Next Article in Journal
Socioeconomic Status and the Gut Microbiome: A TwinsUK Cohort Study
Previous Article in Journal
Complement Activation Contributes to the Pathophysiology of Shiga Toxin-Associated Hemolytic Uremic Syndrome
Previous Article in Special Issue
Structure and Synthesis of Antifungal Disulfide β-Strand Proteins from Filamentous Fungi
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Editorial

Do Antimicrobial Proteins Contribute to Overcoming the Hidden Antifungal Crisis at the Dawn of a Post-Antibiotic Era?

by
László Galgóczy
1,2 and
Florentine Marx
3,*
1
Institute of Plant Biology, Biological Research Centre, Hungarian Academy of Sciences, Temesvári krt. 62, H-6726 Szeged, Hungary
2
Department of Microbiology, Faculty of Science and Informatics, University of Szeged, Közép fasor 52, H-6726 Szeged, Hungary
3
Biocenter, Division of Molecular Biology, Medical University of Innsbruck, Innrain 80-82, A-6020 Innsbruck, Austria
*
Author to whom correspondence should be addressed.
Submission received: 9 January 2019 / Accepted: 10 January 2019 / Published: 11 January 2019
(This article belongs to the Special Issue Antimicrobial Proteins in Filamentous Fungi)
The incidence of fungal infections has been grossly underestimated in the past decades as a consequence of poor identification techniques and a lack of regular epidemiologic surveys in low- and middle-income countries. The most affected areas of the world are the countries where some routinely administrated and effective antifungal drugs are not available, especially for people under living standards [1]. Approximately one quarter of the worldwide population suffers from superficial mycoses [2], which is the fourth most common illness on Earth after headaches and dental caries [3]. In the immunocompetent patient, cutaneous fungal infections are non-life-threatening, but associated with social embarrassment and stigma [1], because they might be recurrent and difficult to eradicate [4]. Patients who suffer from impaired and weak immune status due to infections (e.g. HIV), hematological malignancies, who receive immunosuppressive therapy or necessitate intensive care (e.g. organ transplantation) are at high risk for microbial infections in general and more specifically to develop severe invasive mycoses, such as aspergillosis, candidiasis, cryptococcosis, histoplasmosis, talaromycosis and pneumocystosis with fatal outcome [5,6]. It has been estimated that the number of people that die from fungal diseases per year is > 1.5 million and outreaches the numbers that die from tuberculosis or malaria [2].
Plant pathogenic fungi also pose severe risk to human and animal health, and food supply. Crops, worth billions of € per year, are destroyed by plant pathogenic fungi that constitute the main group of phytopathogens. Mycotoxin-contaminated crops, vegetables, fruits, and animal feeds are continuously increasing in Europe in the last years [7]. Moreover, fungal bio-deterioration of buildings, paintings, object arts and books also have impact on our cultural heritage [8]; furthermore, indoor molds affect human health with impairing the immune status of healthy individuals [9]. Finally, the emergence of novel fungal diseases in plants and animals, as a result of climate change, is affecting global sustainability and biodiversity [10,11].
The number of effective antifungal drugs to prevent fungal contaminations and treat life-threatening mycoses is modest in comparison to the armamentarium available for antibacterial treatment strategies [12]. The high similarity between fungal pathogen and host in terms of cellular morphology, physiology and metabolism hampers the identification and development of new compounds that interfere with targets that are unique to fungi and guarantee improved tolerance or at least minimize severe side effects in the host. Consequently, pesticides and feed supplements in agriculture and animal breeding to prevent or reduce fungal infections belong to the same class of active agents that are also applied in human medical treatment. Since the early 1990’s, when triazoles were introduced in antifungal therapy, a pronounced increase in Candida and Aspergillus species was observed that were less susceptible or even resistant to these licensed drugs [13,14]. In contrast to fungi that are intrinsically resistant against certain drug classes (e.g. non-albicans Candida species, NAC) or are multidrug-resistant (e.g. Candida glabrata [15], Candida auris [16], Fusarium solani [17], and Lomentospora (formerly Scedosporium) prolificans [18]), a resistance mechanism can also be acquired following repeated and long-term exposure to antifungal drugs. This accounts for iatrogenic azole resistance development in fungi but also the emergence of azole-resistant aspergillosis as a consequence of wide agricultural use of this class of antifungal drugs [19]. In the latter case azole therapy may fail in the host who has never been treated with azoles before, but suffers from infection with an azole-resistant Aspergillus fumigatus environmental strain [20].
Thus, early and accurate diagnosis is essential to promote the proper administration of the most effective antifungal therapy and reduce morbidity and mortality. In parallel, effective surveys need to be implemented that ensure the use of fungicides in food production and animal and human medicine. Furthermore, collaborative efforts are required to identify and develop new compounds with high fungal specificity and novel antifungal mechanisms to overcome the limitations of resistance development and adverse toxicity in the host.
