolecular, e polienos, que interagem físicoquimicamente com esteróis da membrana fúngica. A resistência antifúngica pode surgir de várias maneiras. Superexpressão do alvo do medicamento antifúngico, alterações no alvo do medicamento, alterações na biossíntese de esterol, diminuição da concentração intercelular da enzima-alvo e outros processos. Existe uma correlação entre os mecanismos de resistência aos antibacterianos e aos antifúngicos, apesar de a comparação entre os dois ser inevitavelmente limitada por vários parâmetros mencionados na revisão. A extrusão de fármacos através de bombas de membrana foi exaustivamente documentada em células procarióticas e eucarióticas, e o desenvolvimento de novos compostos e estratégias antifúngicas também foi bem caracterizado.

Palavras-chave:
antifúngico; prevenção de doença; mecanismo; estratégias; aplicativo

1. Introduction

Antimicrobial and antifungal resistance has dramatically increased in recent years. Resistance to antifungal and antimicrobial agents has strong correlation for mortality, morbidity, hospital expenses, as well as in society. As a result, researchers have put in a lot of time and effort to figure out how antimicrobial resistance develops, how it can be detected, how it can be prevented, what new treatment options there are for infections caused by resistant organisms, and how to treat infections caused by those organisms. Antibiotic resistance in bacteria has been the primary focus of this investigation for a number of reasons: (a) Majority of nosocomial and community-acquired infections are caused by bacteria; (b) Increasing number of antibacterial classes provides a greater diversity of resistance mechanisms to study; and (c) The ability to transfer resistance determinants from one strain of bacteria to another allows researchers to examine the genetic components of resistance in a wide variety of bacterial species.

Antibacterial drug resistance has received more attention than its fungal counterpart due to certain reasons. The importance of fungi as pathogens was not appreciated until quite recently (Podolsky, 2018PODOLSKY, S.H., 2018. The evolving response to antibiotic resistance (1945-2018). Palgrave Communications, vol. 4, no. 1, pp. 124. http://dx.doi.org/10.1057/s41599-018-0181-x.
http://dx.doi.org/10.1057/s41599-018-018...
) The number of people who died each year from candidiasis, for instance, remained around the same between 1950 and about 1970. There have been many shifts in medical practice since 1970 that have contributed to this alarming increase, including the more common use of immunosuppressive therapies, the overuse of antibacterial agents of broad-spectrum, the prevalence of intravenous devices, and the appearance of viral (chronic immunosuppressive) infections such as AIDS (Beck-Sagué and Jarvis, 1993BECK-SAGUÉ, C.M. and JARVIS, W.R., 1993. Secular trends in the epidemiology of nosocomial fungal infections in the United States, 1980-1990. The Journal of Infectious Diseases, vol. 167, no. 5, pp. 1247-1251. http://dx.doi.org/10.1093/infdis/167.5.1247. PMid:8486965.
http://dx.doi.org/10.1093/infdis/167.5.1...
). These changes, together with the accompanying rise in fungal infections, have accelerated the quest for safer, new, and more effective medicines to combat these diseases.

For nearly three decades, the only drug capable of treating severe pathogens was amphotericin B that is expected to induce substantial renal toxicity. The acceptance of imidazole’s and triazoles in the 1980s and first 1990s represented huge progress in way of treating local or systemic yeast infections safely and properly. The good safety profile of triazoles, especially fluconazole, necessitates their extensive usage. Since its discovery, fluconazole has been used to treat over 16 million people worldwide, including over 300,000 people living with HIV/AIDS in the United States alone (Firacative 2020FIRACATIVE, C., 2020. Invasive fungal disease in humans: are we aware of the real impact? Memorias do Instituto Oswaldo Cruz, vol. 115, e200430. http://dx.doi.org/10.1590/0074-02760200430. PMid:33053052.
http://dx.doi.org/10.1590/0074-027602004...
). As a result of this broad use, more and more cases of antifungal resistance are being recorded (Kaur et al., 2016KAUR, R., DHAKAD, M.S., GOYAL, R. and KUMAR, R., 2016. Emergence of non-albicans Candida species and antifungal resistance in intensive care unit patients. Asian Pacific Journal of Tropical Biomedicine, vol. 6, no. 5, pp. 455-460. http://dx.doi.org/10.1016/j.apjtb.2015.12.019.
http://dx.doi.org/10.1016/j.apjtb.2015.1...
) The clinical implications of resistant bacteria have recently been investigated (Maurice et al., 2018MAURICE, N.M., BEDI, B. and SADIKOT, R.T., 2018. Pseudomonas aeruginosa biofilms: host response and clinical implications in lung infections. American Journal of Respiratory Cell and Molecular Biology, vol. 58, no. 4, pp. 428-439. http://dx.doi.org/10.1165/rcmb.2017-0321TR. PMid:29372812.
http://dx.doi.org/10.1165/rcmb.2017-0321...
; Cândido et al., 2018CÂNDIDO, F.G., VALENTE, F.X., GRZEŚKOWIAK, Ł.M., MOREIRA, A.P.B., ROCHA, D.M.U.P. and ALFENAS, R.D.C.G., 2018. Impact of dietary fat on gut microbiota and low-grade systemic inflammation: mechanisms and clinical implications on obesity. International Journal of Food Sciences and Nutrition, vol. 69, no. 2, pp. 125-143. http://dx.doi.org/10.1080/09637486.2017.1343286. PMid:28675945.
http://dx.doi.org/10.1080/09637486.2017....
; Kahl et al., 2016KAHL, B.C., BECKER, K. and LÖFFLER, B., 2016. Clinical significance and pathogenesis of staphylococcal small colony variants in persistent infections. Clinical Microbiology Reviews, vol. 29, no. 2, pp. 401-427. http://dx.doi.org/10.1128/CMR.00069-15. PMid:26960941.
http://dx.doi.org/10.1128/CMR.00069-15...
). Hence, the medical implications of confrontation are not addressed in this article. As an alternative, we intend to concentrate on the molecular pathways of resistant bacteria. Because antibacterial mechanisms of resistance are much more detailed than antifungal signaling pathways, we have chosen to use well-described resistant bacteria processes as a framework for understanding fungal defence mechanisms, to the extent that such similarities can be applicable logically. We do this in the hopes that people who use the aforementioned medicines clinically and those who may wish to do future research on them will have a better understanding of anti-fungal impedance instruments as a result. Figure 1 representing the general concept about antifungal resistance.

Figure 1
Over view about antifungal resistance.

2. Selection of Antifungal Compounds

There is currently no gold standard for conducting in vitro susceptibility studies on the fungus, and results do not necessarily correlate with those obtained in vivo. So, the individual fungal infection is taken into account first and foremost when choosing an antifungal medication for therapeutic usage. The findings of preclinical and clinical testing with the most prevalent fungal infections clearly indicate the spectrum of activity for the approved antifungal medicines (Cavassin et al., 2021CAVASSIN, F.B., BAÚ-CARNEIRO, J.L., VILAS-BOAS, R.R. and QUEIROZ-TELLES, F., 2021. Sixty years of amphotericin B: an overview of the main antifungal agent used to treat invasive fungal infections. Infectious Diseases and Therapy, vol. 10, no. 1, pp. 115-147. http://dx.doi.org/10.1007/s40121-020-00382-7. PMid:33523419.
http://dx.doi.org/10.1007/s40121-020-003...
). This method is helpful in preventing the selection of antifungals for fungal species that are already resistant to the agent (primary resistance), but it is less effective in preventing the selection of antifungals for species that are already resistant to the agent (secondary resistance; drug-induced resistance).

With the expansion of the arsenal of antifungal drugs, drug resistance has emerged as a major issue (Revie et al., 2018REVIE, N.M., IYER, K.R., ROBBINS, N. and COWEN, L.E., 2018. Antifungal drug resistance: evolution, mechanisms and impact. Current Opinion in Microbiology, vol. 45, pp. 70-76. http://dx.doi.org/10.1016/j.mib.2018.02.005. PMid:29547801.
http://dx.doi.org/10.1016/j.mib.2018.02....
). Primary resistance, rather than secondary resistance, is the norm when it comes to polyene antifungals. That is, the species' susceptibility profiles are often fixed regardless of whether or not the species is exposed to the agent. For instance, Pseudallescheria boydii and Candida lusitaniae, two species recognized for their resistance to the antifungal amphotericin B, are well known yet do not appear to have developed via exposure to the drug (Brüggemann et al., 2022BRÜGGEMANN, R.J., JENSEN, G.M. and LASS-FLÖRL, C., 2022. Liposomal amphotericin B: the past. The Journal of Antimicrobial Chemotherapy, vol. 77, suppl. 2, pp. ii3-ii10. http://dx.doi.org/10.1093/jac/dkac351. PMid:36426673.
http://dx.doi.org/10.1093/jac/dkac351...
). Secondary resistance to amphotericin B in Candida albicans infections is extremely rare despite decades of broad clinical usage (Deorukhkar and Saini 2015DEORUKHKAR, S.C. and SAINI, S., 2015. Non albicans Candida species: a review of epidemiology, pathogenicity and antifungal resistance. Pravara Medical Review, vol. 7, no. 3, pp. 7-15.). Nevertheless, strains of the Candida species are known to exhibit both primary and secondary resistance to 5-fluorocytosine, which is why this treatment can only be used in combination therapy with other antifungal medicines.

The examination of the issue of fungal resistance to azole medications is much more involved. The medically significant yeasts and a few different azole antifungals are examples of both primary and secondary resistance. As a species, Candida krusei is notoriously difficult to treat with fluconazole. Nonetheless, there have been instances of resistance, particularly in HIV-infected hosts who have had many rounds of azole antifungal medication, despite the fact that Candida albicans strains are generally sensitive to fluconazole and certain other azole antifungals. Problems with current susceptibility testing methods and the inability to differentiate between microbiological and clinical drug resistance add more complexity to the issue of drug resistance. The latter is common when a host with a waning immune system reaches the limits of the action of an inhibiting antifungal drug.

Polyenes, azoles, and fluorocytosine have made it possible to cure illnesses that were previously deadly (Julien et al., 2021JULIEN, G.A., ADIYAGA, W.C., SAAKA, R.A.E., SUNYAZI, S.S., BATUIAMU, A.T., ABUGRI, D. and ABUGRI, J. (2021). Dermatophytic diseases: a review of Tinea pedis. medRxiv. In press. http://dx.doi.org/10.1101/2021.06.28.21259664.). The population at risk for opportunistic fungal infections has been expanding as modern medicine continues to extend life via vigorous therapy of other life-threatening illnesses including cancer. Patients in this situation provide a unique difficulty since their host immune systems are generally severely compromised. Chemotherapeutic drugs should kill fungus actively, not only stop them from spreading. Toxin-free fungicidal drugs are still being sought. Immunomodulating drugs, which can correct native host immune deficits, are another area of study.

