8,839
Views
21
CrossRef citations to date
0
Altmetric
Review

Echinocandins – structure, mechanism of action and use in antifungal therapy

ORCID Icon, ORCID Icon, ORCID Icon, ORCID Icon & ORCID Icon
Pages 876-894 | Received 31 Jul 2021, Accepted 02 Mar 2022, Published online: 16 Mar 2022

Abstract

With increasing number of immunocompromised patients as well as drug resistance in fungi, the risk of fatal fungal infections in humans increases as well. The action of echinocandins is based on the inhibition of β-(1,3)-d-glucan synthesis that builds the fungal cell wall. Caspofungin, micafungin, anidulafungin and rezafungin are semi-synthetic cyclic lipopeptides. Their specific chemical structure possess a potential to obtain novel derivatives with better pharmacological properties resulting in more effective treatment, especially in infections caused by Candida and Aspergillus species. In this review we summarise information about echinocandins with closer look on their chemical structure, mechanism of action, drug resistance and usage in clinical practice. We also introduce actual trends in modification of this antifungals as well as new methods of their administration, and additional use in viral and bacterial infections.

Introduction

Diseases caused by fungi are a serious problem. Currently, the number of people affected by fungal infections worldwide exceeds one billionCitation1, and the number of deaths caused by invasive fungal species is comparable to the mortality of tuberculosis − 1.5 million each yearCitation2. About 90% of mortalities are caused by fungi grouped into four genera: Candida, Aspergillus, Cryptococcus, and PneumocystisCitation3 (). There are at least 17 pathogenic species complexes in the genus Candida, but more than 90% of infections are attributed to Candida albicans, Candida glabrata, Candida parapsilosis, Candida tropicalis, and Candida kruseiCitation4. The patients suffering from AIDS, cancer (and associated chemotherapy), leukaemia, patients on immune-compromising drug therapy, and after organ transplantation are particularly vulnerable to mycosesCitation5.

Table 1. The estimated global annual number of selected fungal infectionsCitation1,Citation3.

Until the end of the 20th century, azoles, polyenes, and flucytosine were mainly used to treat mycoses. These classes of drugs can cause serious side effects related to their hepato- and nephrotoxicity. In addition, many fungal strains have developed resistance to these antibiotics, which significantly reduces their efficacyCitation6. In rare cases, cross-resistance to polyenes and azoles may occur, raising concerns about the future of antifungals targeting membrane ergosterol (polyens) and sterol synthesis (azoles)Citation7. In addition, there are frequent drug-drug reactions associated with interactions of the aforementioned drugs and their metabolites in the bodyCitation8. This has prompted the search for alternative agents to combat fungal infections. Echinocandins are a class of antifungal drugs that are fungicidal against many fungi including Candida species, but are fungistatic to the Aspergillus genus. This class of drugs has been found to cause milder side effects compared to polyenes and azolesCitation9. The mechanism of action based on the inhibition of fungal-specific metabolic pathway and limited side effects have resulted in increasing interest and use of this class of drugsCitation10.

The purpose of this paper is to introduce echinocandins as antifungal antibiotics for the reader, details of their chemical structure and proposed modifications, mechanism of action as well as their usage in clinical practice are also described. The use of echinocandins has improved patient outcome; however, there are reports about drug resistance of fungi to this class of antibiotics. Therefore, an additional aim of this study is to summarise the latest information on the phenomenon of fungal drug resistance to echinocandins.

History of the discovery of echinocandins

In 1974, in Switzerland the first antifungal drug of the echinocandin class – echinocandin B was discovered, which showed good antifungal properties but at the same time strong haemolytic effects. To counteract this, cilofungin, a semi-synthetic analogue of echinocandin B with a 4-octyloxybenzoate side chain, was synthesised. This compound significantly reduced haemolytic activity while retaining antifungal propertiesCitation11. However, cilofungin was withdrawn from Phase II clinical trials due to the poor water solubility and toxicity of its co-solventCitation12. An important step in echinocandin research was the discovery of pneumocandin A0 and pneumocandin B0, of which pneumocandin B0 was used to synthesise a new antifungal agentCitation13 ().

Table 2. Names and physical properties of selected echinocandinsCitation12,Citation14–16.

In 1992, caspofungin acetate was first synthesised from pneumocandin B0 and approved for clinical trialsCitation13. The U.S. Food and Drug Administration (FDA) approved caspofungin in January 2001 as a drug for the prevention of fungal infections in adult patients, and the first medical preparations containing caspofungin were introduced to the U.S. market the same year. In Europe, the antibiotic was approved for therapy in 2002 under the trade name CANCIDAS (manufacturer: Merck & Co. Inc., USA, MK-0991)Citation17. In July 2008, it was approved for the treatment of children over 3 months of ageCitation18.

The precursor of micafungin, which was FR901379, was discovered in Japan by Fujisawa Pharmaceutical. To stop FR901379-induced reticulocyte lysis, it underwent various modifications. This resulted in the compound FK463 called micafungin, which had reduced haemolytic activity and was potent against Candida and Aspergillus speciesCitation19. The drug was approved in the U.S. in 2005, and in Europe in 2008. The compound was also approved for the treatment of invasive candidiasis in patients younger than 4 months of ageCitation20. The first formulations of micafungin were produced in Japan and sold under the trade name MYCAMINE (manufacturer: Astellas Pharma Inc., Japan, FK-463)Citation17.

Another of the echinocandins, anidulafungin was obtained by optimising the chemical structure of echinocandin B. Replacement of the naturally occurring fatty acid side chain with an alkoxytriphenyl side chain reduced the haemolytic effect of echinocandin BCitation17. Anidulafungin, as an antifungal compound for the treatment of oesophageal candidiasis, candidemia and deep tissue candidiasis, was approved in the US in 2006 and marketed under the trade name ERAXIS. In Europe, the drug was marketed a year later under the trade name ECALTA (manufacturer: Pfizer Inc. UK)Citation21.

The "ReSTORE" study (ClinicalTrials.gov registration number NCT03667690)Citation16 of rezafungin (CD101), the newest member of the echinocandin class, is currently ongoing. Two post-study Phase III clinical trials are underway to evaluate the efficacy of rezafungin for the treatment of candidemia and invasive candidiasis and for the prevention of invasive fungal infections caused by Candida, Aspergillus and Pneumocystis species (patient recruitment for the trials began in October 2018, with planned completion of the trial set for August 2021)Citation22. Rezafungin is a structural analogue of anidulafungin in which the hemiaminal region at the C5 ornithine position has been replaced with a choline amine ether. With this modification, the stability of the drug in solutions and pharmacokinetics are improved, as well as the drug's action in the body is prolongedCitation10. It is also possible to treat infections caused by fungal strains resistant to other antibiotics such as azoles with rezafunginCitation23.

Chemical structure of selected echinocandins

Echinocandins such as pneumocandin A0 and B0, echinocandin B, and FR901379 are naturally synthesised by filamentous fungi. Caspofungin, micafungin, anidulafungin and rezafungin are semi-synthetic cyclic lipopeptides with antifungal activity (). They have a core composed of a cyclic hexapeptide and are acylated with a different fatty acid attached to the α-amino group of dihydroxyornithine (). This lipid residue is required to anchor the drug to the cell membrane and is therefore essential for bioactivityCitation11.

Figure 1. Echinocandin precursors and their structural modifications resulting in the three representatives approved for clinical use (bold boxes), rezafungin is under phase III clinical trials.

Figure 1. Echinocandin precursors and their structural modifications resulting in the three representatives approved for clinical use (bold boxes), rezafungin is under phase III clinical trials.

Figure 2. Structure of selected compounds of the echinocandin class. Thin boxes indicate the natural product precursors derived from fungal metabolism, bold boxes distinguish semi-synthetic antibiotics.

Figure 2. Structure of selected compounds of the echinocandin class. Thin boxes indicate the natural product precursors derived from fungal metabolism, bold boxes distinguish semi-synthetic antibiotics.

Echinocandin B, pneumocandin B0 and synthesis of caspofungin

Echinocandin B () is a major lipopeptide antifungal antibiotic from a complex produced by Aspergillus nidulans and Aspergillus rugulosus. It contains a linoleic acid side chain and a hexacyclic peptide core built by various amino acid residues such as 3,4-dihydroxyhomotyrosine, 3-hydroxy-4-methylproline, 4,5-dihydroxyornithine, 4-hydroxyproline and two threonine residuesCitation24. These amino acid residues have a significant effect on antifungal activity and determine the physicochemical properties of the echinocandin B nucleus. For example, modified proline and homotyrosine residues are essential for the antifungal efficacy of echinocandinsCitation25. Hydroxyl groups at the three amino acid residues forming the cyclic lipopeptide core, improve the solubility of the drug in water and aid its stability in solutionsCitation12. Taking this into account, during the synthesis of newer echinocandin B derivatives, the core was kept unchanged or underwent minor modificationsCitation26 (). The hydrophobic fatty acid chain attached to the echinocandin B core is crucial for antifungal activity because it acts as a "hook" that allows the drug to anchor in the fungal cell membraneCitation12.