In this respect, nature constitute a rich source for proteins and peptides that promise future drug development. Small, cysteine-rich, cationic proteins that are produced by the most diverse organisms throughout the phylogenetic tree are potential candidates for new antifungal medicines and therapeutic strategies [21]. Most promising thereby are antimicrobial proteins and peptides (AMPs) originating from filamentous fungi belonging to the class Eurotiomycetes (division Ascomycota). Numerous representatives have been extensively analyzed to date, reviewed in [22,23,24,25] and even more await their identification and characterization as the genomes of filamentous ascomycetes harbor information for still unidentified proteins with antimicrobial potential [26,27,28,29]. Deep mining of fungal genome databases and accurate phylogenetic analyses promote these efforts [29]. One obstacle in protein isolation might be the unknown gene regulation and/or very low protein expression levels. This, however, can be overcome by using (heterologous) expression systems that allow the generation and purification of recombinant proteins in sufficient quantity and of high quality that ensure thorough characterization of their structure and function [27,30]. Also, chemical synthesis has been developed for cysteine-rich AMPs, coping with the challenge of correct disulfide bonding within the protein during the synthesis process [31,32].
AMPs from Eurotiomycetes share common features: (i) they are encoded as pre-pro proteins with a signal- and a pro-sequence, which both are cleaved during protein processing and secretion into the culture broth; (ii) the mature proteins consist of ca. 50–60 amino acids (Mw: ~6000 Da), lack any posttranslational modifications, and three to four disulfide bridges stabilize a compact β-fold structure; (iii) all proteins investigated in detail so far exhibit high stability against harsh environmental conditions and (iv) exclusively inhibit the growth of human- and plant-pathogenic fungi in a fungistatic and/or fungicidal manner at µM concentrations [21]. In addition, antiviral activity has been reported in one case [33]. However, some differences exist in species specificity, antifungal efficacy, and mode of action, which are distinct from that of conventional antifungal drugs [21].
Their antifungal mechanisms are intensively studied nowadays. On the one hand, the use of genetically manipulated fungal strains allowed to define pathways involved in protein toxicity [34,35,36,37,38]. On the other hand, comprehensive nuclear magnetic resonance analyses provided detailed knowledge on the solution structure of AMPs and proposed protein motifs that might play an important role in protein function and host interaction [32,33,39,40]. Unfortunately, no fungal-specific target molecule for AMPs has been identified so far, which restrains the development of new, safely applicable AMP-based antifungal strategies. However, the availability of accurate AMP solution structures deposited in the protein data bases (such as the worldwide Protein Data Bank [41]) provide a valuable source for in silico homology modelling and molecular dynamics simulations to propose not only the structure of uncharacterized AMP candidates but also the mechanistic way of action [42]. These structural and functional predictions pave the way towards the modification of natural proteins and peptides applying rational design tools or combinatorial approaches to generate novel AMPs with improved efficacy against plant and human pathogenic fungi [29,32,43,44]. As the chemical synthesis of peptides and proteins becomes more cost-effective [45], AMPs and short peptide derivatives have significant commercial potential on the global market to be produced and applied as new antifungal compounds in the future [46].
In vitro and in vivo experiments approved fungal AMP tolerance of mammalian and plant cells and efficacy in reducing fungal pathogen load in plant, fruit, and animal models, corroborating their suitability for the development of new antifungal strategies in medical treatment, plant protection and food preservation [25,29,33,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65].
Apart from the AMPs originating from fungi of the class Eurotiomycetes, small antimicrobial peptides from various mitosporic filamentous fungi belonging to the genus Trichoderma (class Sordariomycetes) have been characterized that form a unique class of so-called peptaibols. These small peptides (ca. 5–20 amino acid long; Mw: ~500–2200 Da) contain unusual amino acids and are synthesized by non-ribosomal peptide synthases. Their antibacterial and antifungal activity is closely linked with a helical structure, that favors aggregation and ion channel formation in lipid bilayers causing membrane damage [66,67,68,69].
In this Special Issue on "Antimicrobial Proteins in Filamentous Fungi", the contributors addressed achievements in this research field and posed questions that need to be answered to better understand the nature of AMPs and to push forward the development of novel AMP-based antimicrobial strategies and to overcome antimicrobial resistance.
Heredero et al. aimed in their work at the rational design of chimeric protein:peptide molecules with improved antifungal efficacy. The combination of the Penicillium digitatum AMP AfpB with the hexapeptide PAF26 resulted in biotechnologically produced protein chimeras that allowed new insights into AMP design, structure, and function [70].