3. Limitations of Antibacterial and Antifungal Comparison

We argue that contrasting the mechanisms by which bacteria develop resistance to antibiotics and antiseptics is a useful way to gain a better understanding of this issue across bacterial kingdoms, although this comparison will be influenced by a number of factors. The structure of bacteria and fungi significantly diverge from one another such as fungi diploid nature and increased generational time in comparison to bacteria and antifungal and antibacterial drugs now in use selectively inhibit fungi and bacteria by inhibiting their ability to carry out specific tasks. Several antibacterial drugs, for instance, work by blocking the production of peptidoglycan, a crucial building block of bacterial cell walls. On the other hand, the vast majority of antifungal drugs work by interfering with the activity or formation of ergosterol, which is an essential component of the membranes that surround fungus cells. In spite of this, the methods by which fungi gain resistant to inhibitor of ergosterol biosynthesis are startlingly similar to the mechanisms by which bacteria acquire resistance to anticell wall medicines. Analogues of a number of different kinds of antibacterial drugs that are antifungal (protein synthesis inhibitors like aminoglycosides, macrolides, and tetracyclines; topoisomerase inhibitors such as metabolic pathway inhibitors and fluoroquinolones; trimethoprim-sulfamethoxazole) do not exist, so comparisons to other forms of bacterial resistance are limited. Conversely, 5-fluorocytosine (5FC) and other antifungal nucleoside analogues have no therapeutically accessible antibacterial equivalents (despite being among antiviral combinations). Hence, the ability of fungus to produce ribosomal tolerance or topoisomerase genetic changes, nucleoside analogue, as well as the ability of microbes to build immunity to cytotoxic agents remains unknown. Interestingly, Rifampin, an antibiotic RNA polymerase inhibitor, has little inherent activity towards fungi, but when combined with amphotericin B (Beggs et al., 1976BEGGS, W.H., SAROSI, A. and WALKER, M.I., 1976. Synergistic action of amphotericin B and rifampin against Candida species. The Journal of Infectious Diseases, vol. 133, no. 2, pp. 206-209. http://dx.doi.org/10.1093/infdis/133.2.206. PMid:1245767.
http://dx.doi.org/10.1093/infdis/133.2.2...
) it is effective against various fungal species. Rifampin absorption into the fungal cell is boosted by amphotericin B's membrane action (Patra and Baek, 2017PATRA, J.K. and BAEK, K.H., 2017. Green biosynthesis of magnetic iron oxide (Fe3O4) nanoparticles using the aqueous extracts of food processing wastes under photo-catalyzed condition and investigation of their antimicrobial and antioxidant activity. Journal of Photochemistry and Photobiology. B, Biology, vol. 173, pp. 291-300. http://dx.doi.org/10.1016/j.jphotobiol.2017.05.045. PMid:28623821.
http://dx.doi.org/10.1016/j.jphotobiol.2...
). Using murine candidiasis models, (Sugio et al., 2020SUGIO, C.Y.C., GARCIA, A.A.M.N., ALBACH, T., MORAES, G.S., BONFANTE, E.A., URBAN, V.M. and NEPPELENBROEK, K.H., 2020. Candida-associated denture stomatitis and murine models: what is the importance and scientific evidence? Journal of Fungi, vol. 6, no. 2, pp. 70. http://dx.doi.org/10.3390/jof6020070. PMid:32456172.
http://dx.doi.org/10.3390/jof6020070...
) showed similar synergism between amphotericin B and 5FC. Some researchers believe amphotericin B-ergosterol interaction (Nakagawa et al., 2016NAKAGAWA, Y., UMEGAWA, Y., MATSUSHITA, N., YAMAMOTO, T., TSUCHIKAWA, H., HANASHIMA, S., OISHI, T., MATSUMORI, N. and MURATA, M., 2016. The structure of the bimolecular complex between amphotericin B and ergosterol in membranes is stabilized by face-to-face van der Waals interaction with their rigid cyclic cores. Biochemistry, vol. 55, no. 24, pp. 3392-3402. http://dx.doi.org/10.1021/acs.biochem.6b00193. PMid:27227740.
http://dx.doi.org/10.1021/acs.biochem.6b...
) causes membrane disorder, which improves 5FC uptake. Its synergistic activity is analogous to the proposed mechanism of bactericidal synergism of cell wall-active medicines and aminoglycosides versus enterococci, in which streptomycin raises intracellular concentrations when paired with penicillin in vitro over Enterococcus faecalis (Manoharan et al., 2021MANOHARAN, A., OGNENOVSKA, S., PAINO, D., WHITELEY, G., GLASBEY, T., KRIEL, F.H., FARRELL, J., MOORE, K.H., MANOS, J. and DAS, T., 2021. N-acetylcysteine protects bladder epithelial cells from bacterial invasion and displays antibiofilm activity against urinary tract bacterial pathogens. Antibiotics, vol. 10, no. 8, pp. 900. http://dx.doi.org/10.3390/antibiotics10080900. PMid:34438950.
http://dx.doi.org/10.3390/antibiotics100...
). Beggs et al. (1976)BEGGS, W.H., SAROSI, A. and WALKER, M.I., 1976. Synergistic action of amphotericin B and rifampin against Candida species. The Journal of Infectious Diseases, vol. 133, no. 2, pp. 206-209. http://dx.doi.org/10.1093/infdis/133.2.206. PMid:1245767.
http://dx.doi.org/10.1093/infdis/133.2.2...
found that amphotericin B and 5FC interact sequentially against Candida albicans, not in combination. The second constraint to comparing antibacterial and antifungal resistance mechanisms is that fungi's overall resistance mechanisms are unknown. These includes antimicrobial target modification, antibiotic modification, limited contact to the target and fewer grouping of all these etc. Antibiotic modification is the main mechanism of resistance to aminoglycoside β-lactam and antibacterial. Despite a single, unconfirmed report of dermatophytic fungi (Capek and Simek, 1971CAPEK, A. and SIMEK, A., 1971. Antimicrobial agents. XII. Relationship between biochemical resistance and microbial degradation of antimycotics. Folia Microbiologica, vol. 16, no. 6, pp. 472-475. PMid:5143694.) degrading nystatin, antibiotic modification is not a major route of antifungal resistance. Nevertheless, target changes and limited access to targets (often together) appear to be major causes of antifungal drug resistance. Thirdly, our understanding of the processes by which genes are exchanged between bacteria and other bacteria is significantly more advanced than our understanding of the existence of such mechanisms, if any, in fungi. In order to enable the transfer of resistance and virulence determinants across and between species, bacteria use a wide variety of bacteriophages, transposons and plasmids. Hence, there is ample opportunity for resistance to build and spread, even in the absence of direct selection by a particular antimicrobial pressure, and for the rapid emergence of high-level resistance. In contrast, previously documented mechanisms of antifungal resistance often involve the slow, incremental modification of cellular structures or functions to avoid the impact of an antifungal substance to which the fungus has been repeatedly exposed. This is done to counteract the antifungal drug's effects (such as the rising importance of Candida krusei in areas of extensive use in certain medical centers). Finally, the absence of plasmids and standardized bacterial strains for the study of antimicrobial resistance in bacteria is a significant obstacle to evaluating the mechanisms of resistance to antifungals and antibacterial therapies. This ease of access allows for the isolation of resistance determinants in well-characterized backgrounds, allowing researchers to assess the significance of different resistance mechanisms. Because of the higher availability of well-characterized strains and means for DNA delivery, the resistance mechanisms of bacteria may be examined with greater rigor than those of fungi. The initial step in investigating plasmid-mediated beta-lactamases in bacteria is to transfer the plasmid to a well-characterized strain, often Escherichia coli. In this way, membrane alteration and other potentially misleading mechanisms can be controlled for, and the resistance conferred by different -lactamases may be fairly compared. Due to the lack of available standardized systems for studying fungal resistance, most research on this topic is conducted on clinical strains themselves. This makes determining the precise contribution of individual resistance mechanisms to the phenotypic expression of resistance extremely challenging, if not impossible.

3.1. Mechanisms of action and resistance

One of the first steps in comprehending resistance mechanisms is learning how various antimicrobials work. In reality, in many situations, we have gained or improved our understanding of specific mechanisms of action thanks to the explication of resistance mechanisms. Since this is the case, we examine mechanisms of resistance and mechanisms of action for several types of antimicrobials together. For a more in-depth explanation of the mechanisms of action of several antifungal drugs, readers are directed to comprehensive reviews (Shafiei et al., 2020SHAFIEI, M., PEYTON, L., HASHEMZADEH, M. and FOROUMADI, A., 2020. History of the development of antifungal azoles: A review on structures, SAR, and mechanism of action. Bioorganic Chemistry, vol. 104, pp. 104240. http://dx.doi.org/10.1016/j.bioorg.2020.104240. PMid:32906036.
http://dx.doi.org/10.1016/j.bioorg.2020....
; Campoy and Adrio, 2017CAMPOY, S. and ADRIO, J.L., 2017. Antifungals. Biochemical Pharmacology, vol. 133, pp. 86-96. http://dx.doi.org/10.1016/j.bcp.2016.11.019. PMid:27884742.
http://dx.doi.org/10.1016/j.bcp.2016.11....
; Parks and Casey, 1996PARKS, L.W. and CASEY, W.M., 1996. Fungal sterols. In: R. PRASAD and M. GHANNOUM, eds. Lipids of pathogenic fungi. Boca Raton: CRC Press, pp. 63-82.; Gintjee et al., 2020GINTJEE, T.J., DONNELLEY, M.A. and THOMPSON, G.R., 2020. Aspiring antifungals: review of current antifungal pipeline developments. Journal of Fungi, vol. 6, no. 1, pp. 28. http://dx.doi.org/10.3390/jof6010028. PMid:32106450.
http://dx.doi.org/10.3390/jof6010028...
).

3.2. Fungal sterols and antimicrobial agents

Thiocarbamate/allylamine, polyene and azole antifungal drugs attribute their antifungal effects to interference with ergosterol production or direct interaction with this sterol. Fungal cell membranes are primarily composed of the sterol ergosterol (Yang et al., 2015YANG, H., TONG, J., LEE, C.W., HA, S., EOM, S.H. and IM, Y.J., 2015. Structural mechanism of ergosterol regulation by fungal sterol transcription factor Upc2. Nature Communications, vol. 6, no. 1, pp. 6129. http://dx.doi.org/10.1038/ncomms7129. PMid:25655993.
http://dx.doi.org/10.1038/ncomms7129...
).

4. Mechanisms of Action

4.1. Mechanism of action of azole based antimicrobial agents

N-substituted imidazoles have been studied for their antifungal effects (Elejalde et al., 2018ELEJALDE, N.R., MACÍAS, M., CASTILLO, J.C., SORTINO, M., SVETAZ, L., ZACCHINO, S. and PORTILLA, J., 2018. Synthesis and in vitro antifungal evaluation of novel N‐substituted 4‐aryl‐2‐methylimidazoles. ChemistrySelect, vol. 3, no. 18, pp. 5220-5227. http://dx.doi.org/10.1002/slct.201801238.
http://dx.doi.org/10.1002/slct.201801238...
). These chemicals, along with others like ketoconazole, fluconazole, and itraconazole, have proven to be effective medicines against human fungal infections. In example, fluconazole's therapeutic efficacy and safety have led to its widespread use. As a result, azole-resistant microorganisms have emerged, heightening the need to find novel chemicals that are effective against them (Meis et al., 2016MEIS, J.F., CHOWDHARY, A., RHODES, J.L., FISHER, M.C. and VERWEIJ, P.E., 2016. Clinical implications of globally emerging azole resistance in Aspergillus fumigatus. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, vol. 371, no. 1709, pp. 20150460. http://dx.doi.org/10.1098/rstb.2015.0460. PMid:28080986.
http://dx.doi.org/10.1098/rstb.2015.0460...
; Wiederhold and Patterson, 2015WIEDERHOLD, N.P. and PATTERSON, T.F., 2015. Emergence of azole resistance in Aspergillus. Seminars in Respiratory and Critical Care Medicine, vol. 36, no. 5, pp. 673-680. http://dx.doi.org/10.1055/s-0035-1562894. PMid:26398534.
http://dx.doi.org/10.1055/s-0035-1562894...
; Whaley et al., 2017WHALEY, S.G., BERKOW, E.L., RYBAK, J.M., NISHIMOTO, A.T., BARKER, K.S. and ROGERS, P.D., 2017. Azole antifungal resistance in Candida albicans and emerging non-albicans Candida species. Frontiers in Microbiology, vol. 7, pp. 2173. http://dx.doi.org/10.3389/fmicb.2016.02173. PMid:28127295.
http://dx.doi.org/10.3389/fmicb.2016.021...
). In order to treat fungal infections, 10 different azole-related drugs are in the works, as shown by a review of abstracts given at the Interscience Conference on Antimicrobial Agents and Chemotherapy in 1995 and 1996.