Caspofungin is a cyclic, semisynthetic water-soluble lipopeptide (). This compound is a derivative of the naturally occurring hexapeptide in Glarea lozoyensis, pneumocandin B0Citation26,Citation28. This compound was developed as a result of a four-year program of medicinal chemistry to obtained a derivative with improved water solubility that would facilitate the development of an intravenous formulation. Initial attempts to optimise the structure produced derivatives with a cationic aminoethyl ether group, as well as a 3-hydroxyornithine in place of the 3-hydroxyglutamine moiety (). These modifications increased water solubility, stability as well as activity against Candida and Aspergillus spp.Citation27. Initial studies had suggested that improved activity was caused by the presence of cationic groups that can form ion pairs with the negatively charged phosphate group of the phospholipid. The consequence was an increase in the concentration of lipopeptide in the cell membrane where the target, glucan synthase, is located. Replacement of the aminoethyl ether with ethylene diamine resulted in the identification of caspofungin, which was synthesised via a two-step modification of the peptide core of pneumocandin B0. First carboxamide of 3-hydroxyglutamine was selectively reduced to an amine by a two-step method. Second, condensation of the hemiaminal group with ethylenediamine furnished caspofunginCitation29 (). After improving the methodology, caspofungin acetate – CANCIDAS® was obtained by a very efficient three-step synthesisCitation29 (). Despite the success of the medicinal chemistry program, new analogues of caspofungin are still under development. To study echinocandins cellular biology in Candida species, some fluorescent caspofungin’s derivatives were developedCitation30,Citation31 (). It should be emphasised that some of new caspofungin derivatives do not show biological activity (such as, for example, derivatives obtained by cross-metathesis with a larger peptide ring size than in naturally occurring echinocandins)Citation32 (). The possibility of designing macrocyclic compounds by total synthesis is very promising. The use of mild conditions in efficient reactions is an interesting alternative to the semi-synthetic method. The application of total synthesis enables to conveniently plan a wide variety of compounds and to study the relationship between structure and biological activityCitation34.

Figure 3. Medicinal chemistry progression from pneumocandin B0 to caspofungin. The differences in the structure of pneumocandin B0 that lead to caspofungin are marked in green circles as: (a) aminoethyl ether, (b) 3-hydroxyglutamine reduction site, (c) ethylenediamine (according toCitation13,Citation27).

Figure 3. Medicinal chemistry progression from pneumocandin B0 to caspofungin. The differences in the structure of pneumocandin B0 that lead to caspofungin are marked in green circles as: (a) aminoethyl ether, (b) 3-hydroxyglutamine reduction site, (c) ethylenediamine (according toCitation13,Citation27).

Figure 4. Synthesis of caspofungin (according toCitation29).

Figure 4. Synthesis of caspofungin (according toCitation29).

Figure 5. Improved synthesis of caspofungin acetate - CANCIDAS® (according toCitation29).

Figure 5. Improved synthesis of caspofungin acetate - CANCIDAS® (according toCitation29).

Figure 6. Fluorescent derivatives of caspofungin. Structural elements changed in the structure of caspofungin are shown in green boxes (according toCitation14,Citation32).

Figure 6. Fluorescent derivatives of caspofungin. Structural elements changed in the structure of caspofungin are shown in green boxes (according toCitation14,Citation32).

Figure 7. Completely synthetic caspofungin derivatives. The differences in the structure of new compounds relative to caspofungin are shown in green (II., III., IV. according toCitation25; I. according toCitation32; V. according toCitation33).

Figure 7. Completely synthetic caspofungin derivatives. The differences in the structure of new compounds relative to caspofungin are shown in green (II., III., IV. according toCitation25; I. according toCitation32; V. according toCitation33).

Synthesis of micafungin

Another group of echinocandins are compounds with a sulphate moiety instead of a hydroxyl group in dihydroxyhomotyrosine. Micafungin () was formed by enzymatic deacylation of a naturally occurring hexapeptide derived from Coleophoma empetri (FR901379), to which an optimised N-acylated side chain was then introducedCitation35. The side chain contains an isoxazole ring substituted by 3,5-diphenylCitation36. This modification reduces the haemolytic activity of the drug compared to FR901379 while retaining the antifungal properties of the precursorCitation19. The sulphate group in the dihydroxyhomotyrosine side chain in the compound structure increase the water solubility of micafunginCitation37.

Figure 8. Structure comparison of FR901379 (I) and micafungin (II). The side chain of the isoxazole ring that distinguishes micafungin from FR901379 is shown in green circle. The sulphate group responsible for increasing water solubility of both compounds is highlighted in red.

Figure 8. Structure comparison of FR901379 (I) and micafungin (II). The side chain of the isoxazole ring that distinguishes micafungin from FR901379 is shown in green circle. The sulphate group responsible for increasing water solubility of both compounds is highlighted in red.

Modification of echinocandin B – anidulafungin and rezafungin

Anidulafungin is a semi-synthetic derivative of echinocandin BCitation9 (). Anidulafungin, as well as other echinocandins, consists of a peptide nucleus of echinocandin B, composed of amino acid residues. Enzymatical deacylation of echinocandin B using Actinoplanes utahensis culture introduced hydrochloride salt in place of the linoleoyl side chain. Next, this HCl salt was reacted with the activated ester to form anidulafunginCitation38 (). The introduction of an alkoxytriphenyl side chain in place of the alkyl chain of echinocandin B reduces the haemolytic properties of the drug and has a key effect on the intercalation of anidulafungin with the fungal cell membrane, but reduces the solubility of the drug in waterCitation12,Citation38. Poor solubility and low oral bioavailability of anidulafungin cause a necessity of administration by intravenous injectionCitation40,Citation41. To increase the solubility, derivatives with a modified side chain have recently been prepared, and one of them is particularly promisingCitation42 (). According to primary tests, it shows high activity against C. albicans and C. krusei with better water solubility and lower toxicity in vitro in murine macrophages (RAW264.7) than anidulafunginCitation41. The observed effect may be particularly important considering the therapy of immunocompromised patients.

Figure 9. Structure comparison of echinocandin B (I) and anidulafungin (II). The alkoxytriphenyl side chain that distinguishes the structure of anidulafungin from echinocandin B is shown in green circle.

Figure 9. Structure comparison of echinocandin B (I) and anidulafungin (II). The alkoxytriphenyl side chain that distinguishes the structure of anidulafungin from echinocandin B is shown in green circle.

Figure 10. Synthesis of anidulafungin (according toCitation39).

Figure 10. Synthesis of anidulafungin (according toCitation39).

Figure 11. Structure of new semi-synthetic derivative of anidulafungin (according toCitation42).

Figure 11. Structure of new semi-synthetic derivative of anidulafungin (according toCitation42).

Anidulafungin is the precursor of rezafungin (CD101) with an additional choline etherCitation15 (). The choline amine ether moiety at the C5 position of ornithine increased the stability of the compound (reduced degradation of the compound in the hemiaminal region), half-life, solubilityCitation10,Citation22,Citation23. During degradation processes, anidulafungin may undergo a ring-opening process in which the hemiaminal cleaves to form a linear peptide having a terminal amide and aldehyde from the hydroxyl C5 ornithine residue forming reactive intermediates that undergo further metabolism in the body and persist until eliminated in the faeces. The increased chemical stability of rezafungin prevents ring opening resulting in the absence of reactive metabolites that may contribute to toxicityCitation15,Citation43.

Figure 12. Structure comparison of anidulafungin (I) and rezafungin (II). Choline amine ether at the C5 ornithine position distinguishing the two compounds is shown in green circle.

Figure 12. Structure comparison of anidulafungin (I) and rezafungin (II). Choline amine ether at the C5 ornithine position distinguishing the two compounds is shown in green circle.

Mechanism of echinocandin action

Fungal cell wall components include β-(1,3)-d-glucans, β-(1,4)-d-glucans, β-(1,6)-d-glucans, α-glucans, chitin, mannan, and a variety of glycoproteinsCitation42. Glucans are particularly important components in maintaining the cell wall integrity of Candida spp. and Saccharomyces spp. accounting for approximately 50–60% of the cell wall components of these fungi. Besides being an important component of fungal cell wall structure, β-(1,3)-d-glucan is not found in animal cells, so its synthesis is a good target for antifungal antibioticsCitation44.

The synthesis of β-(1,3)-d-glucan is catalysed by UDP-glucose (1,3)-d-glucan-β-(3)-d-glucosyltransferase, referred to as β-(1,3)-d-glucan synthase (EC 2.4.1.34)Citation45. This enzyme uses UDP-glucose as a reaction substrate to form β-(1,3)-d-glycosidic bondsCitation46. The enzyme is a transmembrane heteromeric glycosyltransferase consisting of at least two subunits. The Fks1p subunit (encoded by the FKS1, FKS2, and FKS3 genes) has a catalytic function, while the Rho1p subunit (belonging to the GTPase family) has a regulatory function. Echinocandins binding non-competitively to the Fks1p subunit of the enzyme inhibits its activityCitation47,Citation48. Blocking β-(1,3)-d-glucan biosynthesis leads to structural abnormalities of the fungal cell wall (), resulting in growth inhibition or death by imbalance in osmotic pressureCitation49. The fungicidal or fungistatic effects of echinocandins have been confirmed for most species of the Candida and Aspergillus generaCitation50.

Figure 13. Mechanism of action of echinocandins. (A) normal production of β-(1,3)-d-glucan, (B) echinocandins acting on the FKS1p subunit non-competitively inhibit glucan synthase activity disrupting β-(1,3)-d-glucan synthesis, leading to fungal cell death caused by cell wall instability.