The paper of Marik et al. focuses on the characterization of new peptaibols from Trichoderma gamsii and Trichoderma koningiopsis, belonging to clade Viride, showing variations in antimicrobial activity that can be assigned to differences in the cell wall structure of the target organisms [71].
The study by Delgado et al. proved the potential of a Penicillium chrysogenum AMP, the PgAFP in food protection. PgAFP combined with the food ripening-yeast Debaryomyces hansenii effectively reduced Aspergillus parasiticus growth and aflatoxin contamination in dry-fermented sausage and cheese [72].
Unravelling the AMP structure and function provides important information for the improvement of AMP efficacy and the development of new treatment strategies by rational drug design. An overview on structure peculiarities of β-strand disulfide AMPs from ascomycetous origin and possibilities of their production by modern synthetic chemistry methods and recombinant technology is given in the mini-review by Váradi et al. [73].
Finally, the review of Meyer and Jung [74] compares structure and function of AMPs from Ascomycetes with those of bacterial cannibal toxins and redirects our narrow human conception of AMPs as purely bioactive molecules towards their important role in controlling different cellular processes to ensure optimal fitness of the producing fungal organism.
In conclusion, we want to thank all contributors to this Special Issue on AMPs in filamentous fungi for their participation. The papers published in this Special Issue provide a most valuable overview on the functional and structural complexity of the members of this AMP group and reflects the big efforts made by this community to unravel the most divers aspects in understanding fungal AMPs. This should encourage the worldwide scientific community engaged in AMP science to specifically expand their attention towards AMPs from filamentous fungi, many of which still await identification and characterization.

Funding

L.G. is financed from the Postdoctoral Excellence Programme (PD 120808) and the bilateral Austrian-Hungarian Joint Research Project (ANN 122833) of the Hungarian National Research, Development, and Innovation Office (NKFI Office). F.M. is supported from the Austrian Science Fund FWF (P25894-B20, I1644-B20 and I3132-B21). Research of L.G. has been supported by the János Bolyai Research Scholarship of the Hungarian Academy of Sciences. Present work of L.G. was supported by the UNKP-18-4 New National Excellence Program of the Ministry of Human Capacities.

Acknowledgments

We are most thankful to the excellent work of all peer reviewers, and we also want to acknowledge the support of the staff members of the MDPI Microorganisms Editorial Office for their support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tudela, J.L.R.; Denning, D.W. Recovery from serious fungal infections should be realisable for everyone. Lancet Infect. Dis. 2017, 17, 1111–1113. [Google Scholar] [CrossRef]
  2. Brown, G.D.; Denning, D.W.; Gow, N.A.; Levitz, S.M.; Netea, M.G.; White, T.C. Hidden killers: Human fungal infections. Sci. Transl. Med. 2012, 4, 165rv13. [Google Scholar] [CrossRef] [PubMed]
  3. Vos, T.; Flaxman, A.D.; Naghavi, M.; Lozano, R.; Michaud, C.; Ezzati, M.; Shibuya, K.; Salomon, J.A.; Abdalla, S.; Aboyans, V.; et al. Years lived with disability (YLDs) for 1160 sequelae of 289 diseases and injuries 1990–2010: A systematic analysis for the Global Burden of Disease Study 2010. Lancet 2012, 380, 2163–2196. [Google Scholar] [CrossRef]
  4. Dias, M.F.; Quaresma-Santos, M.V.; Bernardes-Filho, F.; Amorim, A.G.; Schechtman, R.C.; Azulay, D.R. Update on therapy for superficial mycoses: review article part I. An. Bras. Dermatol. 2013, 88, 764–774. [Google Scholar] [CrossRef] [PubMed]