Fungi rely on ergosterol, a bioregulator that controls membrane fluidity, asymmetry, and, by extension, membrane integrity (Chang et al., 2015CHANG, W., ZHANG, M., LI, Y., LI, X., GAO, Y., XIE, Z. and LOU, H., 2015. Lichen endophyte derived pyridoxatin inactivates Candida growth by interfering with ergosterol biosynthesis. Biochimica et Biophysica Acta, vol. 1850, no. 9, pp. 1762-1771. http://dx.doi.org/10.1016/j.bbagen.2015.05.005. PMid:25960388.
http://dx.doi.org/10.1016/j.bbagen.2015....
). Cell membrane integrity is compromised when injected sterols include C-4 methyl groups. Several lines of evidence point to the heme protein as the principal azole target, as it is required for the cytochrome 14-demethylation P-450-dependent of lanosterol (Akhtar and Wright, 2015AKHTAR, M. and WRIGHT, J.N., 2015. Acyl-carbon bond cleaving cytochrome P450 enzymes: CYP17A1, CYP19A1 and CYP51A1. Monooxygenase. Advances in Experimental Medicine and Biology, vol. P450, pp. 107-130. http://dx.doi.org/10.1007/978-3-319-16009-2_4.
http://dx.doi.org/10.1007/978-3-319-1600...
). The plasma membrane's structure and function are altered due to the accumulation of sterol precursors, such as 14-methylated sterols (24-methylenedihydrolanosterol, 4,14-dimethylzymosterol, and lanosterol,), and the depletion of ergosterol. Antifungal effectiveness of newer triazole derivatives like fluconazole, itraconazole, and voriconazole (a triazole in development) is attributed, at least in part, to their ability to block cytochrome P-450-dependent 14-sterol demethylase (Medeiros et al., 2022MEDEIROS, C.I.S., SOUSA, M.N.A., ALMEIDA FILHO, G.G., FREITAS, F.O.R., UCHOA, D.P.L., NOBRE, M.S.C., BEZERRA, A.L.D., ROLIM, L.A.D.M.M., MORAIS, A.M.B., NOGUEIRA, T.B.S.S., NOGUEIRA, R.B.S.S., OLIVEIRA FILHO, A.A. and LIMA, E.O., 2022. Antifungal activity of linalool against fluconazole-resistant clinical strains of vulvovaginal Candida albicans and its predictive mechanism of action. Brazilian Journal of Medical and Biological Research, vol. 55, e11831. http://dx.doi.org/10.1590/1414-431x2022e11831. PMid:35976268.
http://dx.doi.org/10.1590/1414-431x2022e...
). This mechanism of action is supported by compelling data from research in which the structural genes encoding 14-methyl sterol demethylase (Medeiros et al., 2022MEDEIROS, C.I.S., SOUSA, M.N.A., ALMEIDA FILHO, G.G., FREITAS, F.O.R., UCHOA, D.P.L., NOBRE, M.S.C., BEZERRA, A.L.D., ROLIM, L.A.D.M.M., MORAIS, A.M.B., NOGUEIRA, T.B.S.S., NOGUEIRA, R.B.S.S., OLIVEIRA FILHO, A.A. and LIMA, E.O., 2022. Antifungal activity of linalool against fluconazole-resistant clinical strains of vulvovaginal Candida albicans and its predictive mechanism of action. Brazilian Journal of Medical and Biological Research, vol. 55, e11831. http://dx.doi.org/10.1590/1414-431x2022e11831. PMid:35976268.
http://dx.doi.org/10.1590/1414-431x2022e...
) (ERG11) and 5,6 sterol desaturases (ERG3) from C. glabrata were cloned and used to generate knockout mutants of each gene and of both genes together (Enkler et al., 2016ENKLER, L., RICHER, D., MARCHAND, A.L., FERRANDON, D. and JOSSINET, F., 2016. Genome engineering in the yeast pathogen Candida glabrata using the CRISPR-Cas9 system. Scientific Reports, vol. 6, no. 1, pp. 35766. http://dx.doi.org/10.1038/srep35766. PMid:27767081.
http://dx.doi.org/10.1038/srep35766...
). Phenotypic testing revealed that ERG3 deletion mutants exhibited the same level of sensitivity to fluconazole and itraconazole as their parental strains. However, the ERG11 deletion mutant as well as a double mutant in which both genes were deleted showed resistance to itraconazole, fluconazole, and amphotericin B at a concentration of 2 grams per milliliter. This type of evidence suggests that azoles and 14-demethylase have a relationship that is inhibitory to one another.

Although recent azole antifungals are 14-demethylase inhibitors, their mechanisms of action are somewhat diverse (Peyton et al., 2015PEYTON, L.R., GALLAGHER, S. and HASHEMZADEH, M., 2015. Triazole antifungals: a review. Drugs of Today, vol. 51, no. 12, pp. 705-718. http://dx.doi.org/10.1358/dot.2015.51.12.2421058. PMid:26798851.
http://dx.doi.org/10.1358/dot.2015.51.12...
; Feyereisen, 2015FEYEREISEN, R., 2015. Insect P450 inhibitors and insecticides: challenges and opportunities. Pest Management Science, vol. 71, no. 6, pp. 793-800. http://dx.doi.org/10.1002/ps.3895. PMid:25404103.
http://dx.doi.org/10.1002/ps.3895...
). Several membrane-bound enzymes and membrane lipid biosynthesis are all targets of the older imidazole derivatives like miconazole, econazole, and ketoconazole (Gottlieb et al., 2016GOTTLIEB, K., WACHER, V., SLIMAN, J. and PIMENTEL, M., 2016. Inhibition of methanogenic archaea by statins as a targeted management strategy for constipation and related disorders. Alimentary Pharmacology & Therapeutics, vol. 43, no. 2, pp. 197-212. http://dx.doi.org/10.1111/apt.13469. PMid:26559904.
http://dx.doi.org/10.1111/apt.13469...
). Voriconazole treatment of C. albicans cells led to an increase in zymosterol and squalene production (El-Azizi et al., 2015EL-AZIZI, M., FARAG, N. and KHARDORI, N., 2015. Antifungal activity of amphotericin B and voriconazole against the biofilms and biofilm-dispersed cells of Candida albicans employing a newly developed in vitro pharmacokinetic model. Annals of Clinical Microbiology and Antimicrobials, vol. 14, no. 1, pp. 21. http://dx.doi.org/10.1186/s12941-015-0083-3. PMid:25885806.
http://dx.doi.org/10.1186/s12941-015-008...
). It is not known if the buildup of these intermediates is a direct result of 14-demethylase inhibition or a subsequent effect of voriconazole interaction with other (non-14-demethylase) enzymes involved in ergosterol production. The genus examined may also affect the results seen with azoles. Fluconazole and itraconazole both block the 14-demethylase in Cryptococcus neoformans, which leads to a buildup of methylated sterol precursors (Wang et al., 2022WANG, W.Y., CAI, H.Q., QU, S.Y., LIN, W.H., LIANG, C.C., LIU, H., XIE, Z.X. and YUAN, Y.J., 2022. Genomic variation-mediating fluconazole resistance in yeast. Biomolecules, vol. 12, no. 6, pp. 845. http://dx.doi.org/10.3390/biom12060845. PMid:35740970.
http://dx.doi.org/10.3390/biom12060845...
; Debnath et al., 2017DEBNATH, A., CALVET, C.M., JENNINGS, G., ZHOU, W., AKSENOV, A., LUTH, M.R., ABAGYAN, R., NES, W.D., MCKERROW, J.H. and PODUST, L.M., 2017. CYP51 is an essential drug target for the treatment of primary amoebic meningoencephalitis (PAM). PLoS Neglected Tropical Diseases, vol. 11, no. 12, e0006104. http://dx.doi.org/10.1371/journal.pntd.0006104. PMid:29284029.
http://dx.doi.org/10.1371/journal.pntd.0...
). To a far greater extent than in fungi, azoles suppress mammalian cholesterol production at the 14-demethylation stage (Aderiye and Oluwole, 2015ADERIYE, B.I. and OLUWOLE, O.A., 2015. Disruption of fungi cell membranes by polyenes, azoles, allylamines, amino acids and peptides. Int Sci Res J, vol. 1, no. 7, pp. 108-116. http://dx.doi.org/10.18483/IRJSci.13.
http://dx.doi.org/10.18483/IRJSci.13...
). For instance, Hitchcock and Whittle (1993)HITCHCOCK, C. and WHITTLE, P.T. 1993. Chemistry and mode of action of fluconazole. In: J.W. RIPPON and R.A. FROMTLING, eds. Cutaneous antifungal agents: selected compounds in clinical practice and development. New York: Marcel Dekker Inc., pp. 183-197. demonstrated that the half-inhibitory concentration of voriconazole for P-450-dependent 14-sterol demethylase (P-450DM) in rat liver cholesterol was 7.4 M. When tested on fungal P-450DM, however, this antifungal drug showed a 50% inhibitory dosage as low as 0.03 M. Ketoconazole is the drug largely responsible for the clinical manifestations of sterol biosynthesis inhibition in humans. Antifungal drug target locations are summarized along the ergosterol biosynthesis pathway in Figure 2.

Figure 2
The ergosterol production pathway provides a convenient place to summarize the antifungal medication target areas.

4.2. Mechanism of action of allylamines

Allylamines, including such terbinafine and naftifine, have indeed been created as an entirely novel subject of ergosterol biogenesis inhibition that are essentially and compounds distinguishable from the remaining core categories of antifungal substances that hinder ergosterol biosynthesis. Terbinafine has been shown to be exceptionally effective against dermatophytes in both in vivo and in vitro testing environments. In a recent study, the minimum inhibitory concentration (MIC) of terbinafine against 179 clinical isolates of Candida albicans was found to be 1.2 g/ml when the M27 technique from the National Council for Clinical Laboratory Standards was used (Epp, 2011EPP, E. (2011). Functional genomics in Candida albicans: tackling drug resistance and morphology. Montréal: McGill University, 214 p. Thesis in Philosophy.; Rugeles, 2003RUGELES, C.M., 2003. Role of the efflux pumps CaMDR1, CDR1 and CDR2 in the fluconazole susceptibility of primary adhered Candida albicans. Atlanta: Georgia State University. 147 p. Thesis in Biology Microbiology.). Moreover, early information from our group and Ryder and colleagues suggests that terbinafine is effective at least some azole-resistant C. albicans strains (Abaci and Haliki-Uztan, 2011ABACI, O. and HALIKI-UZTAN, A., 2011. Investigation of the susceptibility of Candida species isolated from denture wearers to different antifungal antibiotics. African Journal of Microbiological Research, vol. 5, no. 12, pp. 1398-1403. http://dx.doi.org/10.5897/AJMR10.693.
http://dx.doi.org/10.5897/AJMR10.693...
). Terbinafine seems to be exceptionally active against Cryptococcus neoformans according to the identical assay framework. In an animal study, investigators are looking into the effectiveness of this representative against dispersed candidiasis.

Allylamines function by hindering the early stages of biosynthesis pathway. This reduction happens to coincide with squalene buildup and the lack of any additional sterol in-between. The squalene epoxidation is a human reaction that is catalyzed by squalene epoxidase. It is implied here that allylamine prevents the formation of sterols at this phase in the process. Isolated squalene epoxidase studies show that it serves as the target for allylamine activity. Fungal cell death is brought about mainly by squalene buildup rather than a lack of ergosterol (Vanreppelen et al., 2023VANREPPELEN, G., WUYTS, J., VAN DIJCK, P. and VANDECRUYS, P., 2023. Sources of antifungal drugs. Journal of Fungi, vol. 9, no. 2, pp. 171. http://dx.doi.org/10.3390/jof9020171. PMid:36836286.
http://dx.doi.org/10.3390/jof9020171...
). Squalene high level may rise membrane penetrability, leads to the disturbance of cellular organization (Figure 3).

Figure 3
Cellular membrane and antifungal resistance overview (A and B representing that the figure has two parts). This Figure has been reproduced for the present manuscript

5. Mechanisms of Resistance

5.1. Mechanism of resistance of azole based antimicrobial agents

Modification of azole antimicrobials has not been reported as a mechanism of resistance, as mentioned above. Hence, resistant strains either display a change in quantity or quality of target enzyme, a reduction in access to the target, or both. In the next sections, we will examine these mechanisms in depth, with a summary provided in Figure 4 and Table 1.

Figure 4
Possible resistance mechanisms can be completed in various steps. This figure has been produced for the present manuscript.
Table 1
How azole resistance develops biochemically.