Figure 13. Mechanism of action of echinocandins. (A) normal production of β-(1,3)-d-glucan, (B) echinocandins acting on the FKS1p subunit non-competitively inhibit glucan synthase activity disrupting β-(1,3)-d-glucan synthesis, leading to fungal cell death caused by cell wall instability.

Antifungal activity, metabolism and medicinal use of echinocandins

The range of MIC values of echinocandins varies from 0.007 μg/ml to 32 μg/ml depending on the Candida species. Overall, anidulafungin shows the most potency against most Candida pathogensCitation51 ().

Table 3. Range of MIC and MEC values (μg/ml) of echinocandins against selected Candida and Aspergillus speciesCitation52–55.

Echinocandins are recommended as a treatment for patients suffering from an invasive infection caused mainly by Candida, Aspergillus species and some other pathogenic fungiCitation50 (). These antibiotics also act on biofilm-forming yeasts especially on the Candida genusCitation64. Relative to Candida species, echinocandins exhibit fungicidal activity manifested by significant cell enlargement and distortion, which contributes to inhibition of cell proliferation. Against Aspergillus species (A. fumigatus, A. flavus, A. niger, and A. terreus), echinocandins exert fungistatic effects by causing irregular growth of the hyphae with multiple branched tips and distended cells, preventing the pathogen from spreading beyond the initial site of infectionCitation65–67. Echinocandins are also active against some species of Penicillium and Paecilomyces. To a lesser extent, they show activity against Madurella, Wangiella, Sporothrix, Exophiala, Scedosporium, Pseudallescheria and Fonsecaea generaCitation68. These antibiotics used without additional antifungal compounds are not effective for the treatment of mycoses caused by Mucorales, Cryptococcus, Fusarium, Rhizpous and Trichosporon generaCitation69,Citation70. The cell wall of the above-mentioned fungal genera contains mainly β-(1,6)-d-glucans, which limits their sensitivity to echinocandinsCitation69. The activity of these antibiotics against representatives of the genera Histoplasma, Blastocystis and Coccidioides is also limitedCitation71.

Table 4. Susceptibility of selected fungal species to different antifungal drugsCitation38,Citation56–63.

Echinocandins are given to patients intravenously (over 1 h infusion), because they are poorly absorbed in the gastrointestinal tract (less than 3% of the antibiotic is absorbed after oral administration). They also cannot penetrate the central nervous system or the eyes, and for this reason are not used to treat intraocular inflammation or fungal meningitisCitation72. These antibiotics bind strongly to proteins (97–99%), leading to lower concentrations of the drug available in serum and tissuesCitation36,Citation59. The half-life ranges from 9 to 133 h (), so they are administered once daily (for FDA approved ones) or once weekly (rezafungin). The clinical response of the patient affects the length of treatment. Most patients should be treated 14 days or no more than 7 days after the resolution of symptomsCitation5. Echinocandins are metabolised mostly in the liver as well as in the adrenal glands and spleen through hydrolysis and N-acetylation. Their metabolites are excreted mainly in the bile and faecesCitation18,Citation36. Echinocandins are found in the highest concentrations in the liver, spleen, intestine, kidney, and lung. In these tissues, their concentrations can be two to sixteen times higher than in plasmaCitation79. Echinocandins do not affect the P-glycoprotein family and are weak substrates for cytochrome P450Citation72.

Table 5. Pharmacokinetics of echinocandins for adult patientsCitation18,Citation48,Citation54,Citation60,Citation73–78.

The administered dose in geriatric patients does not require appropriate adjustment. The elderly do not exhibit changes in pharmacokinetics in comparison to younger patients (dose adjustment is not required)Citation80. Echinocandins are not removed renally, so they can be safely administered to elderly patients with renal impairment and patients undergoing dialysis. If a patient is taking several pharmaceuticals during echinocandins therapy, it is not necessary to adjust these medications because echinocandins have little drug-drug interaction. The exception is the special dosage adjustment of caspofungin in patients with hepatic impairmentCitation80,Citation81.

Caspofungin

Caspofungin besides many Candida and Aspergillus species, also has activity against Pneumocystis jirovecii. Its activity against C. parapsilosis and C. guilliermondii, as well as Trichosporon beigelii, Rhizopus arrhizus and Fusarium spp. is moderateCitation17. Studies have also shown antibacterial activity of caspofungin against Enterococcus faeciumCitation82. Caspofungin has nonlinear pharmacokinetics, has an affinity for plasma proteins (97%) and has a half-life of 9–11 h in the human body (), allowing for once-daily dosing. The drug is metabolised in the liver by hydrolysis and N-acylation, and the metabolites are excreted mainly in the faeces (35%) and urine (41%)Citation72,Citation83. Some caspofungin undergoes spontaneous chemical degradation caused by peptide ring-openingCitation84. A small amount of caspofungin is excreted unchanged in the urine (2%)Citation83. In clinical practice, caspofungin is used in patients with invasive aspergillosis (caused by A. fumigatus) and associated neutropenia, as well as in cancer patients, HIV-infected patients, allogeneic haematopoietic stem cell transplant patients, and organ transplant patients, for preventing of fungal infectionsCitation85,Citation86. It may also be used to treat pleural infections, candidemia, oesophageal candidiasis, peritonitis, intra-abdominal abscesses, and abdominal infections caused by Candida speciesCitation13. Caspofungin is used for treatment against voriconazole- and polyene-resistant Aspergillus fumigatusCitation9,Citation49. The recommended dosing regimen in adults consists of a single saturating dose of 70 mg on the first day, followed by 50 mg daily, administered over 1 h, reaching an initial therapeutic plasma concentration of 1 μg/mlCitation87,Citation88 (). Paediatric patients receive a saturating dose of 70 mg and a maintenance dose of 50 mgCitation14,Citation72. Caspofungin may reduce concentrations of rifampicin and tacrolimus in human serumCitation88. In the absence of treatment response or concomitant use of cytochrome P450 inducing drugs (e.g. rifampin, efavirenz, dexamethasone), the daily dose of caspofungin should be increased to 70 mgCitation18. In patients with hypoalbuminemia and liver failure, caspofungin doses should be individually adjustedCitation91. Inhalation administration of the drug is also being studied, which according to data from experiments shows better antifungal properties. Additionally, the inhaled form can be used once a week and has lower hepatotoxicity than intravenous form administered once at high concentrationsCitation92. It is possible to treat animals such as cats suffering from invasive fungal mucositis and sinusitis caused by Aspergillus fumigatus with caspofunginCitation93.

Table 6. Characteristics of echinocandins as medical preparationsCitation5,Citation10,Citation17,Citation18,Citation22,Citation51,Citation89,Citation90.

Micafungin

Micafungin has potent antifungal activity against a broad spectrum of Candida species and azole-resistant Aspergillus speciesCitation19. It binds significantly (>99%) to plasma proteins, mainly albumin and, to a lesser extent, α1-acid glycoprotein, and has a half-life of 11 to 17 h (). Micafungin is not a substrate for P-glycoprotein and does not affect its activity. It is metabolised by arylsulfatase, catechol-O-methyltransferase, and several cytochrome P450 (CYP) isoenzymes (3A4, 1A2, 2B6, and 2 C). It degrades to at least eleven metabolites that are excreted in the bile over many days. Human organism also disposes of them in the urine. Less than one percent of the drug is excreted unchanged via this routeCitation94,Citation95. Micafungin is used to treat patients suffering from invasive candidiasis, candidemia, oesophageal candidiasis, abdominal abscesses, and peritonitis. It is used in the prophylaxis of Candida infections in patients undergoing haematopoietic stem cell transplantation or when a patient is expected to have granulocytopeniaCitation96. Studies show that it is possible to use micafungin as alternative antifungal prophylaxis in patients with acute leukaemia and myelodysplastic syndromeCitation97. Micafungin also affects the immune system by potentiating its response by improving human macrophage activationCitation98. A saturating dose is not required; doses of 100–150 mg daily provide minimum plasma concentrations of approximately 2–2.5 µg/ml on the first day of therapyCitation95. Dosing in adults is 150 mg for oesophageal candidiasis, 100 mg for invasive candidiasis, and 50 mg for prevention of Candida infection (). In paediatric patients, the administered dose ranges from 1 to 4 mg/kg body weight depending on the disease presentCitation53.

Anidulafungin

Anidulafungin is active against a wide range of Candida species, including those resistant to azoles, amphotericin B, and other echinocandins. It is less active against C. guilliermondii and C. parapsilosisCitation99. Out of the echinocandins used, anidulafungin has the greatest activity against species of the genus Aspergillus even in the case of Aspergillus lentulus, which has reduced sensitivity to most antifungal drugsCitation100. This antibiotic is highly bound to plasma proteins (99%) and has a half-life of 24–26 h in the human body (). Anidulafungin is non-enzymatically metabolised in human plasma during biotransformation processes to an open ring peptide lacking antifungal activity, which is excreted mainly in the faecesCitation99,Citation101. Since the degradation of the drug does not occur in the liver and the metabolites are mainly excreted biliary, the compound is safe for the treatment of individuals with liver and kidney failureCitation102. As an antifungal antibiotic, anidulafungin is used to treat oesophageal candidiasis, candidemia, abdominal abscesses, and peritonitis in patients with or without neutropeniaCitation103–105. It is possible to treat pneumonia caused by Pneumocystis jirovecii with anidulafungin in patients who cannot tolerate trimethoprim or sulfamethoxazoleCitation106. Adult patients with invasive candidiasis receive a dose of 100 mg of anidulafungin per day after a 200 mg saturating dose, and for oesophageal candidiasis, 50 mg per day after a 100 mg saturating dose (). Infusion of the saturating dose should last approximately 3 h and the maintenance dose approximately 1.5 hCitation99. Altered clearance and volume of distribution have been noted in patients weighing more than 140 kg which may result in less exposure to echinocandins in infected tissues. In that cases it is recommended to increase the antibiotic dose by 25% compared to the standard amountCitation95. The plasma concentration of anidulafungin is maintained below 1 mg/l throughout the dosing period. Paediatric patients receive a saturating dose of 3 mg/kg body weight followed by a maintenance dose of 1.5 mg/kg body weightCitation5,Citation21. Liposome formulations of anidulafungin show greater efficacy compared to administration of the drug in free form. Such formulations could potentially (so far not approved by FDA) support treatment and reduce the occurrence of drug resistanceCitation107.