  5. Bassetti, M.; Bouza, E. Invasive mould infections in the ICU setting: complexities and solutions. J. Antimicrob. Chemother. 2017, 72, i39–i47. [Google Scholar] [CrossRef] [PubMed]
  6. Bassetti, M.; Garnacho-Montero, J.; Calandra, T.; Kullberg, B.; Dimopoulos, G.; Azoulay, E.; Chakrabarti, A.; Kett, D.; Leon, C.; Ostrosky-Zeichner, L.; et al. Intensive care medicine research agenda on invasive fungal infection in critically ill patients. Intensive Care Med. 2017, 43, 1225–1238. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Magan, N.; Medina, A.; Aldred, D. Possible climate-change effects on mycotoxin contamination of food crops pre- and postharvest. Plant Pathol. 2011, 60, 150–163. [Google Scholar] [CrossRef] [Green Version]
  8. Sterflinger, K.; Pinzari, F. The revenge of time: fungal deterioration of cultural heritage with particular reference to books, paper and parchment. Environ. Microbiol. 2012, 14, 559–566. [Google Scholar] [CrossRef]
  9. Denning, D.W.; Chakrabarti, A. Pulmonary and sinus fungal diseases in non-immunocompromised patients. Lancet Infect. Dis. 2017, 17, e357–e366. [Google Scholar] [CrossRef]
  10. Fisher, M.C.; Henk, D.A.; Briggs, C.J.; Brownstein, J.S.; Madoff, L.C.; McCraw, S.L.; Gurr, S.J. Emerging fungal threats to animal, plant and ecosystem health. Nature 2012, 484, 186–194. [Google Scholar] [CrossRef] [Green Version]
  11. Debourgogne, A.; Dorin, J.; Machouart, M. Emerging infections due to filamentous fungi in humans and animals: only the tip of the iceberg? Environ. Microbiol. Rep. 2016, 8, 332–342. [Google Scholar] [CrossRef] [PubMed]
  12. Robbins, N.; Wright, G.; Cowen, L. Antifungal drugs: The current armamentarium and development of new agents. Microbiol. Spectrum 2016, 4, FUNK-0002-2016. [Google Scholar]
  13. Garcia-Rubio, R.; Cuenca-Estrella, M.; Mellado, E. Triazole resistance in Aspergillus species: An emerging problem. Drugs 2017, 77, 599–613. [Google Scholar] [CrossRef] [PubMed]
  14. Kontoyiannis, D.P. Antifungal resistance: An emerging reality and a global challenge. J. Infect. Dis. 2017, 216, S431–S435. [Google Scholar] [CrossRef] [PubMed]
  15. Healey, K.R.; Zhao, Y.; Perez, W.B.; Lockhart, S.R.; Sobel, J.D.; Farmakiotis, D.; Kontoyiannis, D.P.; Sanglard, D.; Taj-Aldeen, S.J.; Alexander, B.D.; et al. Prevalent mutator genotype identified in fungal pathogen Candida glabrata promotes multi-drug resistance. Nat. Commun. 2016, 7, 11128. [Google Scholar] [CrossRef] [PubMed]
  16. Lockhart, S.R.; Etienne, K.A.; Vallabhaneni, S.; Farooqi, J.; Chowdhary, A.; Govender, N.P.; Colombo, A.L.; Calvo, B.; Cuomo, C.A.; Desjardins, C.A.; et al. Simultaneous emergence of multidrug-resistant Candida auris on 3 continents confirmed by whole-genome sequencing and epidemiological analyses. Clin. Infect. Dis. 2017, 64, 134–140. [Google Scholar] [CrossRef] [PubMed]
  17. Walsh, T.J.; Groll, A.; Hiemenz, J.; Fleming, R.; Roilides, E.; Anaissie, E. Infections due to emerging and uncommon medically important fungal pathogens. Clin. Microbiol. Infect. 2004, 10, 48–66. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Lackner, M.; Hagen, F.; Meis, J.F.; Gerrits van den Ende, A.H.; Vu, D.; Robert, V.; Fritz, J.; Moussa, T.A.; de Hoog, G.S. Susceptibility and diversity in the therapy-refractory genus scedosporium. Antimicrob. Agents Chemother. 2014, 58, 5877–5885. [Google Scholar] [CrossRef] [PubMed]
  19. Verweij, P.E.; Chowdhary, A.; Melchers, W.J.; Meis, J.F. Azole resistance in Aspergillus fumigatus: Can we retain the clinical use of mold-active antifungal azoles? Clin. Infect. Dis. 2016, 62, 362–368. [Google Scholar] [CrossRef] [PubMed]
  20. Moye-Rowley, W.S. Multiple mechanisms contribute to the development of clinically significant azole resistance in Aspergillus fumigatus. Front. Microbiol. 2015, 6, 70. [Google Scholar] [CrossRef]