5.2. Possible resistance mechanisms in microbial cells in various steps

One, the medicine does not fully stop the biological reaction because the target enzyme is overproduced. To make the medicine ineffective, step 2 involves changing the drug's target. An efflux pump removes the medication from the body in step 3. 4. The substance is blocked at the level of the cell membrane and wall. 5 The drug activity inhibits a cell's normal function, but the cell possesses a compensatory bypass pathway. Six, some of the “enzymes” in fungi that are responsible for transforming an inert medication into its active form are blocked. Seven, the cell releases enzymes into the surrounding media that break down the medication.

5.3. Polyene’s mechanism of Actions

For systemic fungal infections, polyene antifungal drugs like amphotericin B were the gold standard from the 1950s till the discoveries of the azoles (Soundarya et al., 2023SOUNDARYA, R., HALAGALI, P., PREETHI, S., VIKRAM, H.P., MEHDI, S. and SINGADI, R., 2023. Quality By Design (QBD) approach in processing of nanoparticles loading antifungal drugs. Journal of Coastal Life Medicine, vol. 11, pp. 279-290.).

Fryberg et al. (1974)FRYBERG, M., OEHLSCHLAGER, A.C. and UNRAU, A.M., 1974. Sterol biosynthesis in antibiotic-resistant yeast: nystatin. Archives of Biochemistry and Biophysics, vol. 160, no. 1, pp. 83-89. http://dx.doi.org/10.1016/S0003-9861(74)80011-1. PMid:4597559.
http://dx.doi.org/10.1016/S0003-9861(74)...
proposed that confrontation develops through the variety of natural sources resistant cells, which are found in relatively small quantities in community. These logically conferring resistance cells create modified sterols with lower affinity for nystatin binding. The rate of development when there is a dose of nystatin is thus determined by the normal development rate (in the absence of nystatin) as well as the rate at which nystatin induces biological membranes destruction. It is believed that this latter rate is a result of nystatin's affinity for the sterols found in membranes; specifically, it is believed that the higher the nystatin-sterol affinity, the higher the rate of membrane damage. As a result, each susceptible and resistant strain should grow at a slower rate than its more susceptible parent. In fact, a distinction between resistant and strains was discovered (Fryberg et al., 1974FRYBERG, M., OEHLSCHLAGER, A.C. and UNRAU, A.M., 1974. Sterol biosynthesis in antibiotic-resistant yeast: nystatin. Archives of Biochemistry and Biophysics, vol. 160, no. 1, pp. 83-89. http://dx.doi.org/10.1016/S0003-9861(74)80011-1. PMid:4597559.
http://dx.doi.org/10.1016/S0003-9861(74)...
). Resistance to polyenes is gradually lost after serial passage on nystatin-free media, implying that the culture has been repopulated by cells producing sterols with a higher affinity for nystatin. The molecular genetics underlying these sterol content shifts are not well understood. Resistance, according to Athar and Winner (1971)ATHAR, M.A. and WINNER, H.I., 1971. Development of resistance by Candida species to polyene antibiotics in vitro. Journal of Medical Microbiology, vol. 4, no. 4, pp. 505-517. http://dx.doi.org/10.1099/00222615-4-4-505. PMid:5316339.
http://dx.doi.org/10.1099/00222615-4-4-5...
, is the result of mutation rather than selection.

The majority of our understanding of the mechanisms of polyene resistance in species is based on research that used mutants created by growing cells under the influence of different doses of antifungals (multistep mutants) (Huang et al., 2023HUANG, T., LI, X., MAIER, M., O’BRIEN-SIMPSON, N.M., HEATH, D.E. and O’CONNOR, A.J., 2023. Using inorganic nanoparticles to fight fungal infections in the antimicrobial resistant era. Acta Biomaterialia, vol. 158, pp. 56-79. http://dx.doi.org/10.1016/j.actbio.2023.01.019. PMid:36640952.
http://dx.doi.org/10.1016/j.actbio.2023....
), (ii) revealing the cells to a steepest descent accumulation, or (iii) generating genetic variants by one-step genetic change with mutagenic agents. Sneha et al. (2023)SNEHA, K.S., NATARAJAN, S., BOAZ, K., RAMAPURAM, J., BALIGA, S., MANAKTALA, N. and CHITRA, N.S., 2023. Candida profile in HIV-positive children needs a dynamic clinical appraisal: a microbiological study. Research Journal of Pharmacy and Technology, vol. 16, no. 1, pp. 423-428. suggested a “biochemical” hypothesis in which resistance develops as a result of changes in the sterol subject matter of the cells, whether quantitatively or qualitatively. According to this concept, in order to guarantee the highest possible level of changed sterol content, the cells involved need to bind significantly less polyene. This decreased binding of polyenes in C. albicans mutants could be attributed to a reduction in the overall ergosterol level in the cell, without concomitant changes in sterol composition; (ii) ergosterol, cholesterol, or stigmasterol by a 3-hydroxy or 3-oxo sterol; or (iii) ergosterol, cholesterol, or stigmasterol by a 4-hydroxy or 4-oxoste (Qin and Dong, 2023QIN, D. and DONG, J., 2023. Multi-level optimization and strategies in microbial biotransformation of nature products. Molecules, vol. 28, no. 6, pp. 2619. http://dx.doi.org/10.3390/molecules28062619. PMid:36985591.
http://dx.doi.org/10.3390/molecules28062...
); or (iii) readjustment, or covering up, of original ergosterol so that tying with polyenes is substituent or thermochemical less favoured.

Various investigators have offered proof that supports all of these potentials. Capek et al. (1974)CAPEK, A., SIMEK, A., BRŮNA, L., SVÁB, A. and BUDĔSÍNSKÝ, Z., 1974. Antimicrobial agents. XXI. Dependence of antifungal activity on the structure of the side chain in the pyrimidine group. Folia Microbiologica, vol. 19, no. 2, pp. 169-171. PMid:4611876. proved that the advancement of inactivating resistance (induced by an adaptation mechanism) in a C. albicans strain was accompanied by a reduction in cell ergosterol content. This reduction in ergosterol content was caused by suppression of synthesis rather than enzymatic degradation of preformed ergosterol. Likewise, Dick et al. (1980)DICK, J.D., MERZ, W.G. and SARAL, R., 1980. Incidence of polyene-resistant yeasts recovered from clinical specimens. Antimicrobial Agents and Chemotherapy, vol. 18, no. 1, pp. 158-163. http://dx.doi.org/10.1128/AAC.18.1.158. PMid:7416742.
http://dx.doi.org/10.1128/AAC.18.1.158...
examined 27 polyene-resistant C. albicans isolates procured from patient with low white blood cells and discovered that their ergosterol content was reduced by 74 to 85%. As a result, reduced ergosterol content may result in reduced susceptibility to polyenes.

Fryberg et al. (1974)FRYBERG, M., OEHLSCHLAGER, A.C. and UNRAU, A.M., 1974. Sterol biosynthesis in antibiotic-resistant yeast: nystatin. Archives of Biochemistry and Biophysics, vol. 160, no. 1, pp. 83-89. http://dx.doi.org/10.1016/S0003-9861(74)80011-1. PMid:4597559.
http://dx.doi.org/10.1016/S0003-9861(74)...
experimented with a number of resistant Candida strains and discovered that more resistant strains inhabited main sterols that resulted from biosynthesis pathway obstruction (leading to ergosterol) at progressively early periods of the process. They came to the conclusion that cultures with 8-sterols had a significantly higher level of resistance to polyenes than cultures with 7-sterols, which have an even higher level of resistance than cultures with 5,7-sterols. These researchers discovered a link among amphotericin B resistance and sterol structure. The conferring resistance following treatment isolate lacked 8,7-sterol isomerase, resulting in a buildup of ergosta-5,8,22-dienol, ergosta-8,22-dienol, fecosterol, and ergosta-8-enol and a decline of ergosterol, the significant sterol in the vulnerable pre - treatment isolate. Mbongo et al. (1998)MBONGO, N., LOISEAU, P.M., BILLION, M.A. and ROBERT-GERO, M., 1998. Mechanism of amphotericin B resistance in Leishmania donovani promastigotes. Antimicrobial Agents and Chemotherapy, vol. 42, no. 2, pp. 352-357. http://dx.doi.org/10.1128/AAC.42.2.352. PMid:9527785.
http://dx.doi.org/10.1128/AAC.42.2.352...
recently offered further proof that the process of amphotericin B resistance in Leishmania donovani includes the replacement of some other sterol for ergosterol in the cell membrane. This replacement results in an alteration in membrane permeability and a decrease in amphotericin B affinity for such surface modification.

Kerridge and colleagues conducted extensive research on the role of parts of cell walls in influencing polyene interactions with their main action area, the cytoplasmic membrane. These researchers examined the polyene vulnerability of exponential- and stationary-phase candidal cells and discovered that static cells were much more resistant than exponential-phase cells. This analysis was explained by the fact that in exponential-phase cells, cell wall constituents are broken down and resynthesised at a rapid rate, leading to better polyene connectivity to the cell membrane. Stationary-phase cells, on the other hand, would've been expected to break down and synthesise cell wall at a much slower rate.

Capek and Simek (1971)CAPEK, A. and SIMEK, A., 1971. Antimicrobial agents. XII. Relationship between biochemical resistance and microbial degradation of antimycotics. Folia Microbiologica, vol. 16, no. 6, pp. 472-475. PMid:5143694. noted on the deterioration of nystatin by a caused enzymes involved evoked by dermatophytic fungi in the early 1970s. This discovery has not been affirmed by any other study. As a result, drug alteration is thought to be an improbable way to become resistant to polyene antimicrobial compounds. Moreover, because polyenes do not need to get into cells, efflux processes are highly improbable to play a role in developing resistance.

A small amount of research has investigated the genetic component of polyene resistance, with the majority focusing on Saccharomyces cerevisiae. Molzahn and Woods (1972)MOLZAHN, S.W. and WOODS, R.A., 1972. Polyene resistance and the isolation of sterol mutants in Saccharomyces cerevisiae. Journal of General Microbiology, vol. 72, no. 2, pp. 339-348. http://dx.doi.org/10.1099/00221287-72-2-339. PMid:4562308.
http://dx.doi.org/10.1099/00221287-72-2-...
isolated and characterised S. cerevisiae mutants (n = 103) that were resilient to polyenes such as nystatin, filipin, and pemaricin. Mutation was assigned to four unrelated genes: pol5, pol3, pol2, and pol1. The polyene in use for mutated isolation was determined to have connections to I the severity of cross-resistance and (ii) the variety of mutations with genetic variations at specific pol genes. The latter sterol, on the other hand, wasn't really identified among the various mutants, whereas ergosterol was absent in the pol2 mutant and present at really low concentrations in the pol3 mutant. Even though the communication between both the pol genes is unknown, data derived from UV absorption spectra recommended that all these mutations have an epigenetic modification connection, that is, those who act in sequence instead of parallel.

  1. Antimicrobial impedance correlation. It is hard to make parallels with antibiotic resistant mechanisms because very little recognised about the pathways through which microorganisms modify their ergosterol material in connection with polyene tolerance. Polyene action is comparable to that of the glycopeptide antibiotics vancomycin and teicoplanin in that it tends to involve direct contact with a systemic cellular component (instead of an enzyme or an aspect of the synthesis of proteins machines and equipment like a ribosome). Glycopeptide antibacterial drugs function by attaching to the pentapeptide peptidoglycan prediction' terminal D-alanyl-D-alanine. This adhesion hinders the catalysis of the terminal D-alanine, which offers the energy needed for the establishment of the relationship that creates the cross-bridge among different peptide side changes, as well as the transglycosylation required for peptidoglycan biosynthetic pathway.

It is uncertain how substantial glycopeptide antimicrobial resistance is in comparison to polyene resistance in fungi. This is because the mechanism of glycopeptide antimicrobial resistance is the direct consequence of the gaining of a tolerance operon, as was discussed above. Maybe of much greater importance is the fact that glycopeptide resistance was only recently discovered in Staphylococcus haemolyticus. This resistance, which is comparable to polyene resistance, develops as a result of a combination of serial passing on antibiotic plates and, presumably, regular exposure to vancomycin in individuals who have been treated with fluid resuscitation for renal failure. Support levels appear to be connected with replacements in the bridge connecting the peptide side chains, despite the fact that the precise mechanisms that underlie this type of resistance are still a mystery. Because of these variations in bridge content, the MICs for certain antibacterial agents are likely to be greater. This is because the collaborative binding of glycopeptides to the target may be inhibited.