Rezafungin

Rezafungin is the first representative of the second generation class of echinocandinsCitation10. It is effective against Candida spp., Aspergillus spp., Trichophyton mentagrophytes, Trichophyton rubrum, and Microsporum gypseumCitation50. Rezafungin is also effective against isolates in which there is confirmed resistance to other echinocandins or azoles (C. auris, C. parapsilosis, C. glabrata)Citation48,Citation73. Rezafungin is currently in Phase III of clinical trials in patients with candidemia and invasive candidiasisCitation48. The antibiotic is also effective in preventing Pneumocystis spp. pneumoniaCitation23. A study in a mouse model showed that 3 weeks of rezafungin prophylaxis was as effective as the standard treatment of pneumocystis pneumonia with trimethoprim or sulfamethoxazoleCitation22. Rezafungin also shows very potent activity against A. fumigatus, even against cryptic multidrug-resistant strains carrying a mutation in the CYP51A gene (causing azole resistance) which may suggest an effective treatment for azole-resistant aspergillosisCitation10,Citation108. Rezafungin binds strongly to proteins (99%), has a long half-life (80 h after the first dose, after subsequent dose 150 h), and its high safety profile allows for high doses (400 mg first dose, subsequent 200 mg) once a weekCitation22,Citation23,Citation108,Citation109 (). The pharmacokinetics of rezafungin are relatively linear regardless of dose and its prolonged action enhanced drug penetration into infected tissuesCitation22,Citation110. Rezafungin has a better safety profile compared to other echinocandinsCitation10. The solubility in water is above 150 mg/ml and in acetate, lactate and tris buffers is 45 mg/mlCitation15. Rezafungin is rapidly distributed to the kidney and liver and its urinary excretion is negligible. It shows better distribution to hepatic abscess areas compared to micafungin (the concentration for rezafungin was 44.5 µg/g and for micafungin was 3.4 µg/g 24 h after drug administration)Citation111. Rezafungin at a concentration of 0.25 µg/ml inhibits biofilm formation in C. albicans by deformation of the cells and release of their contentsCitation10,Citation112.

Lyophilised rezafungin powder shows much less degradation during storage compared to other echinocandins. Storage of rezafungin at 40 °C, at room temperature in 5% dextrose, 0.9% NaCl solution or sterile water for more than 1 year showed minimal degradation of the compound (less than 7%) and no epimerizationCitation10,Citation15,Citation23. Infusion solutions can be stored without stabilisers and without fear of photolysis or spontaneous degradationCitation15. Rezafungin can potentially be topical and subcutaneously appliedCitation10,Citation113.

Synergistic action of echinocandins with other antibiotics

Combinations of echinocandins along with other antifungal antibiotics are currently being investigated. Results of recent studies indicate an additive effect of caspofungin and voriconazole against an echinocandin-resistant strain of C. glabrata. Moreover, using the checkerboard test, combinations of caspofungin with azoles and amphotericin B showed an increase in fungicidal effect from 17.65% to 29.41% against this speciesCitation114. The combination of anidulafungin and isavuconazole increases efficacy against azole-resistant A. fumigatus, which may benefit patients with invasive aspergillosis when azole therapy does not workCitation115. Nikkomycin Z, along with caspofungin or micafungin, has shown enhanced activity against biofilms produced by C. albicans and C. parapsilosis. More effective activity against biofilms was shown by the combination of nikkomycin Z with micafungin compared to the combination with caspofunginCitation116. The combination of colistin, which has no antifungal properties, with caspofungin resulted in increased potency against C. auris. This is presumed to be due to the alteration of the fungal cell wall structure by echinocandins, making it easier for colistin to interact with the fungal cell membraneCitation117. Combination therapy with voriconazole and anidulafungin has also been shown to be equally effective and less costly in patients with a haemolytic disease or haematopoietic cell transplantation compared to voriconazole monotherapyCitation118. New research suggests a synergistic effect of caspofungin and isavuconazole, but further studies are requiredCitation119. Farnesol, which is a quorum-sensing molecule that enhances the effects of antifungal drugs, exhibits synergistic effects with echinocandins against C. auris biofilmsCitation120.

During studies on echinocandin-resistant species, compounds that potentiate their effects were also discovered. The most promising was DTPA (pentetic acid), which modulates echinocandin resistance phenotypes by chelating metal cations, especially magnesium and zinc excluding iron. This compound yielded beneficial relationships in mouse models of candidiasis. A chelator that potentiates the effects of echinocandins may represent a novel therapeutic strategy for combating drug-resistant fungiCitation121.

Side effects of echinocandins

Side effects of treatment with echinocandins are much milder compared to other antifungal antibiotics. Troublesome adverse reactions that force withdrawal occur less frequently with echinocandins than with other systemic antifungal drugsCitation18. For example, amphotericin B administered intravenously can cause seizures, violent fevers, chills, myalgia, and hyperkalemia. In addition, the use of this antibiotic poses the likelihood of permanent kidney and liver damage due to its hepato- and nephrotoxicityCitation122. The most common side effects associated with intravenous infusion of echinocandins are facial flushing, edoema, rash, pruritus, thrombophlebitis, bronchospasm, dyspnoea, decreased blood pressure, and feverCitation9. The symptoms listed above can be seen with all echinocandins, but the incidence varies depending on the drug administered. Fever is a common side effect reported in approximately 35% of caspofungin-treated patients, whereas it is reported in only 1% of micafungin-treated patients. To reduce side effects, the rate of antibiotic infusion may be reducedCitation36,Citation123. Common disordered effects such as nausea, vomiting, and diarrhoea occur in 7% of patients, and 3–25% of patients treated with caspofungin are diagnosed with phlebitis. Less than 2% of patients experience these complaints after treatment with anidulafungin and micafunginCitation14. Caspofungin shows a higher frequency of liver dysfunction (1–15%) compared to other echinocandins. Micafungin may increase risk of liver cancerCitation72. Complications such as anaemia, neutropenia, leukopoenia, and thrombocytopenia comprise less than 10% of all adverse effects of echinocandinsCitation9. Echinocandins should be avoided during pregnancy because they exhibit embryotoxic and teratogenic effectsCitation124.

Antiviral activity of echinocandins

Studies on micafungin have demonstrated antiviral activity against enterovirus 71 (EV71), chikungunya virus (CHIKV) and dengue virus serotype 2 (DENV-2)Citation125–127. For the first time, micafungin as an antiviral agent was used against the proliferation and replication of enterovirus 71 (EV71) replicons with an IC50 (concentration of drug that reduce viral activity by 50%) value of 6.35 µg/mlCitation126. Micafungin also has antiviral activity against other enteroviruses such as group B coxsackievirus type 3 (CVB3) and human rhinovirus (HRV). Studies on the effect of micafungin against CHIKV have provided information on the binding of the antibiotic to the virus envelope proteins. Micafungin may affect the later stages of viral infection. The drug reduces the cytopathic effect of the virus, its replication and impairs cell-to-cell transmission. The IC50 value of micafungin against CHIKV S27 ranged from 17.2–20.63 µM, which was higher than the IC50 value against EV71. Micafungin is also effective against other alphaviruses such as SINV and SFVCitation127.

Recently, a pioneering study on the action of micafungin against dengue virus serotype 2 (DENV-2) has emerged. It is speculated that the mechanism of antiviral action of micafungin may be related to the degradation of the virion by binding the antibiotic to the envelope protein of DENV-2. The action of micafungin on the dengue virus results in the reduction of viral RNA levels. Micafungin mainly acts during the first stages of dengue virus infection. It inhibits DENV-2 binding and entry at doses ranging from 12.5–100 µM. Moreover, micafungin showed activity against other serotypes of the dengue virus (DENV-1, DENV-3 and DENV-4). Other echinocandins have also been shown to act on the dengue virus. Micafungin has the ability to suppress infections caused by arboviruses (CHIKV and DENV). The data presented by Chen et al. 2021 in the in vitro cellular model cannot be directly applied in clinical practice at this time therefore further studies require in vivo testingCitation125.

Resistance of fungi to echinocandins and “eagle-like effect”

Fungi possess adaptive mechanisms by which they compete with other microorganisms for the resources of the ecological niche they occupy. These adaptations can result in increased resistance to antifungal drugsCitation128. Multidrug-resistant Candida species have spread worldwide, which in the future may affect methods used to treat infectionsCitation129. Acquired resistance to echinocandins of various strains of C. albicans, C. dubliniensis, C. kefyr, C. glabrata, C. krusei, C. tropicalis and C. lusitaniae has been increasingly described. It is speculated that prolonged or repeated exposure to these antibiotics is a major factor in the acquisition of resistance by Candida spp. to echinocandinsCitation130.