  21. Hegedüs, N.; Marx, F. Antifungal proteins: More than antimicrobials? Fungal. Biol. Rev. 2013, 26, 132–145. [Google Scholar] [CrossRef] [PubMed]
  22. Marx, F. Small, basic antifungal proteins secreted from filamentous ascomycetes: a comparative study regarding expression, structure, function and potential application. Appl. Microbiol. Biotechnol. 2004, 65, 133–142. [Google Scholar] [CrossRef] [PubMed]
  23. Marx, F.; Binder, U.; Leiter, E.; Pócsi, I. The Penicillium chrysogenum antifungal protein PAF, a promising tool for the development of new antifungal therapies and fungal cell biology studies. Cell. Mol. Life Sci. 2008, 65, 445–454. [Google Scholar] [CrossRef]
  24. Meyer, V. A small protein that fights fungi: AFP as a new promising antifungal agent of biotechnological value. Appl. Microbiol. Biotechnol. 2008, 78, 17–28. [Google Scholar] [CrossRef] [PubMed]
  25. Delgado, J.; Owens, R.A.; Doyle, S.; Asensio, M.A.; Núñez, F. Manuscript title: antifungal proteins from moulds: analytical tools and potential application to dry-ripened foods. Appl. Microbiol. Biotechnol. 2016, 100, 6991–7000. [Google Scholar] [CrossRef] [PubMed]
  26. Garrigues, S.; Gandía, M.; Marcos, J.F. Occurrence and function of fungal antifungal proteins: A case study of the citrus postharvest pathogen Penicillium digitatum. Appl. Microbiol. Biotechnol. 2016, 100, 2243–2256. [Google Scholar] [CrossRef] [PubMed]
  27. Tóth, L.; Kele, Z.; Borics, A.; Nagy, L.G.; Váradi, G.; Virágh, M.; Takó, M.; Vágvölgyi, C.; Galgóczy, L. NFAP2, a novel cysteine-rich anti-yeast protein from Neosartorya fischeri NRRL 181: Isolation and characterization. AMB Express 2016, 6, 75. [Google Scholar] [CrossRef]
  28. Paege, N.; Jung, S.; Schäpe, P.; Müller-Hagen, D.; Ouedraogo, J.P.; Heiderich, C.; Jedamzick, J.; Nitsche, B.M.; van den Hondel, C.A.; Ram, A.F.; et al. A transcriptome meta-analysis proposes novel biological roles for the antifungal protein AnAFP in Aspergillus niger. PLoS ONE 2016, 11, e0165755. [Google Scholar] [CrossRef]
  29. Sonderegger, C.; Váradi, G.; Galgóczy, L.; Kocsubé, S.; Posch, W.; Borics, A.; Dubrac, S.; Tóth, G.K.; Wilflingseder, D.; Marx, F. The evolutionary conserved γ-core motif influences the anti-Candida activity of the Penicillium chrysogenum antifungal protein PAF. Front. Microbiol. 2018, 9, 1655. [Google Scholar] [CrossRef]
  30. Sonderegger, C.; Galgóczy, L.; Garrigues, S.; Fizil, Á.; Borics, A.; Manzanares, P.; Hegedüs, N.; Huber, A.; Marcos, J.F.; Batta, G.; et al. A Penicillium chrysogenum-based expression system for the production of small, cysteine-rich antifungal proteins for structural and functional analyses. Microb. Cell. Fact. 2016, 15, 192. [Google Scholar] [CrossRef]
  31. Váradi, G.; Tóth, G.K.; Kele, Z.; Galgóczy, L.; Fizil, Á.; Batta, G. Synthesis of PAF, an antifungal protein from P. chrysogenum, by native chemical ligation: native disulfide pattern and fold obtained upon oxidative refolding. Chemistry 2013, 19, 12684–12692. [Google Scholar]
  32. Tóth, L.; Váradi, G.; Borics, A.; Batta, G.; Kele, Z.; Vendrinszky, Á.; Tóth, R.; Ficze, H.; Tóth, G.K.; Vágvölgyi, C.; et al. Anti-candidal activity and functional mapping of recombinant and synthetic Neosartorya fischeri antifungal protein 2 (NFAP2). Front. Microbiol. 2018, 9, 393. [Google Scholar] [CrossRef] [PubMed]
  33. Huber, A.; Hajdu, D.; Bratschun-Khan, D.; Gáspári, Z.; Varbanov, M.; Philippot, S.; Fizil, Á.; Czajlik, A.; Kele, Z.; Sonderegger, C.; et al. New antimicrobial potential and structural properties of PAFB: A cationic, cysteine-rich protein from Penicillium chrysogenum Q176. Sci. Rep. 2018, 8, 1751. [Google Scholar] [CrossRef] [PubMed]
  34. Martín-Urdiroz, M.; Martínez-Rocha, A.L.; Di Pietro, A.; Martínez-del-Pozo, A.; Roncero, M.I. Differential toxicity of antifungal protein AFP against mutants of Fusarium oxysporum. Int. Microbiol. 2009, 12, 115–121. [Google Scholar] [PubMed]