5.4. Mechanism of resistance of allylamines

Even though diagnostic failure has indeed been noticed in treated patients with terbinafine, allylamine resistance among human fungal pathogens has not been identified as being associated with diagnostic use of terbinafine and naftifine. Resistance may well be anticipated with enhanced consumption of this agent, as vanden Bossche et al. (1992)VANDEN BOSSCHE, H., MARICHAL, H.P., ODDS, F., LE JEUNE, L. and COENE, M.-C., 1992. Characterization of an azole-resistant Candida glabrata isolate. Antimicrobial Agents and Chemotherapy, vol. 36, no. 12, pp. 2602-2610. http://dx.doi.org/10.1128/AAC.36.12.2602. PMid:1482129.
http://dx.doi.org/10.1128/AAC.36.12.2602...
noted a C. glabrata strain that's become resistant to fluconazole and demonstrated cross-resistance to terbinafine. According to other researchers, CDR1 can use terbinafine as a substrate. As a result, the device for developing resistance to allylamines has already been prevalent in some fungal strains. (a) Antibacterial resistance correlation. Because allylamine-resistant fungi are still poorly understood, correlations of resistant strains are moot. It is important to note, even so, that the various sites of action of azoles, polyenes, and allylamines depict the sequences actions on the formation of cell walls displayed by various antimicrobial agents, which include phosphomycin (a phosphoenolpyruvate analogue which behaves at an initial level in peptidoglycan synthesis), penicillin (which acts at an initial stage), and vancomycin (that also behaves at an intermediate step) (which acts at the final step in cross-linking). underlying biological mechanisms of antibacterial drugs have been elucidated by analyzing the buildup of particular precursors after acute poisoning, comparable to the investigation of the formation of cell walls in bacteria. Subsequently all of the antibiotics doing at different steps of the same procedure, it is possibly not astonishing that exact mutations will result in cross-resistance to numerous of the compounds.

5.5. Echinocandins mechanism of resistance

Resistance to echinocandins is an extremely unusual occurrence (Espinel-Ingroff, 2008ESPINEL-INGROFF, A., 2008. Mechanisms of resistance to antifungal agents: yeasts and filamentous fungi. Revista Iberoamericana de Micologia, vol. 25, no. 2, pp. 101-106. http://dx.doi.org/10.1016/S1130-1406(08)70027-5. PMid:18473504.
http://dx.doi.org/10.1016/S1130-1406(08)...
). For instance, it is estimated that more than 97% of clinical isolates belonging to the Candida genus are susceptible to these medications (Pfaller et al., 2008PFALLER, M., BOYKEN, L., HOLLIS, R.J., KROEGER, J., MESSER, S.A., TENDOLKAR, S. and DIEKEMA, D.J., 2008. In vitro susceptibility of invasive isolates of Candida spp. to anidulafungin, caspofungin, and micafungin: six years of global surveillance. Journal of Clinical Microbiology, vol. 46, no. 1, pp. 150-156. http://dx.doi.org/10.1128/JCM.01901-07. PMid:18032613.
http://dx.doi.org/10.1128/JCM.01901-07...
). This is the case for the most majority of Candida species. In contrast to the acquired resistance shown in other fungi, the intrinsic echinocandin resistance seen in Cryptococcus neoformans is not associated with a mutation in either the FKS2 or FKS1 gene. Echinocandins do, in fact, inhibit the enzyme 1–3-glucan synthase in C. neoformans; yet, this yeast is still capable of growing even in the presence of high doses of these medications. This species of C. neoformans appears to have a unique cell-wall polysaccharides composition, which may be the cause of its resistance to echinocandins.

6. 5-Fluorocytosine Mechanism of Resistance

5FC resistance is a condition that occurs quite frequently. The development of resistance can either be innate, as is the case with Candida tropicalis, or it can be acquired by the selection of resistant mutants following antifungal treatment. 7-8% of clinical isolates belonging to the Candida genus are resistant to 5FC. Nevertheless, this frequency jumps to 22% when only nonalbicans Candida species are included. There is a resistance to 5FC in about one to two percent of the clinical isolates of Cryptococcus neoformans (Medoff and Kobayashi, 1980MEDOFF, G. and KOBAYASHI, G.S., 1980. Strategies in the treatment of systemic fungal infections. The New England Journal of Medicine, vol. 302, no. 3, pp. 145-155. http://dx.doi.org/10.1056/NEJM198001173020304. PMid:6985703.
http://dx.doi.org/10.1056/NEJM1980011730...
).

6.1. Activity of compounds against cell wall

The fungus cell walls include substances that are distinctive to the fungal kingdom, such as mannan, chitin, and glucans. Because these constituents are not naturally occurring, they have indeed been identified as potential targets with selective toxic effects. Our understanding of the structure of cell walls of medically important fungi is primarily based on research with Candida albicans. This yeast's cell wall is a multi - layered structure made up of chitin, -glucan, and mannoprotein, with the latter two constituting up to 80% of the wall mass (Martínez et al., 2020MARTÍNEZ, J.M., DELSO, C., ÁLVAREZ, I. and RASO, J., 2020. Pulsed electric field‐assisted extraction of valuable compounds from microorganisms. Comprehensive Reviews in Food Science and Food Safety, vol. 19, no. 2, pp. 530-552. http://dx.doi.org/10.1111/1541-4337.12512. PMid:33325176.
http://dx.doi.org/10.1111/1541-4337.1251...
). The external layers are made up of mannan, mannoprotein, and (1,6)-glucan, while the internal layers are mostly made up of chitin and (1,3)-glucan with some mannoprotein. Throughout the last 30 years, a variety of substances with the capacity to influence the cell membranes of fungi has indeed been identified and explained. We will limit this overview to glucan synthesized inhibition because at least one antifungal substance from this class is currently being tested in clinical studies (MK-0991), being developed by scientists. Chitin synthesis inhibition, such as nikkomycins, have indeed been thoroughly researched, but no companies are manufacturing has been established.

6.2. Inhibitors of glucan

Of the three compound categories (aculeacins, papulacandins and echinocandins) that are specific inhibitors of fungal 3-glucan synthase, only echinocandins are currently being aggressively followed in clinical studies to assess their well-being, reduced side effects, and effectiveness against candidiasis. This is being done in order to determine whether or not they are effective against candidiasis. Both in vitro and in vivo testing has shown that echinocandins, which appear to be lipopeptides, possess antifungal properties that are effective against Candida and Aspergillus species.

6.3. Mechanism of Actions

In a manner that is non-competitive, glucan inhibition inhibits -(1,3)-glucan synthetase. This protein has a molecular weight of 210 kilodaltons and is a key component of membranes. These chemicals inhibit the production of the systemic glucan component in fungi without having an effect on the formation of nucleic acid or mannan. The inhibition of glucan biosynthesis pathways has downstream effects on other aspects of live cells, such as a reduction in the amount of ergosterol and lanosterol that are present and an increase in the amount of chitin that is present in cell walls. Fungi experience cytogenetic and microscopic changes as a result of the inhibition of -(1,3)-glucan synthetase (Kurtz and Douglas, 1997KURTZ, M.B. and DOUGLAS, C.M., 1997. Lipopeptide inhibitors of fungal glucan synthase. Journal of Medical and Veterinary Mycology, vol. 35, no. 2, pp. 79-86. http://dx.doi.org/10.1080/02681219780000961. PMid:9147267.
http://dx.doi.org/10.1080/02681219780000...
). These changes include the growth of pseudohyphae, the hardening of cell walls, and the failure of buds to detach from parent cells. Moreover, cells are becoming osmotically responsive, and the process of lysis is mostly restricted to the growing terminals of budding cells.

6.4. Mechanism of resistance to glucan synthesis inhibitors

Because glucan synthesised inhibition has not been used in clinical trials, conferring resistance mutations actually results from therapeutic strategies are not available. As a result, understanding of the processes of glucan synthesised inhibitor opposition is completely dependent on the evaluation of experimentally mutants. The topic that follows is informed by the results of research lab mutagenesis experiments performed by Kurtz and colleagues, who examined resistant mutants of S. cerevisiae. The goal of lipopeptides, such as echinocandins, is glucan synthase (a heterodimeric enzyme encrypted by FKS1 and RHO1 in S. cerevisiae) (Maligie and Selitrennikoff, 2005MALIGIE, M. and SELITRENNIKOFF, C., 2005. Cryptococcus neoformans resistance to echinocandins: (1,3)β-glucan synthase activity is sensitive to echinocandins. Antimicrobial Agents and Chemotherapy, vol. 49, no. 7, pp. 2851-2856. http://dx.doi.org/10.1128/AAC.49.7.2851-2856.2005. PMid:15980360.
http://dx.doi.org/10.1128/AAC.49.7.2851-...
). FKS2, some other gene found in S. cerevisiae, is morphologically similar to FKS1. FKS1 genetic mutations given this potential in vitro susceptibility to lipopeptides. Mutants in another cell membrane synthesis gene, GNS1, which encrypts an enzymatic activity implicated in fatty acid deformation, are associated with low-level resistance (10-fold). FKS2 genetic changes do not acquire resistance. Furthermore, stimulation of MDR-like genes or the classification of pathway bypass mutants do not seem to have any significance as lipopeptide resistance mechanisms. Finally, because lipopeptides do not cross the cell membrane, entry methods may not be involved in their action and thus cannot be involved in the fungal cell's response to them. These research results, in addition to the low rate of mutated gene per creation of fungal cells, imply that, at least in vitro, S. cerevisiae develops resistance to lipopeptide antimycotic agencies via mutants that change the protein encoded by FKS1, that is the primary target of the receptor and is thought to be the catalysed aspect of the wall of the fungal cell glucan synthase (Jallow and Govender, 2021JALLOW, S. and GOVENDER, N.P., 2021. Ibrexafungerp: a first-in-class oral triterpenoid glucan synthase inhibitor. Journal of Fungi, vol. 7, no. 3, pp. 163. http://dx.doi.org/10.3390/jof7030163. PMid:33668824.
http://dx.doi.org/10.3390/jof7030163...
). Even farther research with C. albicans mutations revealed that resistance to pneumocandin in S. cerevisiae and C. albicans is very similar, that once fact that Candida is diploid is considered. The discovery that the C. albicans mutations retrieved by Kurtz and colleagues were unaffected in their vulnerability to other medically used antifungal agents such as fluconazole, itraconazole, 5FC, and amphotericin B when confirmed in vitro and in animal models of candidiasis is clinically important (Maesaki et al., 2000MAESAKI, S., HOSSAIN, M.A., MIYAZAKI, Y., TOMONO, K., TASHIRO, T. and KOHNO, S., 2000. Efficacy of FK463, a (1, 3)-β-d-glucan synthase inhibitor, in disseminated azole-resistant Candida albicans infection in mice. Antimicrobial Agents and Chemotherapy, vol. 44, no. 6, pp. 1728-1730. http://dx.doi.org/10.1128/AAC.44.6.1728-1730.2000. PMid:10817741.
http://dx.doi.org/10.1128/AAC.44.6.1728-...
).

6.5. Correlation with anti-bacterial resistance

While there aren't any nucleoside analogues between many antimicrobial properties, there really are agents that need to be chemically modified in order to be active within the bacterial cell. One of them is metronidazole, a 5-nitroimidazole compound for whom the activation is dependent on nitro group decrease in the oxygen-deprived environment. Metronidazole resistance is uncommon (possibly due to its rare use in contrast to other agents) and is thought to be caused by lowered uptake or a slower rate of decrease. A more appropriate comparison includes herpes simplex virus resistance to the anti - viral complex acyclovir. This substance is phosphorylated intracellularly by thymidine kinase, which is encrypted by a virus. Cellular enzymes then transform acyclovir monophosphate to acyclovir triphosphate, at which direction the compound is incorporated into emulating viral DNA and turns as a cofactor. The most common process of acyclovir resistance is a lack of or changed substrate specific of the viral thymidine kinase. Because of these changes, acyclovir is not converted to its activated state classified the cell.

