Skip to main content
Elsevier Sponsored Documents logoLink to Elsevier Sponsored Documents
. 2013 Jan;26(4):132–145. doi: 10.1016/j.fbr.2012.07.002

Antifungal proteins: More than antimicrobials?

Nikoletta Hegedüs 1, Florentine Marx 1,
PMCID: PMC3569713  PMID: 23412850

Abstract

Antimicrobial proteins (AMPs) are widely distributed in nature. In higher eukaryotes, AMPs provide the host with an important defence mechanism against invading pathogens. AMPs of lower eukaryotes and prokaryotes may support successful competition for nutrients with other microorganisms of the same ecological niche. AMPs show a vast variety in structure, function, antimicrobial spectrum and mechanism of action. Most interestingly, there is growing evidence that AMPs also fulfil important biological functions other than antimicrobial activity. The present review focuses on the mechanistic function of small, cationic, cysteine-rich AMPs of mammals, insects, plants and fungi with antifungal activity and specifically aims at summarizing current knowledge concerning additional biological properties which opens novel aspects for their future use in medicine, agriculture and biotechnology.

Keywords: Cysteine-rich antifungal proteins, Defensins, Development, Growth, Immunity, Signalling

Highlights

► Mammals, insects, plants and fungi produce various antimicrobial proteins (AMPs). ► This review focuses on small, cysteine-rich antifungal proteins. ► Their antifungal spectrum, structure and mode of action are described. ► Additional biological functions based on their signalling activity are highlighted. ► The multitude of functions endows AMPs with great medical/biotechnological potential.

1. Introduction

Small proteins with antimicrobial activity, so called antimicrobial proteins (AMPs), are produced by organisms throughout all kingdoms comprising prokaryotes, lower and higher eukaryotes. AMPs are secreted proteins that efficiently inhibit the growth of viruses, bacteria, fungi and parasites. In unicellular organisms, AMPs might provide their hosts the advantage to successfully compete with organisms that possess similar nutritional and ecological requirements. In multicellular organisms, AMPs constitute a primitive mechanism of innate immunity and form the first line of defence to protect their hosts from microbial attack. The innate immunity represents an evolutionarily ancient and widespread defence mechanism found in plants, insects and vertebrates. In addition, vertebrates developed the adaptive immune system – a sophisticated mechanism that uses antibodies and killer cells to recognize and eliminate invading microorganisms and allows immunological memory and self versus non-self recognition. The innate immune response is fast, and complements the adaptive immunity. Thus, both mechanisms combine to form an optimal and efficient defence system that supports the fitness of the host. The fact that closely related AMPs are widely distributed over different eukaryotic kingdoms, i.e. the class of defensins, suggests that ancestral AMP genes existed in basal eukaryotes even before fungal and insect lineages diverged (Lehrer and Ganz, 1999; Lehrer, 2007; Zhu, 2008).

AMPs are gene-encoded and they are either constitutively expressed or rapidly transcribed upon induction. In higher eukaryotes invading microbes and their products, e.g. lipopolysaccharides (Mendez-Samperio et al., 2007; Amlie-Lefond et al., 2005), or host cellular compounds, such as butyrate (Murakami et al., 2002), cytokines (Wolk et al., 2004; Wilson et al., 2007; Lai and Gallo, 2009), and vitamins (Schauber et al., 2006) stimulate AMP production.

Due to the vast variety in function, structure, expression pattern, target organisms and producing hosts, the classification of AMPs is difficult and somewhat arbitrary to date. Mostly, AMPs are classified according to their functional and/or structural properties. Both characteristics are determined by the primary sequence of the protein which very often shows a high number of certain amino acids such as glycine, cysteine, histidine, proline, tyrosine, arginine, lysine and serine.

Although only a defined number of AMPs has been structurally analysed by nuclear magnetic resonance (NMR), calorimetric dichroism (CD) or X-ray crystallography, the secondary and tertiary structure of numerous AMPs has been predicted from primary sequence homologies. The most common classes contain proteins with α-helical, β-sheet or mixed α-helical/β-sheet structures (Zhu, 2008; Dimarcq et al., 1998; Giangaspero et al., 2001; Tossi et al., 2000).

The best functionally characterized AMPs are bactericidal, whereas the properties of antifungal AMPs and their mode-of-action are less well studied. In the online database of AMPs at http://aps.unmc.edu/AP/main.php around 1900 AMPs of different origin are registered. From these, more than 1500 AMPs (79 %) have been assigned antibacterial activity compared to 648 antifungal AMPs (34 %). This classification, however, is redundant and antibacterial AMPs may also show antifungal activity that has not been investigated so far.