  35. Binder, U.; Oberparleiter, C.; Meyer, V.; Marx, F. The antifungal protein PAF interferes with PKC/MPK and cAMP/PKA signalling of Aspergillus nidulans. Mol. Microbiol. 2010, 75, 294–307. [Google Scholar] [CrossRef] [PubMed]
  36. Binder, U.; Chu, M.; Read, N.D.; Marx, F. The antifungal activity of the Penicillium chrysogenum protein PAF disrupts calcium homeostasis in Neurospora crassa. Eukaryot. Cell. 2010, 9, 1374–1382. [Google Scholar] [CrossRef] [PubMed]
  37. Binder, U.; Bencina, M.; Eigentler, A.; Meyer, V.; Marx, F. The Aspergillus giganteus antifungal protein AFPNN5353 activates the cell wall integrity pathway and perturbs calcium homeostasis. BMC Microbiol. 2011, 11, 209. [Google Scholar] [CrossRef]
  38. Gandía, M.; Garrigues, S.; Hernanz-Koers, M.; Manzanares, P.; Marcos, J.F. Differential roles, crosstalk and response to the antifungal protein AfpB in the three mitogen-activated protein kinases (MAPK) pathways of the citrus postharvest pathogen Penicillium digitatum. Fungal. Genet. Biol. 2018. [Google Scholar] [CrossRef]
  39. Campos-Olivas, R.; Bruix, M.; Santoro, J.; Lacadena, J.; Martinez del Pozo, A.; Gavilanes, J.G.; Rico, M. NMR solution structure of the antifungal protein from Aspergillus giganteus: Evidence for cysteine pairing isomerism. Biochemistry 1995, 34, 3009–3021. [Google Scholar] [CrossRef]
  40. Batta, G.; Barna, T.; Gáspári, Z.; Sándor, S.; Kövér, K.E.; Binder, U.; Sarg, B.; Kaiserer, L.; Chhillar, A.K.; Eigentler, A.; et al. Functional aspects of the solution structure and dynamics of PAF--a highly-stable antifungal protein from Penicillium chrysogenum. FEBS J. 2009, 276, 2875–2890. [Google Scholar] [CrossRef]
  41. Berman, H.; Henrick, K.; Nakamura, H. Announcing the worldwide Protein Data Bank. Nat. Struct. Biol. 2003, 10, 980. [Google Scholar] [CrossRef] [PubMed]
  42. Utesch, T.; de Miguel Catalina, A.; Schattenberg, C.; Paege, N.; Schmieder, P.; Krause, E.; Miao, Y.; McCammon, J.A.; Meyer, V.; et al. A computational modeling approach predicts interaction of the antifungal protein AFP from Aspergillus giganteus with fungal membranes via its γ-core motif. mSphere 2018, 3, e00377-18. [Google Scholar] [CrossRef] [PubMed]
  43. Marcos, J.F.; Muñoz, A.; Pérez-Payá, E.; Misra, S.; López-García, B. Identification and rational design of novel antimicrobial peptides for plant protection. Annu. Rev. Phytopathol. 2008, 46, 273–301. [Google Scholar] [CrossRef] [PubMed]
  44. Kim, H.; Jang, J.H.; Kim, S.C.; Cho, J.H. De novo generation of short antimicrobial peptides with enhanced stability and cell specificity. J. Antimicrob. Chemother. 2014, 69, 121–132. [Google Scholar] [CrossRef] [PubMed]
  45. Behrendt, R.; White, P.; Offer, J. Advances in Fmoc solid-phase peptide synthesis. J. Pept. Sci. 2016, 22, 4–27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Duncan, V.M.S.; O’Neil, D.A. Commercialization of antifungal peptides. Fungal. Biol. Rev. 2013, 26, 156–165. [Google Scholar] [CrossRef]
  47. Oldach, K.H.; Becker, D.; Lörz, H. Heterologous expression of genes mediating enhanced fungal resistance in transgenic wheat. Mol. Plant Microbe Interact. 2001, 14, 832–838. [Google Scholar] [CrossRef] [PubMed]
  48. Vila, L.; Lacadena, V.; Fontanet, P.; Martinez del Pozo, A.; San Segundo, B. A protein from the mold Aspergillus giganteus is a potent inhibitor of fungal plant pathogens. Mol. Plant Microbe Interact. 2001, 14, 1327–1331. [Google Scholar] [CrossRef] [PubMed]
  49. Moreno, A.B.; Del Pozo, A.M.; Borja, M.; Segundo, B.S. Activity of the antifungal protein from Aspergillus giganteus against Botrytis cinerea. Phytopathology 2003, 93, 1344–1353. [Google Scholar] [CrossRef]