In response to echinocandins, fungi activate adaptive mechanisms that induce cell wall repair. Signals of cell wall instability are transmitted to the Rho1 subunit of β-(1,3)-d-glucan synthase, whose function is to control glucan synthase and coordinate the activity of protein kinase C (PKC)Citation131. PKC controls the activity of other proteins responsible for maintaining the integrity of the fungal cell by synthesising a compensatory cell wall, which is mainly composed of chitin () Citation132. Increased levels of chitin synthesis in response to echinocandins can also be controlled by the mitogen-activated protein kinase (MAPK), high-osmolarity glycerol response (HOG), and calcineurin pathwaysCitation46. Calcineurin, upon calcium activation, causes dephosphorylation of the protein transcription factor Crz1 (Calcineurin-Responsive Zinc Finger), which, upon translocation to the cell nucleus, induces FKS2 expression by binding to calcium-dependent response elements (CDREs) in promoter sequencesCitation46,Citation131.

Figure 14. Mechanisms that adapt fungi to echinocandins. The protein kinase C (PKC), calcineurin, high-osmolarity glycerol (HOG) response, and mitogen-activated kinase (MAPK) pathways induce the synthesis of a compensatory cell wall composed of chitin. FKS mutations in the β-(1,3)-d-glucan synthase "hot spot" alter the enzyme's sensitivity to echinocandins and allow glucan production despite the presence of this antibiotic.

Figure 14. Mechanisms that adapt fungi to echinocandins. The protein kinase C (PKC), calcineurin, high-osmolarity glycerol (HOG) response, and mitogen-activated kinase (MAPK) pathways induce the synthesis of a compensatory cell wall composed of chitin. FKS mutations in the β-(1,3)-d-glucan synthase "hot spot" alter the enzyme's sensitivity to echinocandins and allow glucan production despite the presence of this antibiotic.

Many studies on echinocandin resistance involve mutations in the catalytic subunit of glucan synthase (Fks). Three genes encoding this subunit are known: FKS1, FKS2, and FKS3. Variations in the gene encoding the Fks3 subunit affect the enzyme activity more weakly compared to the other genes in this group and therefore are not essential for the development of drug resistanceCitation133. Most Candida species acquire resistance to echinocandins through mutations in the Fks1 subunit gene, but in C. glabrata, resistance-causing mutations can also occur in the FKS2 geneCitation134. Acquired changes in the structure of the Fks subunit are mostly amino acid substitutions, but deletions and alternative "stop" codons are also found in C. glabrataCitation130. Mutations that determine fungal resistance to echinocandins are located in two highly conserved regions of genes encoding the Fks subunit of glucan synthase called "hot spots" (HS). The point mutations are grouped into two "hot spot" regions: HS1 encoding amino acids at positions 641–649 of the enzyme and HS2 responsible for encoding amino acid residues 1345–1365 in the Fks1 protein in Candida albicans and homologous regions of Fks2 in C. glabrata. The FKS2 "hot spot" mutation was also detected in echinocandin-resistant S. cerevisiae strainsCitation133. All members of the C. parapsilosis group (C. parapsilosis, C. orthopsilosis, and C. metapsilosis) have a natural DNA polymorphism at position Pro 649 of glucan synthase, resulting in a proline substitution by alanine, making them less susceptible to caspofungin. In C. guilliermondii these mutations are located at Met 663 and Ala 634 and serve similar resultsCitation135. Glucan synthase gene sequence changes associated with the phenotype with the strongest resistance to echinocandins occur at positions encoding Ser 645 and Phe 641 and account for 80% of all mutations detected in C. albicans in the Fks1 subunitCitation136. The most common and strongest phenotype substitution involves a change in serine at position 645 to phenylalanine, proline, or tyrosineCitation129. The Ser 663 change in the Fks2 subunit in C. glabrata that is equivalent to Ser 645 in C. albicans is the most relevant amino acid substitution from the perspective of acquiring echinocandin resistance in this speciesCitation137. Mutations in A. fumigatus occur in the AfFKS1 gene of glucan synthase, which, along with increased chitin production, develops resistance of Aspergillus to echinocandinsCitation138.

Single amino acid substitutions in the glucan synthase sequence in various fungi reduce drug sensitivity by 50–3000-fold and increase MIC values by 5–100-fold. Amino acid substitutions in C. albicans Fks1 resulted in altered cell wall morphologyCitation139. C. albicans strains with a homozygous FKS1 "hot spot" mutation exhibit thicker cell walls, which may be partially attributed to a compensatory increase in wall chitin content. Mutants with high chitin content in the cell wall show reduced growth rates in a liquid medium and impaired ability to transform blastospores into hyphaeCitation140. The increase in MIC values for echinocandins depends on the position and specific amino acid substitution in the glucan synthase structure. The most significant increase in MIC values was found for changes involving the first and fifth amino acids (phenylalanine and serine, respectively) in the "hot spot 1" regions of the genes encoding FKS1 or FKS2. In most cases, FKS sequence changes cause cross-resistance to all echinocandinsCitation130. Resistance to echinocandins can change with the level of FKS gene expression in C. albicansCitation133,Citation140. FKS2 expression in C. glabrata is calcineurin-dependent, meaning that echinocandin resistance can be reduced by including calcineurin inhibitors FK506 in treatmentCitation141.

The observation of the specific strain of C. albicans which was able to grow in the presence of very high concentrations of caspofungin, significantly exceeding the MIC values, prompted attempts to explain this phenomenonCitation142. It was found that a similar phenomenon can also be observed for other species such as C. parapsilosis, C. glabrata, C. tropicalis and C. krusei, as well as for other antibiotics from the echinocandinsCitation143. This paradoxical effect of fungal growth at very high concentrations of echinocandins is referred to as “Eagle-like effect" and is defined as the reduced activity of echinocandins against fungi when exposed to a dose well above the MIC. This effect is suspected to be due to adaptation as a result of stimulation of calcineurin pathways and chitin synthesis to maintain cell wall integrityCitation144. Another hypothesis is the overproduction of polysaccharides that may complement β-(1,3)-D glucan deficiencies in the cell wall. "Eagle-like effect" is often observed in Aspergillus species (e.g. A. fumigatus) during exposure to high concentrations of caspofungin (above 1 mg/l)Citation143. The described effect was found in vitro and during clinical caspofungin treatment of patients with invasive pulmonary aspergillosisCitation67. The study showed that high doses of caspofungin on the order of 150 mg per day did not significantly improve treatment outcomes compared to patients treated with standard therapy (50 mg caspofungin per day). In addition, in cases of infections caused by C. glabrata and C. tropicalis, more failed treatments were reported at the 150 mg daily dose of caspofungin compared to the 50 mg daily dose, which may confirm the occurrence of the "Eagle-like effect" in clinical practiceCitation145.

Conclusion

Echinocandins have become a good alternative to azoles and polyenes in the treatment of severe fungal infections caused in particular by Candida and Aspergillus species due to their unique mechanism of action and relatively mild side effects compared to other antifungal drugs. These antibiotics block the synthesis of one of the main components of the fungal cell wall (β-(1,3)-d-glucan) by binding to the Fks subunit of β-(1,3)-d-glucan synthase. Their action results in cell wall defects and the death of fungal cells. The metabolic pathway that echinocandins targets does not occur in human cells, which limits side effects. They are also administered to patients with a weakened immune system, e.g. people suffering from AIDS, cancer, neutropenia, as well as to transplant patients who are particularly at risk of fungal infections due to immunosuppressive therapy. On the other hand, echinocandins are administered only intravenously, which limits the treatment process only to hospital conditions. They are also embryotoxic so they cannot be administered to pregnant women. They are also not effective in the treatment of fungal infections caused by species with limited content of β-(1,3)-d-glucan in the cell wall.

However, taking into account the limited number of available antifungal agents and the characteristic chemical structure of echinocandins, the road opens up to various structural modifications, and thus to obtain new and effective drugs. An example is rezafungin, which is currently in Phase III clinical trials. Due to its improved pharmacokinetics as well as better stability in solutions, it is a good example of the development of this class of antibiotics. Despite the relatively short history of medical use, echinocandins have been approved by the FDA and EMA as first-line drugs for oesophageal candidiasis, invasive candidemia and the prevention of mycoses in transplant patients, which undoubtedly confirms their high usefulness in the fight against mycoses. The cost of one dose of the drug varies between ± $50 (caspofungin) and ± $100 (anidulafungin, micafungin), which makes the therapy relatively expensive, but the increasing use of these drugs as well as the introduction of new preparations from this class may reduce the cost of therapy in the near future. Fungal resistance to echinocandins is found relatively rarely compared to other antifungal antibiotics, yet we already know the basic mechanisms of this phenomenon, which may help in the work on the synthesis of new representatives of more effective echinocandins.

Summarising this basic review of information on echinocandins, it can be stated that despite the relatively short period of their use, these drugs have already found a significant place in clinical practice and are now a very promising and developing class of antifungal antibiotics.

Disclosure statement

The authors report no conflict of interest.

Additional information

Funding

This study was supported by the Ministry of Education and Science, Poland as a part of subsidies for maintaining the research potential granted to the Faculty of Biology (SWB-8) and Faculty of Chemistry of the University of Białystok.