6.6. Development of novel antifungal methods

The current antifungal medicines have a restricted ability to treat infections, particularly systemic infections, and there have been no significant advances in antifungal therapeutics created in recent years. Thus, there is a need for innovative treatments that target pathogenic fungus. Throughout the course of the past few years, a number of different strategies have been developed in an effort to discover new solutions. Researchers want to find novel antifungal medications, and they expect to do it either by evaluating medical chemicals that already exist, by testing compounds derived from natural sources like plants, the sea, or microbes, or by conducting systematic screenings of chemical compound libraries. Both in vitro and in vivo studies are conducted by researchers with the goal of better understanding the fundamental biology of fungal microorganisms. Interactions between hosts and pathogenic fungi are essential to the survival of all fungal diseases. By focusing on this interaction, it is possible to develop innovative medicines that have the potential to be used either independently or in conjunction with other antifungal medications. A combination like this could possibly play a role in the evolution of antifungal medication resistance.

6.7. Compounds for novel antifungal agents

Natural compounds (NP) or natural bioactive chemicals derived from plants, other microbes, or marine species have been studied extensively for their antifungal effects (Di Santo, 2010DI SANTO, R., 2010. Natural products as antifungal agents against clinically relevant pathogens. Natural Product Reports, vol. 27, no. 7, pp. 1084-1098. http://dx.doi.org/10.1039/b914961a. PMid:20485730.
http://dx.doi.org/10.1039/b914961a...
). Some of these chemicals are studied due to the fact that they activate crucial fungal trigger systems, while others are explored purely on the basis of their antifungal potential. Although promising findings were found in this research, no chemical has yet been developed that is ready for human clinical trials. The antifungal action of the fungistatic agent fluconazole (FLC) on Candida albicans was also investigated in other investigations, with an emphasis on in vitro screens of various medicines now used in clinical practice. Many drugs, including calcineurin (Marchetti et al., 2000MARCHETTI, O., MOREILLON, P., GLAUSER, M.P., BILLE, J. and SANGLARD, D., 2000. Potent synergism of the combination of fluconazole and cyclosporine in Candida albicans. Antimicrobial Agents and Chemotherapy, vol. 44, no. 9, pp. 2373-2381. http://dx.doi.org/10.1128/AAC.44.9.2373-2381.2000. PMid:10952582.
http://dx.doi.org/10.1128/AAC.44.9.2373-...
) and Tor pathway (Cruz et al., 2001CRUZ, M.C., GOLDSTEIN, A.L., BLANKENSHIP, J., DEL POETA, M., PERFECT, J.R., MCCUSKER, J.H., BENNANI, Y.L., CARDENAS, M.E. and HEITMAN, J., 2001. Rapamycin and less immunosuppressive analogs are toxic to Candida albicans and Cryptococcus neoformans via FKBP12-dependent inhibition of TOR. Antimicrobial Agents and Chemotherapy, vol. 45, no. 11, pp. 3162-3170. http://dx.doi.org/10.1128/AAC.45.11.3162-3170.2001. PMid:11600372.
http://dx.doi.org/10.1128/AAC.45.11.3162...
) inhibitors, efflux pump inhibitors (compounds generated from milbemycin), and antibodies against heat-shock 90 protein (HSP90) were found as a result. It has been established that calcineurin pathway inhibitors significantly reduce fungal virulence and are fully active in vivo in the potentiation of fluconazole. To find more effective antifungal agents, industrial laboratories also systematically screened libraries of chemical compounds. Recently, a new, very efficient glucan synthase inhibitor, designated as compound C, was found thanks to high-throughput screening of the historical Schering-Plough drug collection. Candida albicans and Candida sp. magnifica. Caenorhabditis elegans infected with C. elegans treated with drugs was used to conduct another form of high-throughput screen of chemical libraries. Albatross (Pukkila-Worley et al., 2009PUKKILA-WORLEY, R., HOLSON, E., WAGNER, F. and MYLONAKIS, E., 2009. Antifungal drug discovery through the study of invertebrate model hosts. Current Medicinal Chemistry, vol. 16, no. 13, pp. 1588-1595. http://dx.doi.org/10.2174/092986709788186237. PMid:19442135.
http://dx.doi.org/10.2174/09298670978818...
). One of the key challenges in modern antimicrobial discovery is eliminated since compounds may be evaluated for both antifungal efficacy and host toxicity at the same time. In a preliminary search for antifungal chemicals, this new C. elegans system discovered 15 chemicals that increased the time nematodes could survive after being exposed to the medically significant human disease C. albicans. One of these compounds, caffeic acid phenethyl ester (CAPE), showed antifungal efficacy against many fungal species in vitro and in a mouse model of systemic candidiasis (Pukkila-Worley et al., 2009PUKKILA-WORLEY, R., HOLSON, E., WAGNER, F. and MYLONAKIS, E., 2009. Antifungal drug discovery through the study of invertebrate model hosts. Current Medicinal Chemistry, vol. 16, no. 13, pp. 1588-1595. http://dx.doi.org/10.2174/092986709788186237. PMid:19442135.
http://dx.doi.org/10.2174/09298670978818...
). This complete animal model may also help in the discovery of drugs that alter infection-specific expression of fungal virulence proteins.

6.8. Control measures against antifungal resistance

Fungicidal resistance inhibition and suppression techniques have yet to be characterised. Moves closer similar to those suggested for antibacterial drugs could be proposed. These indicators involve (I) prudent antifungal use, (ii) suitable medicating with a focus on wanting to avoid care with low antimicrobial activity dosage, (iii) counselling with configurations of established means, (iv) action with the suitable anti - fungal (in cases where the pathogenetic agent is known), and (v) conducting cross - sectional studies to evaluate the true prevalence of antifungal resistance. It should be noted that information supporting the use of the proposed reforms is largely suffering from a lack, and ongoing research could offer detailed guidance in the coming days. Furthermore, recent advancements in rapid fungi diagnosis may help decrease the application of improper antifungal agents to behave microbes that really are resilient to a specific agent. Regrettably, advancement in continuing to develop fungi-specific diagnostic methods has been slow. The recognized leader of a comparison technique to evaluate yeast antifungal susceptibility is positive and delivers resources for conducting investigation research.

7. Future Prospects

Over the past two–three decades, and especially in the post-genomic era, research on fungal pathogenesis mechanisms may lead to the development of novel anti-virulence agents that disarm fungal cells rather than inhibiting their growth. These agents may be used to treat fungal infections in alternative ways. This is a promising treatment for opportunistic pathogenic fungi. Anti-virulence methods automatically increase antifungal targets. Anti-virulence techniques also induce less antifungal medication resistance due to lesser selective pressure. C. albicans has been the organism most studied on this topic, with filamentation and biofilm formation, two of the most important virulence mechanisms in the pathogenesis of candidiasis, recognized as high-value targets for such innovative anti-virulence techniques. Repurposing as a fast track to antifungal medication development has grown in popularity during the past decade. Repurposing (also called repositioning) investigates new therapeutic indications for existing medications, unlike de novo drug development, which is long, expensive, and difficult. Repurposing identifies pharmaceuticals with known pharmacokinetics, pharmacodynamics, and human safety, making it cheaper, faster, and more successful than de novo discovery. Hence, repurposing can accelerate the deployment of innovative antifungal medications and shorten the bench-to-bedside time. Multi-drug-resistant and emerging viruses, like C, may require this. auris infections in recent years. Numerous studies have repurposed medicines as antifungals. Recently, these initiatives have taken advantage of repurposing libraries, a collection of hundreds of current medications that can be easily screened for novel antifungal activity. Finally, nanotechnology, using nanoscale materials, has been gaining popularity as an alternate technique for antifungal medication development. These “nanomaterials” or “nanoantibiotics” are single structures shorter than 100 nm in at least one dimension. Nanomaterials' endurance, chemical reactivity, biocompatibility, conductivity, and reduced toxicity are drawing attention. Most study on this area has examined metal nanoparticles, produced by various ways, and their direct antifungal action. Silver nanoparticles are particularly popular due to their antibacterial characteristics. Smaller nanoparticles have more antifungal action due to their structure, surface area, and size. Nanoparticles can kill drug-resistant fungal biofilms.

8. Conclusion

Impedance to antibacterial drugs is a rational and unavoidable result of employing these agencies to cure infections in humans. The accessibility of gene function tools has resulted in a rapid growth of our knowledge of the processes that cause anti - bacterial resistance to arise and expanded, and it pledges to tremendously notify the creation of new and efficient substances for prospective use. With enhanced use as well as accessibility to different categories of antifungals, it is anticipated that the amount and wide range of fungal strains conferring resistance to these agencies will start increasing. Research continues into antifungal mechanisms of resistance, as well as the advancement of experimental models (similar to those used in bacteria) in which independent antimicrobial compounds could be investigated, will be crucial parts of a strategic plan to restrict the emergence of resistance to these agencies and create safer and much more effective substances in the future.