  50. Coca, M.; Bortolotti, C.; Rufat, M.; Peñas, G.; Eritja, R.; Tharreau, D.; del Pozo, A.M.; Messeguer, J.; San Segundo, B. Transgenic rice plants expressing the antifungal AFP protein from Aspergillus giganteus show enhanced resistance to the rice blast fungus Magnaporthe grisea. Plant Mol. Biol. 2004, 54, 245–259. [Google Scholar] [CrossRef]
  51. Theis, T.; Marx, F.; Salvenmoser, W.; Stahl, U.; Meyer, V. New insights into the target site and mode of action of the antifungal protein of Aspergillus giganteus. Res. Microbiol. 2005, 156, 47–56. [Google Scholar] [CrossRef]
  52. Szappanos, H.; Szigeti, G.P.; Pál, B.; Rusznák, Z.; Szucs, G.; Rajnavölgyi, E.; Balla, J.; Balla, G.; Nagy, E.; Leiter, E.; et al. The Penicillium chrysogenum-derived antifungal peptide shows no toxic effects on mammalian cells in the intended therapeutic concentration. Naunyn Schmiedebergs Arch. Pharmacol. 2005, 371, 122–132. [Google Scholar] [CrossRef]
  53. Moreno, A.B.; Peñas, G.; Rufat, M.; Bravo, J.M.; Estopà, M.; Messeguer, J.; San Segundo, B. Pathogen-induced production of the antifungal AFP protein from Aspergillus giganteus confers resistance to the blast fungus Magnaporthe grisea in transgenic rice. Mol. Plant Microbe Interact. 2005, 18, 960–972. [Google Scholar] [CrossRef]
  54. Szappanos, H.; Szigeti, G.P.; Pál, B.; Rusznák, Z.; Szucs, G.; Rajnavölgyi, E.; Balla, J.; Balla, G.; Nagy, E.; Leiter, E.; et al. The antifungal protein AFP secreted by Aspergillus giganteus does not cause detrimental effects on certain mammalian cells. Peptides 2006, 27, 1717–1725. [Google Scholar] [CrossRef]
  55. Moreno, A.B.; Martínez Del Pozo, A.; San Segundo, B. Biotechnologically relevant enzymes and proteins. Antifungal mechanism of the Aspergillus giganteus AFP against the rice blast fungus Magnaporthe grisea. Appl. Microbiol. Biotechnol. 2006, 72, 883–895. [Google Scholar] [CrossRef]
  56. Barna, B.; Leiter, E.; Hegedus, N.; Bíró, T.; Pócsi, I. Effect of the Penicillium chrysogenum antifungal protein (PAF) on barley powdery mildew and wheat leaf rust pathogens. J. Basic Microbiol. 2008, 48, 516–520. [Google Scholar] [CrossRef]
  57. Barakat, H.; Spielvogel, A.; Hassan, M.; El-Desouky, A.; El-Mansy, H.; Rath, F.; Meyer, V.; Stahl, U. The antifungal protein AFP from Aspergillus giganteus prevents secondary growth of different Fusarium species on barley. Appl. Microbiol. Biotechnol. 2010, 87, 617–624. [Google Scholar] [CrossRef]
  58. Palicz, Z.; Jenes, A.; Gáll, T.; Miszti-Blasius, K.; Kollár, S.; Kovács, I.; Emri, M.; Márián, T.; Leiter, E.; Pócsi, I.; et al. In vivo application of a small molecular weight antifungal protein of Penicillium chrysogenum (PAF). Toxicol. Appl. Pharmacol. 2013, 269, 8–16. [Google Scholar] [CrossRef]
  59. Delgado, J.; Acosta, R.; Rodríguez-Martín, A.; Bermúdez, E.; Núñez, F.; Asensio, M.A. Growth inhibition and stability of PgAFP from Penicillium chrysogenum against fungi common on dry-ripened meat products. Int. J. Food Microbiol. 2015, 205, 23–29. [Google Scholar] [CrossRef]
  60. Palicz, Z.; Gáll, T.; Leiter, É.; Kollár, S.; Kovács, I.; Miszti-Blasius, K.; Pócsi, I.; Csernoch, L.; Szentesi, P. Application of a low molecular weight antifungal protein from Penicillium chrysogenum (PAF) to treat pulmonary aspergillosis in mice. Emerg. Microbes Infect. 2016, 5, e114. [Google Scholar] [CrossRef]
  61. Delgado, J.; Owens, R.A.; Doyle, S.; Núñez, F.; Asensio, M.A. Quantitative proteomics reveals new insights into calcium-mediated resistance mechanisms in Aspergillus flavus against the antifungal protein PgAFP in cheese. Food Microbiol. 2017, 66, 1–10. [Google Scholar] [CrossRef]
  62. Garrigues, S.; Gandía, M.; Popa, C.; Borics, A.; Marx, F.; Coca, M.; Marcos, J.F.; Manzanares, P. Efficient production and characterization of the novel and highly active antifungal protein AfpB from Penicillium digitatum. Sci. Rep. 2017, 7, 14663. [Google Scholar] [CrossRef]