References

  • Bongomin F, Gago S, Oladele RO, et al. Global and multi-national prevalence of fungal diseases-estimate precision. J Fungi 2017;3:57.
  • MacNeil A, Glaziou P, Sismanidis C, et al. Global epidemiology of tuberculosis and progress toward achieving global targets – 2017. Morb Mortal Wkly Rep 2019;68:263–6.
  • Brown GD, Denning DW, Gow NA, et al. Hidden killers: human fungal infections. Sci Transl Med 2012;4:165rv13.
  • Neoh CF, Slavin M, Chen SC, et al. Echinocandins in the treatment of candidaemia and invasive candidiasis: clinical and economic perspectives. Int J Antimicrob Agents 2014;43:207–14.
  • Vazquez JA, Sobel JD. Anidulafungin: a novel echinocandin. Clin Infect Dis 2006;43:215–22.
  • Wall G, Lopez-Ribot JL. Current antimycotics, new prospects, and future approaches to antifungal therapy. Antibiotics 2020;9:445.
  • Hull CM, Parker JE, Bader O, et al. Facultative sterol uptake in an ergosterol-deficient clinical isolate of Candida glabrata harboring a missense mutation in ERG11 and exhibiting cross-resistance to azoles and amphotericin B. Antimicrob Agents Chemother 2012;56:4223–32.
  • Lindsay J, Teh BW, Micklethwaite K, et al. Azole antifungals and new targeted therapies for hematological malignancy. Curr Opin Infect Dis 2019;32:538–45.
  • Mroczyńska M, Brillowska-Dąbrowska A. Review on current status of echinocandins use. Antibiotics 2020;9:227.
  • Garcia-Effron G. Rezafungin-mechanisms of action, susceptibility and resistance: similarities and differences with the other echinocandins. J Fungi 2020;6:262.
  • Hüttel W. Echinocandins: structural diversity, biosynthesis, and development of antimycotics. Appl Microbiol Biotechnol 2021;105:55–66.
  • Patil A, Majumdar S. Echinocandins in antifungal pharmacotherapy. J Pharm Pharmacol 2017;69:1635–60.
  • Balkovec JM, Hughes DL, Masurekar PS, et al. Discovery and development of first in class antifungal caspofungin (CANCIDAS®)-a case study. Nat Prod Rep 2014;31:15–34.
  • Cappelletty D, Eiselstein-McKitrick K. The echinocandins. Pharmacotherapy 2007;27:369–88.
  • Krishnan BR, James KD, Polowy K, et al. CD101, a novel echinocandin with exceptional stability properties and enhanced aqueous solubility. J Antibiot 2017;70:130–5.
  • ClinicalTrials. Study of rezafungin compared to caspofungin in subjects with candidemia and/or invasive candidiasis (ReSTORE). Available from: https://clinicaltrials.gov/ct2/show/NCT03667690?cond=NCT03667690&draw=2&rank=1. [last accessed 1 Jul 2021]
  • Stan CD, Tuchiluş C, Stan CI. Echinocandins-new antifungal agents. Rev Med Chir Soc Med Nat Iasi 2014;118:528–36.
  • Chen SC, Slavin MA, Sorrell TC. Echinocandin antifungal drugs in fungal infections: a comparison. Drugs 2011;71:11–41.
  • Hashimoto S. Micafungin: a sulfated echinocandin. J Antibiot 2009;62:27–35.
  • Taormina G, Gopinath R, Moore J, et al. A regulatory review approach for evaluation of micafungin for treatment of neonatal candidiasis. Clin Infect Dis 2021;73:2335–40.
  • St. Germain RM, Ellis JM. Anidulafungin. Formulary 2006;41:387–403.
  • Zhao Y, Perlin DS. Review of the novel echinocandin antifungal rezafungin: animal studies and clinical data. J Fungi 2020;6:192.
  • Sofjan AK, Mitchell A, Shah DN, et al. Rezafungin (CD101), a next-generation echinocandin: a systematic literature review and assessment of possible place in therapy. J Glob Antimicrob Resist 2018;14:58–64.
  • Hu ZC, Li WJ, Zou SP, et al. Mutagenesis of echinocandin B overproducing Aspergillus nidulans capable of using starch as main carbon source. Prep Biochem Biotechnol 2020;50:745–52.
  • Yao J, Liu H, Zhou T, et al. Total synthesis and structure–activity relationships of caspofungin-like macrocyclic antifungal lipopeptides. Tetrahedron 2012;68:3074–85.
  • Chen L, Yue Q, Li Y, et al. Engineering of Glarea lozoyensis for exclusive production of the pneumocandin B0 precursor of the antifungal drug caspofungin acetate. Appl Environ Microbiol 2015;81:1550–8.
  • Bouffard FA, Zambias RA, Dropinski JF, et al. Synthesis and antifungal activity of novel cationic pneumocandin B(o) derivatives. J Med Chem 1994;37:222–5.
  • Cândido ES, Affonseca F, Cardoso MH, et al. Echinocandins as biotechnological tools for treating Candida auris infections. J Fungi 2020;6:185.
  • Leonard WR, Jr, Belyk KM, Conlon DA, et al. Synthesis of the antifungal beta-1,3-glucan synthase inhibitor CANCIDAS (caspofungin acetate) from pneumocandin B0. J Org Chem 2007;72:2335–43.
  • Pratt A, Garcia-Effron G, Zhao Y, et al. Evaluation of fungal-specific fluorescent labeled echinocandin probes as diagnostic adjuncts. Med Mycol 2013;51:103–7.
  • Jaber QZ, Bibi M, Ksiezopolska E, et al. Elevated vacuolar uptake of fluorescently labeled antifungal drug caspofungin predicts echinocandin resistance in pathogenic yeast. ACS Cent Sci 2020;6:1698–712.
  • Mulder MP, Kruijtzer JA, Breukink EJ, et al. Synthesis and evaluation of novel macrocyclic antifungal peptides. Bioorg Med Chem 2011;19:6505–17.
  • Mulder MP, Fodran P, Kemmink J, et al. Mutual influence of backbone proline substitution and lipophilic tail character on the biological activity of simplified analogues of caspofungin. Org BioMol Chem 2012;10:7491–502.
  • Yao J, Liu H, Zhou T, et al. Total synthesis and structure-activity relationships of new echinocandin-like antifungal cyclolipohexapeptides. Eur J Med Chem 2012;50:196–208.
  • Tomishima M, Ohki H, Yamada A, et al. Novel echinocandin antifungals. Part 2: optimization of the side chain of the natural product FR901379. Discovery of micafungin. Bioorg Med Chem Lett 2008;18:2886–90.
  • Denning DW. Echinocandin antifungal drugs. Lancet 2003;362:1142–51.
  • Marena GD, Dos Santos Ramos MA, Bauab TM, et al. Biological properties and analytical methods for micafungin: a critical review. Crit Rev Anal Chem 2021;51:312–28.
  • Norris T, VanAlsten J, Hubbs S, et al. Commercialization and late-stage development of a semisynthetic antifungal API: anidulafungin/D-fructose (Eraxis). Org Proc Res and Develop 2008;12:447–55.
  • Singh SB, Herath K, Nielsen Kahn J, et al. Synthesis and antifungal evaluation of pentyloxyl-diphenylisoxazoloyl pneumocandins and echinocandins. Bioorg Med Chem Lett 2013;23:3253–6.
  • Torre PD, Reboli AC. Anidulafungin: a new echinocandin for candidal infections. Expert Rev anti Infect Ther 2007;5:45–52.
  • Zhu B, Dong Y, Ma J, et al. The synthesis and activity evaluation of N-acylated analogs of echinocandin B with improved solubility and lower toxicity. J Pept Sci 2020;26:e3278.
  • Gow NAR, Latge JP, Munro CA. The fungal cell wall: structure, biosynthesis, and function. MiCrobiol Spectr 2017;5:1.
  • Ong V, Hough G, Schlosser M, et al. Preclinical evaluation of the stability, safety, and efficacy of CD101, a novel echinocandin. Antimicrob Agents Chemother 2016;60:6872–9.
  • Garcia-Rubio R, de Oliveira HC, Rivera J, et al. The fungal cell wall: Candida, Cryptococcus, and Aspergillus species. Front Microbiol 2019;10:2993.
  • Sucher AJ, Chahine EB, Balcer HE. Echinocandins: the newest class of antifungals. Ann Pharmacother 2009;43:1647–57.
  • Steinbach WJ, Reedy JL, Cramer RA, et al. Harnessing calcineurin as a novel anti-infective agent against invasive fungal infections. Nat Rev Microbiol 2007;5:418–30.
  • Liu W, Yuan L, Wang S. Recent progress in the discovery of antifungal agents targeting the cell wall. J Med Chem 2020;63:12429–59.
  • Aguilar-Zapata D, Petraitiene R, Petraitis V. Echinocandins: the expanding antifungal armamentarium. Clin Infect Dis 2015;61: S604–S611.
  • Gamaletsou MN, Walsh TJ, Sipsas NV. Invasive fungal infections in patients with hematological malignancies: emergence of resistant pathogens and new antifungal therapies. Turk J Haematol 2018;35:1–11.
  • Perlin DS, Cell wall-modifying antifungal drugs. In: Latgé JP, ed. The fungal cell wall. Current topics in microbiology and immunology. vol 425. Cham: Springer; 2019.
  • Eschenauer G, Depestel DD, Carver PL. Comparison of echinocandin antifungals. Ther Clin Risk Manag 2007;3:71–97.