References

  • ABACI, O. and HALIKI-UZTAN, A., 2011. Investigation of the susceptibility of Candida species isolated from denture wearers to different antifungal antibiotics. African Journal of Microbiological Research, vol. 5, no. 12, pp. 1398-1403. http://dx.doi.org/10.5897/AJMR10.693
    » http://dx.doi.org/10.5897/AJMR10.693
  • ADERIYE, B.I. and OLUWOLE, O.A., 2015. Disruption of fungi cell membranes by polyenes, azoles, allylamines, amino acids and peptides. Int Sci Res J, vol. 1, no. 7, pp. 108-116. http://dx.doi.org/10.18483/IRJSci.13
    » http://dx.doi.org/10.18483/IRJSci.13
  • AKHTAR, M. and WRIGHT, J.N., 2015. Acyl-carbon bond cleaving cytochrome P450 enzymes: CYP17A1, CYP19A1 and CYP51A1. Monooxygenase. Advances in Experimental Medicine and Biology, vol. P450, pp. 107-130. http://dx.doi.org/10.1007/978-3-319-16009-2_4
    » http://dx.doi.org/10.1007/978-3-319-16009-2_4
  • ATHAR, M.A. and WINNER, H.I., 1971. Development of resistance by Candida species to polyene antibiotics in vitro Journal of Medical Microbiology, vol. 4, no. 4, pp. 505-517. http://dx.doi.org/10.1099/00222615-4-4-505 PMid:5316339.
    » http://dx.doi.org/10.1099/00222615-4-4-505
  • BECK-SAGUÉ, C.M. and JARVIS, W.R., 1993. Secular trends in the epidemiology of nosocomial fungal infections in the United States, 1980-1990. The Journal of Infectious Diseases, vol. 167, no. 5, pp. 1247-1251. http://dx.doi.org/10.1093/infdis/167.5.1247 PMid:8486965.
    » http://dx.doi.org/10.1093/infdis/167.5.1247
  • BEGGS, W.H., SAROSI, A. and WALKER, M.I., 1976. Synergistic action of amphotericin B and rifampin against Candida species. The Journal of Infectious Diseases, vol. 133, no. 2, pp. 206-209. http://dx.doi.org/10.1093/infdis/133.2.206 PMid:1245767.
    » http://dx.doi.org/10.1093/infdis/133.2.206
  • BRÜGGEMANN, R.J., JENSEN, G.M. and LASS-FLÖRL, C., 2022. Liposomal amphotericin B: the past. The Journal of Antimicrobial Chemotherapy, vol. 77, suppl. 2, pp. ii3-ii10. http://dx.doi.org/10.1093/jac/dkac351 PMid:36426673.
    » http://dx.doi.org/10.1093/jac/dkac351
  • CAMPOY, S. and ADRIO, J.L., 2017. Antifungals. Biochemical Pharmacology, vol. 133, pp. 86-96. http://dx.doi.org/10.1016/j.bcp.2016.11.019 PMid:27884742.
    » http://dx.doi.org/10.1016/j.bcp.2016.11.019
  • CÂNDIDO, F.G., VALENTE, F.X., GRZEŚKOWIAK, Ł.M., MOREIRA, A.P.B., ROCHA, D.M.U.P. and ALFENAS, R.D.C.G., 2018. Impact of dietary fat on gut microbiota and low-grade systemic inflammation: mechanisms and clinical implications on obesity. International Journal of Food Sciences and Nutrition, vol. 69, no. 2, pp. 125-143. http://dx.doi.org/10.1080/09637486.2017.1343286 PMid:28675945.
    » http://dx.doi.org/10.1080/09637486.2017.1343286
  • CAPEK, A. and SIMEK, A., 1971. Antimicrobial agents. XII. Relationship between biochemical resistance and microbial degradation of antimycotics. Folia Microbiologica, vol. 16, no. 6, pp. 472-475. PMid:5143694.
  • CAPEK, A., SIMEK, A., BRŮNA, L., SVÁB, A. and BUDĔSÍNSKÝ, Z., 1974. Antimicrobial agents. XXI. Dependence of antifungal activity on the structure of the side chain in the pyrimidine group. Folia Microbiologica, vol. 19, no. 2, pp. 169-171. PMid:4611876.
  • CAVASSIN, F.B., BAÚ-CARNEIRO, J.L., VILAS-BOAS, R.R. and QUEIROZ-TELLES, F., 2021. Sixty years of amphotericin B: an overview of the main antifungal agent used to treat invasive fungal infections. Infectious Diseases and Therapy, vol. 10, no. 1, pp. 115-147. http://dx.doi.org/10.1007/s40121-020-00382-7 PMid:33523419.
    » http://dx.doi.org/10.1007/s40121-020-00382-7
  • CHANG, W., ZHANG, M., LI, Y., LI, X., GAO, Y., XIE, Z. and LOU, H., 2015. Lichen endophyte derived pyridoxatin inactivates Candida growth by interfering with ergosterol biosynthesis. Biochimica et Biophysica Acta, vol. 1850, no. 9, pp. 1762-1771. http://dx.doi.org/10.1016/j.bbagen.2015.05.005 PMid:25960388.
    » http://dx.doi.org/10.1016/j.bbagen.2015.05.005
  • CRUZ, M.C., GOLDSTEIN, A.L., BLANKENSHIP, J., DEL POETA, M., PERFECT, J.R., MCCUSKER, J.H., BENNANI, Y.L., CARDENAS, M.E. and HEITMAN, J., 2001. Rapamycin and less immunosuppressive analogs are toxic to Candida albicans and Cryptococcus neoformans via FKBP12-dependent inhibition of TOR. Antimicrobial Agents and Chemotherapy, vol. 45, no. 11, pp. 3162-3170. http://dx.doi.org/10.1128/AAC.45.11.3162-3170.2001 PMid:11600372.
    » http://dx.doi.org/10.1128/AAC.45.11.3162-3170.2001
  • DEBNATH, A., CALVET, C.M., JENNINGS, G., ZHOU, W., AKSENOV, A., LUTH, M.R., ABAGYAN, R., NES, W.D., MCKERROW, J.H. and PODUST, L.M., 2017. CYP51 is an essential drug target for the treatment of primary amoebic meningoencephalitis (PAM). PLoS Neglected Tropical Diseases, vol. 11, no. 12, e0006104. http://dx.doi.org/10.1371/journal.pntd.0006104 PMid:29284029.
    » http://dx.doi.org/10.1371/journal.pntd.0006104
  • DEORUKHKAR, S.C. and SAINI, S., 2015. Non albicans Candida species: a review of epidemiology, pathogenicity and antifungal resistance. Pravara Medical Review, vol. 7, no. 3, pp. 7-15.
  • DI SANTO, R., 2010. Natural products as antifungal agents against clinically relevant pathogens. Natural Product Reports, vol. 27, no. 7, pp. 1084-1098. http://dx.doi.org/10.1039/b914961a PMid:20485730.
    » http://dx.doi.org/10.1039/b914961a
  • DICK, J.D., MERZ, W.G. and SARAL, R., 1980. Incidence of polyene-resistant yeasts recovered from clinical specimens. Antimicrobial Agents and Chemotherapy, vol. 18, no. 1, pp. 158-163. http://dx.doi.org/10.1128/AAC.18.1.158 PMid:7416742.
    » http://dx.doi.org/10.1128/AAC.18.1.158
  • EL-AZIZI, M., FARAG, N. and KHARDORI, N., 2015. Antifungal activity of amphotericin B and voriconazole against the biofilms and biofilm-dispersed cells of Candida albicans employing a newly developed in vitro pharmacokinetic model. Annals of Clinical Microbiology and Antimicrobials, vol. 14, no. 1, pp. 21. http://dx.doi.org/10.1186/s12941-015-0083-3 PMid:25885806.
    » http://dx.doi.org/10.1186/s12941-015-0083-3
  • ELEJALDE, N.R., MACÍAS, M., CASTILLO, J.C., SORTINO, M., SVETAZ, L., ZACCHINO, S. and PORTILLA, J., 2018. Synthesis and in vitro antifungal evaluation of novel N‐substituted 4‐aryl‐2‐methylimidazoles. ChemistrySelect, vol. 3, no. 18, pp. 5220-5227. http://dx.doi.org/10.1002/slct.201801238
    » http://dx.doi.org/10.1002/slct.201801238
  • ENKLER, L., RICHER, D., MARCHAND, A.L., FERRANDON, D. and JOSSINET, F., 2016. Genome engineering in the yeast pathogen Candida glabrata using the CRISPR-Cas9 system. Scientific Reports, vol. 6, no. 1, pp. 35766. http://dx.doi.org/10.1038/srep35766 PMid:27767081.
    » http://dx.doi.org/10.1038/srep35766
  • EPP, E. (2011). Functional genomics in Candida albicans: tackling drug resistance and morphology Montréal: McGill University, 214 p. Thesis in Philosophy.
  • ESPINEL-INGROFF, A., 2008. Mechanisms of resistance to antifungal agents: yeasts and filamentous fungi. Revista Iberoamericana de Micologia, vol. 25, no. 2, pp. 101-106. http://dx.doi.org/10.1016/S1130-1406(08)70027-5 PMid:18473504.
    » http://dx.doi.org/10.1016/S1130-1406(08)70027-5
  • FEYEREISEN, R., 2015. Insect P450 inhibitors and insecticides: challenges and opportunities. Pest Management Science, vol. 71, no. 6, pp. 793-800. http://dx.doi.org/10.1002/ps.3895 PMid:25404103.
    » http://dx.doi.org/10.1002/ps.3895
  • FIRACATIVE, C., 2020. Invasive fungal disease in humans: are we aware of the real impact? Memorias do Instituto Oswaldo Cruz, vol. 115, e200430. http://dx.doi.org/10.1590/0074-02760200430 PMid:33053052.
    » http://dx.doi.org/10.1590/0074-02760200430
  • FRYBERG, M., OEHLSCHLAGER, A.C. and UNRAU, A.M., 1974. Sterol biosynthesis in antibiotic-resistant yeast: nystatin. Archives of Biochemistry and Biophysics, vol. 160, no. 1, pp. 83-89. http://dx.doi.org/10.1016/S0003-9861(74)80011-1 PMid:4597559.
    » http://dx.doi.org/10.1016/S0003-9861(74)80011-1
  • GINTJEE, T.J., DONNELLEY, M.A. and THOMPSON, G.R., 2020. Aspiring antifungals: review of current antifungal pipeline developments. Journal of Fungi, vol. 6, no. 1, pp. 28. http://dx.doi.org/10.3390/jof6010028 PMid:32106450.
    » http://dx.doi.org/10.3390/jof6010028
  • GOTTLIEB, K., WACHER, V., SLIMAN, J. and PIMENTEL, M., 2016. Inhibition of methanogenic archaea by statins as a targeted management strategy for constipation and related disorders. Alimentary Pharmacology & Therapeutics, vol. 43, no. 2, pp. 197-212. http://dx.doi.org/10.1111/apt.13469 PMid:26559904.
    » http://dx.doi.org/10.1111/apt.13469
  • HITCHCOCK, C. and WHITTLE, P.T. 1993. Chemistry and mode of action of fluconazole. In: J.W. RIPPON and R.A. FROMTLING, eds. Cutaneous antifungal agents: selected compounds in clinical practice and development New York: Marcel Dekker Inc., pp. 183-197.
  • HUANG, T., LI, X., MAIER, M., O’BRIEN-SIMPSON, N.M., HEATH, D.E. and O’CONNOR, A.J., 2023. Using inorganic nanoparticles to fight fungal infections in the antimicrobial resistant era. Acta Biomaterialia, vol. 158, pp. 56-79. http://dx.doi.org/10.1016/j.actbio.2023.01.019 PMid:36640952.
    » http://dx.doi.org/10.1016/j.actbio.2023.01.019
  • JALLOW, S. and GOVENDER, N.P., 2021. Ibrexafungerp: a first-in-class oral triterpenoid glucan synthase inhibitor. Journal of Fungi, vol. 7, no. 3, pp. 163. http://dx.doi.org/10.3390/jof7030163 PMid:33668824.
    » http://dx.doi.org/10.3390/jof7030163
  • JULIEN, G.A., ADIYAGA, W.C., SAAKA, R.A.E., SUNYAZI, S.S., BATUIAMU, A.T., ABUGRI, D. and ABUGRI, J. (2021). Dermatophytic diseases: a review of Tinea pedis medRxiv. In press. http://dx.doi.org/10.1101/2021.06.28.21259664.
  • KAHL, B.C., BECKER, K. and LÖFFLER, B., 2016. Clinical significance and pathogenesis of staphylococcal small colony variants in persistent infections. Clinical Microbiology Reviews, vol. 29, no. 2, pp. 401-427. http://dx.doi.org/10.1128/CMR.00069-15 PMid:26960941.
    » http://dx.doi.org/10.1128/CMR.00069-15
  • KAUR, R., DHAKAD, M.S., GOYAL, R. and KUMAR, R., 2016. Emergence of non-albicans Candida species and antifungal resistance in intensive care unit patients. Asian Pacific Journal of Tropical Biomedicine, vol. 6, no. 5, pp. 455-460. http://dx.doi.org/10.1016/j.apjtb.2015.12.019
    » http://dx.doi.org/10.1016/j.apjtb.2015.12.019
  • KURTZ, M.B. and DOUGLAS, C.M., 1997. Lipopeptide inhibitors of fungal glucan synthase. Journal of Medical and Veterinary Mycology, vol. 35, no. 2, pp. 79-86. http://dx.doi.org/10.1080/02681219780000961 PMid:9147267.
    » http://dx.doi.org/10.1080/02681219780000961