Most interestingly, the number of reports that document new additional functions of AMPs beyond their antimicrobial activity is constantly increasing. These features might arise from signalling functions of AMPs that accompany the activity of AMPs to interfere with the cell proliferation of microbes. For example, antimicrobial peptides from bacteria are part of the quorum-sensing mechanism that helps microorganisms to communicate and co-ordinate their behaviour by accumulating signalling molecules in the extracellular environment. This microbial communication system regulates e.g. symbiosis, biofilm formation, conjugation, sporulation, virulence, motility and the production of various secondary metabolites (Raina et al., 2009; Maroti et al., 2011). In fungi, plants and insects, AMPs have been related to development and differentiation (Stotz et al., 2009a,b; Eigentler et al., 2012; Hegedus et al., 2011a), symbiotic interaction (Maroti et al., 2011) and root hair extension (Allen et al., 2008). In vertebrates, especially mammals, most diverse biological effects of AMPs have been reported, such as endotoxin neutralization (Rosenfeld et al., 2010), signalling (Salzet, 2002), regulation of the immune response by chemotactic and immunomodulating activities (Wiesner and Vilcinskas, 2010; Yang et al., 2002), induction of angiogenesis and wound repair (Baroni et al., 2009) and protection against cancer (Yang et al., 2002).

This review focuses on the properties and mode-of-action of small (5–8 kDa), cysteine-rich antifungal AMPs produced in mammals, insects, plants and fungi of related structure and highlights additional biological functions apart from antimicrobial activity. We want to apologize in advance for not being able to refer to all excellent publications available in this field due to space limitation. Readers specifically interested in antibacterial AMPs are directed to other excellent reports and reviews (Lehrer, 2007; Wiesner and Vilcinskas, 2010; Bulet and Stocklin, 2005; Yang et al., 2004; Selsted and Ouellette, 2005; Lay and Anderson, 2005; Wong et al., 2007; Papagianni, 2003; Boman, 2003; Brogden, 2005; Reddy et al., 2004; Guani-Guerra et al., 2010; Taylor et al., 2008; Mygind et al., 2005).

2. A short general overview on the structure and mode-of-action of antifungal cysteine-rich AMPs

The most prominent group within the antifungal AMPs with a close structural relationship constitutes defensins from plants, insects and mammals (Taylor et al., 2008; Aerts et al., 2008). Defensins contain six to eight cysteines which form intramolecular disulfide bonds and stabilize an antiparallel β-sheet conformation flanked by an α-helical segment, also called cysteine stabilized α/β motif (CSα/β) (Bulet and Stocklin, 2005; Selsted and Ouellette, 2005; Taylor et al., 2008). Defensin-like antifungal AMPs of fungal origin are steadily increasing in number and show similar structural features as defensins but lack the α-helix (Campos-Olivas et al., 1995; Batta et al., 2009). The compact structure of defensins and defensin-like AMPs confers resistance towards extreme temperature, pH and protease-mediated degradation (Batta et al., 2009; Bulet et al., 1999; Landon et al., 1997; Hajji et al., 2010).

Based on their cationic character, defensins are thought to interact with negatively charged plasma membrane components of sensitive microorganisms. Two general models try to explain the mechanism of antimicrobial action of defensins: (A) the permeabilization of the cell membrane by (i) the carpet model and (ii) the pore model (Brogden, 2005). Whereas in (i) several protein molecules insert into the membrane forming a pore, in (ii) the protein molecules oligomerize and form a multimeric pore. Both models are primarily based on AMP-bacterial interaction and describe the disintegration of the plasma membrane, cell leakage and cell death by necrosis. (B) Alternatively, the membrane interaction of defensins and defensin-like AMPs may not primarily damage the plasma membrane of target cells. Instead, antifungal protein interaction with specific lipid and/or protein components of the plasma membrane leads to a selective ion permeability of the membrane and to the formation of transient pores and/or results in (active) protein transport into the host cell where these antifungals interact with intracellular targets (Brogden, 2005; Marx et al., 2008; Thevissen et al., 2003a). This antifungal activity increases the intracellular level of reactive oxygen species (ROS) and triggers programmed cell death (PCD) (Leiter et al., 2005; Aerts et al., 2011, 2009). Thus, plasma membrane leakage could occur at a later time point after protein contact with the fungal cell as a secondary effect of extensive intracellular ROS formation, as proposed for plant defensins (Thevissen et al., 2003a).

3. Mammalian AMPs

Among mammalian AMPs, the defensins are best characterized and categorized in three subfamilies, α-, β- and θ-defensins. Whereas the θ-defensins were primarily found in primates (Schutte et al., 2002), α- and β-defensins are produced by humans and differ in the position of the disulfide bridges (Schneider et al., 2005). α-defensins are products of neutrophils and Paneth cells of the small intestine (Mallow et al., 1996; Ganz et al., 1985) and human β-defensins 1–4 (hBD1, hBD2, hBD3, hBD4) are secreted by epidermal, epithelial and mesenchymal cells (Ali et al., 2001; Varoga et al., 2006; Zhao et al., 1996). We will focus on this latter defensin group hBD1-4 that shows the closest homology to insect and plant defensins. Six conserved cysteine residues form three disulfide bridges and arrange three antiparallel β-sheets flanked by an α-helical segment (Pazgier et al., 2006). Apart from their toxicity against gram-positive and gram-negative bacteria and viruses, these proteins possess antifungal activity against Candida albicans, Candida krusei and Candida parapsilosis (Feng et al., 2005; Hoover et al., 2003; Hazrati et al., 2006; Sun et al., 2005). β-defensins form the first line of defence against microbial infection and link innate with adaptive immunity. They are constitutively expressed or are induced by microbes, cytokines (TNF-α, IL-1) or even by the status of cellular differentiation (Liu et al., 2002). Near the site of microbial invasion β-defensins form chemotactic gradients to recruit immature dendritic cells and memory T cells (Yang et al., 2002, 1999). Immunomodulation is regulated via the Toll signalling pathway. For example, the murine β-defensin 2 (mDF2beta) is an endogenous ligand to the Toll-like receptor 4 and induces dendritic cell maturation (Biragyn et al., 2002).