  • Martos AI, Romero A, González MT, et al. Evaluation of the Etest method for susceptibility testing of Aspergillus spp. and Fusarium spp. to three echinocandins. Med Mycol 2010;48:858–61.
  • Kauffman CA, Pappas PG, Sobel JD, et al. Essentials of clinical mycology. New York: Springer; 2011.
  • Lockhart SR, Zimbeck AJ, Baddley JW, et al. In vitro echinocandin susceptibility of Aspergillus isolates from patients enrolled in the T\transplant-associated infection surveillance network. Antimicrob Agents Chemother 2011;55:3944–6.
  • Pfaller MA, Espinel-Ingroff A, Bustamante B, et al. Multicenter study of anidulafungin and micafungin MIC distributions and epidemiological cutoff values for eight Candida species and the CLSI M27-A3 broth microdilution method. Antimicrob Agents Chemother 2014;58:916–22.
  • Guarro J, Gams W, Pujol I, et al. Acremonium species: new emerging fungal opportunists-in vitro antifungal susceptibilities and review. Clin Infect Dis 1997;25:1222–9.
  • Pujol I, Aguilar C, Gené J, et al. In vitro antifungal susceptibility of Alternaria spp. and Ulocladium spp. J Antimicrob Chemother 2000;46:337.
  • Rodriguez-Tudela JL, Diaz-Guerra TM, Mellado E, et al. Susceptibility patterns and molecular identification of Trichosporon species. Antimicrob Agents Chemother 2005;49:4026–34.
  • Ashley ESD, Lewis R, Lewis JS, et al. Pharmacology of systemic antifungal agents. Clin Inf Dis 2006;43:S28–S39.
  • Perdomo H, Sutton DA, García D, et al. Spectrum of clinically relevant Acremonium species in the United States. J Clin Microbiol 2011;49:243–56.
  • Vitale RG, de Hoog GS, Schwarz P, et al. Antifungal susceptibility and phylogeny of opportunistic members of the order mucorales. J Clin Microbiol 2012;50:66–75.
  • da Cunha KC, Sutton DA, Fothergill AW, et al. In vitro antifungal susceptibility and molecular identity of 99 clinical isolates of the opportunistic fungal genus Curvularia. Diagn Microbiol Infect Dis 2013;76:168–74.
  • Sandoval-Denis M, Sutton DA, Cano-Lira JF, et al. Phylogeny of the clinically relevant species of the emerging fungus Trichoderma and their antifungal susceptibilities. J Clin Microbiol 2014;52:2112–25.
  • Sherry L, Ramage G, Kean R, et al. Biofilm-forming capability of highly virulent, multidrug-resistant candida auris. Emerg Infect Dis 2017;23:328–31.
  • Nishiyama Y, Uchida K, Yamaguchi H. Morphological changes of Candida albicans induced by micafungin (FK463), a water-soluble echinocandin-like lipopeptide. J Electron Microsc 2002;51:247–55.
  • Meletiadis J, Siopi M, Kanioura L, et al. Development and multicentre validation of an agar-based screening method for echinocandin susceptibility testing of Aspergillus species. J Antimicrob Chemother 2019;74:2247–54.
  • Aruanno M, Glampedakis E, Lamoth F. Echinocandins for the treatment of invasive Aspergillosis: from laboratory to bedside. Antimicrob Agents Chemother 2019;63:e00399–19.
  • Odabasi Z, Paetznick VL, Rodriguez JR, et al. In vitro activity of anidulafungin against selected clinically important mold isolates. Antimicrob Agents Chemother 2004;48:1912–5.
  • Hoffman JA, Walsh TJ. Echinocandins in children. Pediatr Infect Dis J 2011;30:508–9.
  • Diekema DJ, Messer SA, Hollis RJ, et al. Activities of caspofungin, itraconazole, posaconazole, ravuconazole, voriconazole, and amphotericin B against 448 recent clinical isolates of filamentous fungi. J Clin Microbiol 2003;41:3623–6.
  • Bal AM. The echinocandins: three useful choices or three too many? Int J Antimicrob Agents 2010;35:13–8.
  • Hashemian SM, Farhadi T, Velayati AA. Caspofungin: a review of its characteristics, activity, and use in intensive care units. Expert Rev Anti Infect Ther 2020;18:1213–20.
  • Lakota EA, Ong V, Flanagan S, et al. Population pharmacokinetic analyses for rezafungin (CD101) efficacy using phase 1 data. Antimicrob Agents Chemother 2018;62:e02603–17.
  • Ong V, Flanagan S, Sandison T, et al. 1286. Pharmacokinetics, excretion, and mass balance of [14C]-rezafungin following intravenous (IV) administration in healthy adults. Open Forum Infect Dis 2020;7:S658–S659.
  • Muilwijk EW, Lempers VJ, Burger DM, et al. Impact of special patient populations on the pharmacokinetics of echinocandins. Expert Rev Anti Infect Ther 2015;13:799–815.
  • Sandison T, Ong V, Lee J,et al. Safety and pharmacokinetics of CD101 IV, a novel echinocandin, in healthy adults. Antimicrob Agents Chemother 2017;61:e01627–16.
  • Lempers VJ, Schouten JA, Hunfeld NG, et al. Altered micafungin pharmacokinetics in intensive care unit patients. Antimicrob Agents Chemother 2015;59:4403–9.
  • Liu P, Ruhnke M, Meersseman W, et al. Pharmacokinetics of anidulafungin in critically ill patients with candidemia/invasive candidiasis. Antimicrob Agents Chemother 2013;57:1672–6.
  • Hajdu R, Thompson R, Sundelof JG, et al. Preliminary animal pharmacokinetics of the parenteral antifungal agent MK-0991 (L-743,872). Antimicrob Agents Chemother 1997;41:2339–44.
  • Flevari A, Theodorakopoulou M, Velegraki A, et al. Treatment of invasive candidiasis in the elderly: a review. Clin Interv Aging 2013;8:1199–208.
  • Dekkers BGJ, Veringa A, Marriott DJE, et al. Invasive candidiasis in the elderly: considerations for drug therapy. Drugs Aging 2018;35:781–9.
  • Isnard C, Hernandez SB, Guérin F, et al. Unexpected cell wall alteration-mediated bactericidal activity of the antifungal caspofungin against vancomycin-resistant Enterococcus faecium. Antimicrob Agents Chemother 2020;64:e01261–20.
  • Keating G, Figgitt D. Caspofungin: a review of its use in oesophageal candidiasis, invasive candidiasis and invasive aspergillosis. Drugs 2003;63:2235–63.
  • Houšť J, Spížek J, Havlíček V. Antifungal Drugs. Metabolites 2020;10:106.
  • McCormack PL, Perry CM. Caspofungin: a review of its use in the treatment of fungal infections. Drugs 2005;65:2049–68.
  • Yang Q, Zhang T, Zhao D, et al. Factors influencing caspofungin plasma concentrations in kidney transplant patients with high incidence of invasive fungal infections. J Clin Pharm Ther 2020;45:72–80.
  • Nguyen TH, Hoppe-Tichy T, Geiss HK, et al. Factors influencing caspofungin plasma concentrations in patients of a surgical intensive care unit. J Antimicrob Chemother 2007;60:100–6.
  • Prabhakar H. Pharmacology in clinical neurosciences. Singapore: Springer; 2020.
  • Arendrup MC, Meletiadis J, Zaragoza O, et al. Multicentre determination of rezafungin (CD101) susceptibility of Candida species by the EUCAST method. Clin Microbiol Infect 2018;24:1200–4.
  • Miesel L, Cushion MT, Ashbaugh A, et al. Efficacy of rezafungin in prophylactic mouse models of invasive candidiasis, aspergillosis, and Pneumocystis pneumonia. Antimicrob Agents Chemother 2021;65:e01992–20.
  • Kurland S, Furebring M, Löwdin E, et al. Pharmacokinetics of caspofungin in critically ill patients in relation to liver dysfunction: differential impact of plasma albumin and bilirubin levels. Antimicrob Agents Chemother 2019;63:e02466–18.
  • Yu IG, O'Brien SE, Ryckman DM. Pharmacokinetic and pharmacodynamic comparison of intravenous and inhaled caspofungin. J Aerosol Med Pulm Drug Deliv 2021;34:197–203.
  • Leshinsky J, McLachlan A, Foster DJR, et al. Pharmacokinetics of caspofungin acetate to guide optimal dosing in cats. PLoS One 2017;12:e0178783.
  • Carter NJ, Keating GM. Micafungin: a review of its use in the prophylaxis and treatment of invasive Candida infections in pediatric patients. Paediatr Drugs 2009;11:271–91.
  • Wasmann RE, Muilwijk EW, Burger DM, et al. Clinical pharmacokinetics and pharmacodynamics of micafungin. Clin Pharmacokinet 2018;57:267–86.
  • Scott LJ. Micafungin: a review of its use in the prophylaxis and treatment of invasive Candida infections. Drugs 2012;72:2141–65.
  • Epstein DJ, Seo SK, Huang YT, et al. Micafungin versus posaconazole prophylaxis in acute leukemia or myelodysplastic syndrome: a randomized study. J Infect 2018;77:227–34.