  • MAESAKI, S., HOSSAIN, M.A., MIYAZAKI, Y., TOMONO, K., TASHIRO, T. and KOHNO, S., 2000. Efficacy of FK463, a (1, 3)-β-d-glucan synthase inhibitor, in disseminated azole-resistant Candida albicans infection in mice. Antimicrobial Agents and Chemotherapy, vol. 44, no. 6, pp. 1728-1730. http://dx.doi.org/10.1128/AAC.44.6.1728-1730.2000 PMid:10817741.
    » http://dx.doi.org/10.1128/AAC.44.6.1728-1730.2000
  • MALIGIE, M. and SELITRENNIKOFF, C., 2005. Cryptococcus neoformans resistance to echinocandins: (1,3)β-glucan synthase activity is sensitive to echinocandins. Antimicrobial Agents and Chemotherapy, vol. 49, no. 7, pp. 2851-2856. http://dx.doi.org/10.1128/AAC.49.7.2851-2856.2005 PMid:15980360.
    » http://dx.doi.org/10.1128/AAC.49.7.2851-2856.2005
  • MANOHARAN, A., OGNENOVSKA, S., PAINO, D., WHITELEY, G., GLASBEY, T., KRIEL, F.H., FARRELL, J., MOORE, K.H., MANOS, J. and DAS, T., 2021. N-acetylcysteine protects bladder epithelial cells from bacterial invasion and displays antibiofilm activity against urinary tract bacterial pathogens. Antibiotics, vol. 10, no. 8, pp. 900. http://dx.doi.org/10.3390/antibiotics10080900 PMid:34438950.
    » http://dx.doi.org/10.3390/antibiotics10080900
  • MARCHETTI, O., MOREILLON, P., GLAUSER, M.P., BILLE, J. and SANGLARD, D., 2000. Potent synergism of the combination of fluconazole and cyclosporine in Candida albicans Antimicrobial Agents and Chemotherapy, vol. 44, no. 9, pp. 2373-2381. http://dx.doi.org/10.1128/AAC.44.9.2373-2381.2000 PMid:10952582.
    » http://dx.doi.org/10.1128/AAC.44.9.2373-2381.2000
  • MARTÍNEZ, J.M., DELSO, C., ÁLVAREZ, I. and RASO, J., 2020. Pulsed electric field‐assisted extraction of valuable compounds from microorganisms. Comprehensive Reviews in Food Science and Food Safety, vol. 19, no. 2, pp. 530-552. http://dx.doi.org/10.1111/1541-4337.12512 PMid:33325176.
    » http://dx.doi.org/10.1111/1541-4337.12512
  • MAURICE, N.M., BEDI, B. and SADIKOT, R.T., 2018. Pseudomonas aeruginosa biofilms: host response and clinical implications in lung infections. American Journal of Respiratory Cell and Molecular Biology, vol. 58, no. 4, pp. 428-439. http://dx.doi.org/10.1165/rcmb.2017-0321TR PMid:29372812.
    » http://dx.doi.org/10.1165/rcmb.2017-0321TR
  • MBONGO, N., LOISEAU, P.M., BILLION, M.A. and ROBERT-GERO, M., 1998. Mechanism of amphotericin B resistance in Leishmania donovani promastigotes. Antimicrobial Agents and Chemotherapy, vol. 42, no. 2, pp. 352-357. http://dx.doi.org/10.1128/AAC.42.2.352 PMid:9527785.
    » http://dx.doi.org/10.1128/AAC.42.2.352
  • MEDEIROS, C.I.S., SOUSA, M.N.A., ALMEIDA FILHO, G.G., FREITAS, F.O.R., UCHOA, D.P.L., NOBRE, M.S.C., BEZERRA, A.L.D., ROLIM, L.A.D.M.M., MORAIS, A.M.B., NOGUEIRA, T.B.S.S., NOGUEIRA, R.B.S.S., OLIVEIRA FILHO, A.A. and LIMA, E.O., 2022. Antifungal activity of linalool against fluconazole-resistant clinical strains of vulvovaginal Candida albicans and its predictive mechanism of action. Brazilian Journal of Medical and Biological Research, vol. 55, e11831. http://dx.doi.org/10.1590/1414-431x2022e11831 PMid:35976268.
    » http://dx.doi.org/10.1590/1414-431x2022e11831
  • MEDOFF, G. and KOBAYASHI, G.S., 1980. Strategies in the treatment of systemic fungal infections. The New England Journal of Medicine, vol. 302, no. 3, pp. 145-155. http://dx.doi.org/10.1056/NEJM198001173020304 PMid:6985703.
    » http://dx.doi.org/10.1056/NEJM198001173020304
  • MEIS, J.F., CHOWDHARY, A., RHODES, J.L., FISHER, M.C. and VERWEIJ, P.E., 2016. Clinical implications of globally emerging azole resistance in Aspergillus fumigatus. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, vol. 371, no. 1709, pp. 20150460. http://dx.doi.org/10.1098/rstb.2015.0460 PMid:28080986.
    » http://dx.doi.org/10.1098/rstb.2015.0460
  • MOLZAHN, S.W. and WOODS, R.A., 1972. Polyene resistance and the isolation of sterol mutants in Saccharomyces cerevisiae Journal of General Microbiology, vol. 72, no. 2, pp. 339-348. http://dx.doi.org/10.1099/00221287-72-2-339 PMid:4562308.
    » http://dx.doi.org/10.1099/00221287-72-2-339
  • NAKAGAWA, Y., UMEGAWA, Y., MATSUSHITA, N., YAMAMOTO, T., TSUCHIKAWA, H., HANASHIMA, S., OISHI, T., MATSUMORI, N. and MURATA, M., 2016. The structure of the bimolecular complex between amphotericin B and ergosterol in membranes is stabilized by face-to-face van der Waals interaction with their rigid cyclic cores. Biochemistry, vol. 55, no. 24, pp. 3392-3402. http://dx.doi.org/10.1021/acs.biochem.6b00193 PMid:27227740.
    » http://dx.doi.org/10.1021/acs.biochem.6b00193
  • PARKS, L.W. and CASEY, W.M., 1996. Fungal sterols. In: R. PRASAD and M. GHANNOUM, eds. Lipids of pathogenic fungi Boca Raton: CRC Press, pp. 63-82.
  • PATRA, J.K. and BAEK, K.H., 2017. Green biosynthesis of magnetic iron oxide (Fe3O4) nanoparticles using the aqueous extracts of food processing wastes under photo-catalyzed condition and investigation of their antimicrobial and antioxidant activity. Journal of Photochemistry and Photobiology. B, Biology, vol. 173, pp. 291-300. http://dx.doi.org/10.1016/j.jphotobiol.2017.05.045 PMid:28623821.
    » http://dx.doi.org/10.1016/j.jphotobiol.2017.05.045
  • PEYTON, L.R., GALLAGHER, S. and HASHEMZADEH, M., 2015. Triazole antifungals: a review. Drugs of Today, vol. 51, no. 12, pp. 705-718. http://dx.doi.org/10.1358/dot.2015.51.12.2421058 PMid:26798851.
    » http://dx.doi.org/10.1358/dot.2015.51.12.2421058
  • PFALLER, M., BOYKEN, L., HOLLIS, R.J., KROEGER, J., MESSER, S.A., TENDOLKAR, S. and DIEKEMA, D.J., 2008. In vitro susceptibility of invasive isolates of Candida spp. to anidulafungin, caspofungin, and micafungin: six years of global surveillance. Journal of Clinical Microbiology, vol. 46, no. 1, pp. 150-156. http://dx.doi.org/10.1128/JCM.01901-07 PMid:18032613.
    » http://dx.doi.org/10.1128/JCM.01901-07
  • PODOLSKY, S.H., 2018. The evolving response to antibiotic resistance (1945-2018). Palgrave Communications, vol. 4, no. 1, pp. 124. http://dx.doi.org/10.1057/s41599-018-0181-x
    » http://dx.doi.org/10.1057/s41599-018-0181-x
  • PUKKILA-WORLEY, R., HOLSON, E., WAGNER, F. and MYLONAKIS, E., 2009. Antifungal drug discovery through the study of invertebrate model hosts. Current Medicinal Chemistry, vol. 16, no. 13, pp. 1588-1595. http://dx.doi.org/10.2174/092986709788186237 PMid:19442135.
    » http://dx.doi.org/10.2174/092986709788186237
  • QIN, D. and DONG, J., 2023. Multi-level optimization and strategies in microbial biotransformation of nature products. Molecules, vol. 28, no. 6, pp. 2619. http://dx.doi.org/10.3390/molecules28062619 PMid:36985591.
    » http://dx.doi.org/10.3390/molecules28062619
  • REVIE, N.M., IYER, K.R., ROBBINS, N. and COWEN, L.E., 2018. Antifungal drug resistance: evolution, mechanisms and impact. Current Opinion in Microbiology, vol. 45, pp. 70-76. http://dx.doi.org/10.1016/j.mib.2018.02.005 PMid:29547801.
    » http://dx.doi.org/10.1016/j.mib.2018.02.005
  • RUGELES, C.M., 2003. Role of the efflux pumps CaMDR1, CDR1 and CDR2 in the fluconazole susceptibility of primary adhered Candida albicans Atlanta: Georgia State University. 147 p. Thesis in Biology Microbiology.
  • SHAFIEI, M., PEYTON, L., HASHEMZADEH, M. and FOROUMADI, A., 2020. History of the development of antifungal azoles: A review on structures, SAR, and mechanism of action. Bioorganic Chemistry, vol. 104, pp. 104240. http://dx.doi.org/10.1016/j.bioorg.2020.104240 PMid:32906036.
    » http://dx.doi.org/10.1016/j.bioorg.2020.104240
  • SNEHA, K.S., NATARAJAN, S., BOAZ, K., RAMAPURAM, J., BALIGA, S., MANAKTALA, N. and CHITRA, N.S., 2023. Candida profile in HIV-positive children needs a dynamic clinical appraisal: a microbiological study. Research Journal of Pharmacy and Technology, vol. 16, no. 1, pp. 423-428.
  • SOUNDARYA, R., HALAGALI, P., PREETHI, S., VIKRAM, H.P., MEHDI, S. and SINGADI, R., 2023. Quality By Design (QBD) approach in processing of nanoparticles loading antifungal drugs. Journal of Coastal Life Medicine, vol. 11, pp. 279-290.
  • SUGIO, C.Y.C., GARCIA, A.A.M.N., ALBACH, T., MORAES, G.S., BONFANTE, E.A., URBAN, V.M. and NEPPELENBROEK, K.H., 2020. Candida-associated denture stomatitis and murine models: what is the importance and scientific evidence? Journal of Fungi, vol. 6, no. 2, pp. 70. http://dx.doi.org/10.3390/jof6020070 PMid:32456172.
    » http://dx.doi.org/10.3390/jof6020070
  • VANDEN BOSSCHE, H., MARICHAL, H.P., ODDS, F., LE JEUNE, L. and COENE, M.-C., 1992. Characterization of an azole-resistant Candida glabrata isolate. Antimicrobial Agents and Chemotherapy, vol. 36, no. 12, pp. 2602-2610. http://dx.doi.org/10.1128/AAC.36.12.2602 PMid:1482129.
    » http://dx.doi.org/10.1128/AAC.36.12.2602
  • VANREPPELEN, G., WUYTS, J., VAN DIJCK, P. and VANDECRUYS, P., 2023. Sources of antifungal drugs. Journal of Fungi, vol. 9, no. 2, pp. 171. http://dx.doi.org/10.3390/jof9020171 PMid:36836286.
    » http://dx.doi.org/10.3390/jof9020171
  • WANG, W.Y., CAI, H.Q., QU, S.Y., LIN, W.H., LIANG, C.C., LIU, H., XIE, Z.X. and YUAN, Y.J., 2022. Genomic variation-mediating fluconazole resistance in yeast. Biomolecules, vol. 12, no. 6, pp. 845. http://dx.doi.org/10.3390/biom12060845 PMid:35740970.
    » http://dx.doi.org/10.3390/biom12060845
  • WHALEY, S.G., BERKOW, E.L., RYBAK, J.M., NISHIMOTO, A.T., BARKER, K.S. and ROGERS, P.D., 2017. Azole antifungal resistance in Candida albicans and emerging non-albicans Candida species. Frontiers in Microbiology, vol. 7, pp. 2173. http://dx.doi.org/10.3389/fmicb.2016.02173 PMid:28127295.
    » http://dx.doi.org/10.3389/fmicb.2016.02173
  • WIEDERHOLD, N.P. and PATTERSON, T.F., 2015. Emergence of azole resistance in Aspergillus Seminars in Respiratory and Critical Care Medicine, vol. 36, no. 5, pp. 673-680. http://dx.doi.org/10.1055/s-0035-1562894 PMid:26398534.
    » http://dx.doi.org/10.1055/s-0035-1562894
  • YANG, H., TONG, J., LEE, C.W., HA, S., EOM, S.H. and IM, Y.J., 2015. Structural mechanism of ergosterol regulation by fungal sterol transcription factor Upc2. Nature Communications, vol. 6, no. 1, pp. 6129. http://dx.doi.org/10.1038/ncomms7129 PMid:25655993.
    » http://dx.doi.org/10.1038/ncomms7129

Publication Dates

  • Publication in this collection
    23 Feb 2024
  • Date of issue
    2023

History

  • Received
    24 May 2023
  • Accepted
    28 Sept 2023
Instituto Internacional de Ecologia R. Bento Carlos, 750, 13560-660 São Carlos SP - Brasil, Tel. e Fax: (55 16) 3362-5400 - São Carlos - SP - Brazil
E-mail: bjb@bjb.com.br