  63. Garrigues, S.; Gandía, M.; Castillo, L.; Coca, M.; Marx, F.; Marcos, J.F.; Manzanares, P. Three antifungal proteins from Penicillium expansum: Different patterns of production and antifungal activity. Front. Microbiol. 2018, 9, 2370. [Google Scholar] [CrossRef]
  64. Shi, X.; Cordero, T.; Garrigues, S.; Marcos, J.F.; Daròs, J.A.; Coca, M. Efficient production of antifungal proteins in plants using a new transient expression vector derived from tobacco mosaic virus. Plant Biotechnol. J. 2018. [Google Scholar] [CrossRef]
  65. Kovács, R.; Holzknecht, J.; Hargitai, Z.; Papp, C.; Farkas, A.; Borics, A.; Tóth, L.; Váradi, G.; Tóth, G.K.; Kovács, I.; et al. In vivo applicability of Neosartorya fischeri antifungal protein 2 (NFAP2) in treatment of vulvovaginal candidiasis. Antimicrob. Agents Chemother. 2018. [Google Scholar] [CrossRef]
  66. Kubicek, C.P.; Komoń-Zelazowska, M.; Sándor, E.; Druzhinina, I.S. Facts and challenges in the understanding of the biosynthesis of peptaibols by Trichoderma. Chem. Biodivers. 2007, 4, 1068–1082. [Google Scholar] [CrossRef]
  67. Leitgeb, B.; Szekeres, A.; Manczinger, L.; Vágvölgyi, C.; Kredics, L. The history of alamethicin: a review of the most extensively studied peptaibol. Chem. Biodivers. 2007, 4, 1027–1051. [Google Scholar] [CrossRef]
  68. Oh, S.U.; Yun, B.S.; Lee, S.J.; Kim, J.H.; Yoo, I.D. Atroviridins A-C and neoatroviridins A-D, novel peptaibol antibiotics produced by Trichoderma atroviride F80317. I. Taxonomy, fermentation, isolation and biological activities. J. Antibiot. (Tokyo) 2002, 55, 557–564. [Google Scholar] [CrossRef]
  69. Szekeres, A.; Leitgeb, B.; Kredics, L.; Antal, Z.; Hatvani, L.; Manczinger, L.; Vágvölgyi, C. Peptaibols and related peptaibiotics of Trichoderma. A review. Acta Microbiol. Immunol. Hung. 2005, 52, 137–168. [Google Scholar] [CrossRef]
  70. Heredero, M.; Garrigues, S.; Gandía, M.; Marcos, J.F.; Manzanares, P. Rational design and biotechnological production of novel AfpB-PAF26 chimeric antifungal proteins. Microorganisms 2018, 6, E106. [Google Scholar] [CrossRef]
  71. Marik, T.; Tyagi, C.; Racić, G.; Rakk, D.; Szekeres, A.; Vágvölgyi, C.; Kredics, L. New 19-residue peptaibols from Trichoderma clade Viride. Microorganisms 2018, 6, E85. [Google Scholar] [CrossRef]
  72. Delgado, J.; Rodríguez, A.; García, A.; Núñez, F.; Asensio, M.A. Inhibitory effect of PgAFP and protective cultures on Aspergillus parasiticus growth and aflatoxins production on dry-fermented sausage and cheese. Microorganisms 2018, 6, E69. [Google Scholar] [CrossRef]
  73. Váradi, G.; Tóth, G.K.; Batta, G. Structure and synthesis of antifungal disulfide β-strand proteins from filamentous fungi. Microorganisms 2019, 7, E5. [Google Scholar]
  74. Meyer, V.; Jung, S. Antifungal peptides of the AFP family revisited: Are these cannibal toxins? Microorganisms 2018, 6, E50. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Galgóczy, L.; Marx, F. Do Antimicrobial Proteins Contribute to Overcoming the Hidden Antifungal Crisis at the Dawn of a Post-Antibiotic Era? Microorganisms 2019, 7, 16. https://0-doi-org.brum.beds.ac.uk/10.3390/microorganisms7010016

AMA Style

Galgóczy L, Marx F. Do Antimicrobial Proteins Contribute to Overcoming the Hidden Antifungal Crisis at the Dawn of a Post-Antibiotic Era? Microorganisms. 2019; 7(1):16. https://0-doi-org.brum.beds.ac.uk/10.3390/microorganisms7010016

Chicago/Turabian Style

Galgóczy, László, and Florentine Marx. 2019. "Do Antimicrobial Proteins Contribute to Overcoming the Hidden Antifungal Crisis at the Dawn of a Post-Antibiotic Era?" Microorganisms 7, no. 1: 16. https://0-doi-org.brum.beds.ac.uk/10.3390/microorganisms7010016

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

Article Metrics

Back to TopTop