  • Guirao-Abad JP, Sánchez-Fresneda R, Machado F, et al. Micafungin enhances the human macrophage response to Candida albicans through β-glucan exposure. Antimicrob Agents Chemother 2018;62:e02161–17.
  • Murdoch D, Plosker GL. Anidulafungin. Anidulafungin Drugs 2004;64:2249–58.
  • Staab JF, Kahn JN, Marr KA. Differential Aspergillus lentulus echinocandin susceptibilities are Fksp independent. Antimicrob Agents Chemother 2010;54:4992–8.
  • Kofla G, Ruhnke M. Pharmacology and metabolism of anidulafungin, caspofungin and micafungin in the treatment of invasive candidosis: review of the literature. Eur J Med Res 2011;16:159–66.
  • Xie R, McFadyen L, Raber S, et al. Population analysis of anidulafungin in infants to older adults with confirmed or suspected invasive candidiasis. Clin Pharmacol Ther 2020;108:316–25.
  • Kullberg BJ, Vasquez J, Mootsikapun P, et al. Efficacy of anidulafungin in 539 patients with invasive candidiasis: a patient-level pooled analysis of six clinical trials. J Antimicrob Chemother 2017;72:2368–77.
  • Sganga G, Wang M, Capparella MR, et al. Evaluation of anidulafungin in the treatment of intra-abdominal candidiasis: a pooled analysis of patient-level data from 5 prospective studies. Eur J Clin Microbiol Infect Dis 2019;38:1849–56.
  • Cornely FB, Cornely OA, Salmanton-García J, et al. Attributable mortality of candidemia after introduction of echinocandins. Mycoses 2020;63:1373–81.
  • Chen PY, Yu CJ, Chien JY, et al. Anidulafungin as an alternative treatment for Pneumocystis jirovecii pneumonia in patients who cannot tolerate trimethoprim/sulfamethoxazole. Int J Antimicrob Agents 2020;55:105820.
  • Vera-González N, Bailey-Hytholt CM, Langlois L, et al. Anidulafungin liposome nanoparticles exhibit antifungal activity against planktonic and biofilm Candida albicans. J Biomed Mater Res A 2020;108:2263–76.
  • Wiederhold NP, Locke JB, Daruwala P, et al. Rezafungin (CD101) demonstrates potent in vitro activity against Aspergillus, including azole-resistant Aspergillus fumigatus isolates and cryptic species. J Antimicrob Chemother 2018;73:3063–7.
  • Gintjee TJ, Donnelley MA, Thompson GR. 3rd., Aspiring antifungals: review of current antifungal pipeline developments. J Fungi 2020;6:28.
  • Zhao Y, Perez WB, Jiménez-Ortigosa C, et al. CD101: a novel long-acting echinocandin. Cell Microbiol 2016;18:1308–16.
  • Stover KR, Cleary JD. Antifungal penetration and distribution into organs and tissue. Curr Fun Infe Rep 2020;14:279–88.
  • Chandra J, Ghannoum MA. CD101, a novel echinocandin, possesses potent antibiofilm activity against early and mature Candida albicans biofilms. Antimicrob Agents Chemother 2018;62:e01750–17.
  • Nyirjesy P, Alessio C, Jandourek A, et al. CD101 topical compared with oral fluconazole for acute vulvovaginal candidiasis: a randomized controlled trial. J Low Genit Tract Dis 2019;23:226–9.
  • Khalifa HO, Majima H, Watanabe A, et al. In vitro characterization of twenty-one antifungal combinations against echinocandin-resistant and -susceptible Candida glabrata. J Fungi 2021;7:108.
  • Buil JB, Brüggemann RJM, Bedin Denardi L, et al. In vitro interaction of isavuconazole and anidulafungin against azole-susceptible and azole-resistant Aspergillus fumigatus isolates. J Antimicrob Chemother 2020;75:2582–6.
  • Kovács R, Nagy F, Tóth Z, et al. Synergistic effect of nikkomycin Z with caspofungin and micafungin against Candida albicans and Candida parapsilosis biofilms. Lett Appl Microbiol 2019;69:271–8.
  • Bidaud AL, Djenontin E, Botterel F, et al. Colistin interacts synergistically with echinocandins against Candida auris. Int J Antimicrob Agents 2020;55:105901.
  • Grau S, Azanza JR, Ruiz I, et al. Cost-effectiveness analysis of combination antifungal therapy with voriconazole and anidulafungin versus voriconazole monotherapy for primary treatment of invasive aspergillosis in Spain. Clinicoecon Outcomes Res 2017;9:39–47.
  • Nagy F, Tóth Z, Nyikos F, et al. In vitro and in vivo interaction of caspofungin with isavuconazole against Candida auris planktonic cells and biofilms. Med Mycol 2021;59:1015–23.
  • Nagy F, Tóth Z, Daróczi L, et al. Farnesol increases the activity of echinocandins against Candida auris biofilms. Med Mycol 2020;58:404–7.
  • Polvi EJ, Averette AF, Lee SC, et al. Metal chelation as a powerful strategy to probe cellular circuitry governing fungal drug resistance and morphogenesis. PLoS Genet 2016;12:e1006350.
  • Laniado-Laborín R, Cabrales-Vargas MN. Amphotericin B: side effects and toxicity. Rev Iberoam Micol 2009;26:223–7.
  • Swaminathan S, Kamat S, Pinto NA. Echinocandins: their role in the management of Candida biofilms. Indian J Med Microbiol 2018;36:87–92.
  • Pilmis B, Jullien V, Sobel J, et al. Antifungal drugs during pregnancy: an updated review. J Antimicrob Chemother 2015;70:14–22.
  • Chen YC, Lu JW, Yeh CT, et al. Micafungin inhibits dengue virus infection through the disruption of virus binding, entry, and stability. Pharmaceuticals 2021;14:338.
  • Kim C, Kang H, Kim DE, et al. Antiviral activity of micafungin against enterovirus 71. Virol J 2016;13:99.
  • Ho YJ, Liu FC, Yeh CT, et al. Micafungin is a novel anti-viral agent of chikungunya virus through multiple mechanisms. Antiviral Res 2018;159:134–42.
  • Healey KR, Perlin DS. Fungal resistance to echinocandins and the MDR phenomenon in Candida glabrata. J Fungi 2018;4:105.
  • Pristov KE, Ghannoum MA. Resistance of Candida to azoles and echinocandins worldwide. Clin Microbiol Infect 2019;25:792–8.
  • Arendrup MC, Perlin DS. Echinocandin resistance: an emerging clinical problem?. Curr Opin Infect Dis 2014;27:484–92.
  • Levin DE. Cell wall integrity signaling in Saccharomyces cerevisiae. Microbiol Mol Biol Rev 2005;69:262–91.
  • Lesage G, Sdicu AM, Ménard P, et al. Analysis of beta-1,3-glucan assembly in Saccharomyces cerevisiae using a synthetic interaction network and altered sensitivity to caspofungin. Genetics 2004;167:35–49.
  • Suwunnakorn S, Wakabayashi H, Kordalewska M, et al. FKS2 and FKS3 genes of opportunistic human pathogen Candida albicans influence echinocandin susceptibility. Antimicrob Agents Chemother 2018;62:e02299–17.
  • Hou X, Healey KR, Shor E, et al. Novel FKS1 and FKS2 modifications in a high-level echinocandin resistant clinical isolate of Candida glabrata. Emerg Microbes Infect 2019;8:1619–25.
  • Perlin DS. Echinocandin resistance in Candida. Clin Infect Dis 2015;61:S612–S617.
  • Perlin DS. Current perspectives on echinocandin class drugs. Future Microbiol 2011;6:441–57.
  • Garcia-Effron G, Lee S, Park S, et al. Effect of Candida glabrata FKS1 and FKS2 mutations on echinocandin sensitivity and kinetics of 1,3-beta-D-glucan synthase: implication for the existing susceptibility breakpoint. Antimicrob Agents Chemother 2009;53:3690–9.
  • Sharma C, Chowdhary A. Molecular bases of antifungal resistance in filamentous fungi. Int J Antimicrob Agents 2017;50:607–16.
  • Cowen LE, Sanglard D, Howard SJ, et al. Mechanisms of antifungal drug resistance. Cold Spring Harb Perspect Med 2014;5:a019752.
  • Ben-Ami R, Garcia-Effron G, Lewis RE, et al. Fitness and virulence costs of Candida albicans FKS1 hot spot mutations associated with echinocandin resistance. J Infect Dis 2011;204:626–35.
  • Yu SJ, Chang YL, Chen YL. Calcineurin signaling: lessons from Candida species. FEMS Yeast Res 2015;15:fov016.
  • Stevens DA, White TC, Perlin DS, et al. Studies of the paradoxical effect of caspofungin at high drug concentrations. Diagn Microbiol Infect Dis 2005;51:173–8.
  • Chamilos G, Lewis RE, Albert N, et al. Paradoxical effect of echinocandins across Candida species in vitro: evidence for echinocandin-specific and Candida species-related differences. Antimicrob Agents Chemother 2007;51:2257–9.
  • Vanstraelen K, Lagrou K, Maertens J, et al. The eagle-like effect of echinocandins: what's in a name?. Expert Rev Anti Infect Ther 2013;11:1179–91.
  • Emri T, Majoros L, Tóth V, et al. Echinocandins: production and applications. Appl Microbiol Biotechnol 2013;97:3267–84.