Apart from functions in the immune response, β-defensins regulate other most diverse cellular processes. For example, β-defensins affect sperm function including the initiation of motility and capacitation (Tollner et al., 2004; Yudin et al., 2005; Zhou et al., 2004). hBD2 activates and triggers the degranulation of mast cells, which results in the release of histamine and prostaglandin D2, two messenger molecules that regulate many physiological processes (Befus et al., 1999; Niyonsaba et al., 2001). hBD2 also stimulates the migration, proliferation and tube formation of endothelial cells in wounds and accelerates wound closure (Baroni et al., 2009). Furthermore, hBD2–4 promote wound healing by supporting the migration and proliferation of keratinocytes (Niyonsaba et al., 2007).

Notably, human β-defensins also have been associated with the pathogenesis of diseases. For example, hBD2 and hBD3 were originally purified from inflammatory skin disease psoriasis and proinflammatory cytokines induce their transcription (Harder et al., 2001), whereas the loss of hBD1 favours cancer development (Sun et al., 2006). hBD2 has been implicated in Crohn's disease (Fellermann et al., 2006).

Thus defensins act not only as potent antimicrobial agents, but also as immunological adjuvants, similar to chemokines (Salzet, 2002), and as signalling molecules which regulate important cellular functions, ultimately contributing to the fitness of the host.

4. Insect AMPs

The antibacterial features of insect defensins are well documented (Bulet et al., 1999; Cociancich et al., 1993; Lehrer and Ganz, 1996; Otvos, 2000), whereas the number of antifungal insect defensins comprises only four members which show a high structural relatedness to mammalian and plant defensins: drosomycin from the fruit fly Drosophila melanogaster (Fehlbaum et al., 1994; Zhang and Zhu, 2009), heliomicin from the tobacco budworm Heliothis virescens (Lamberty et al., 1999), termicin from the fungus-growing termites Pseudocanthothermes spiniger and Reticulitermes flavipes (Lamberty et al., 2001) and gallerimycin from the greater wax moth larvae Galleria mellonella (Lee et al., 2004; Schuhmann et al., 2003). Most insect AMPs are produced after microbial challenge in the fat body, equivalent to the mammalian liver, and certain haemocytes before they are released into the hemolymph.

Only modest information exists on gallerimycin and heliomicin. Gallerimycin has antifungal activity against Aspergillus niger and the entomopathogenic fungus Metharhizium anisopliae, but not against yeasts and bacteria (Schuhmann et al., 2003; Mak et al., 2010). Similarly, heliomicin is inactive against bacteria, but toxic against yeasts like C. albicans, Pichia pastoris and Cryptococcus neoformans and filamentous fungi such as Neurospora crassa, Fusarium sp., Aspergillus fumigatus, Trichoderma viridae (Lamberty et al., 1999). In analogy to plant defensins (RsAFP2), heliomicin directly binds to fungal plasma membrane ceramides, i.e. glucosylceramides. Consequently, yeast deletion mutants in the glucosylceramide biosynthesis gene GCS1 were 20-fold less sensitive than the wild-type strains (Thevissen et al., 2004).

The mode-of-action of drosomycin and termicin is better described. Although drosomycin is strictly antifungal at micromolar concentrations against filamentous fungi (N. crassa, Fusarium sp., Alternaria sp., Botrytis cinerea) and yeasts (Saccharomyces cerevisiae) (Zhang and Zhu, 2009; Tian et al., 2008), bacterial challenge can induce drosomycin gene expression. The systemic gene expression is regulated by the Toll pathway (Zhang and Zhu, 2009). In addition, a local response in epithelia situated in the respiratory, digestive and reproductive tracts can also induce drosomycin expression independently of the Toll pathway (Ferrandon et al., 1998; Tzou et al., 2000), whereby the inducible response in the respiratory tract is controlled by the Immune Deficiency (IMD) pathway (Zhang and Zhu, 2009).

Drosomycin lyses fungal hyphae suggesting a mode-of-action based on membrane permeabilization and pore formation in susceptible fungi (Fehlbaum et al., 1994). Membrane components, such as sphingolipids, seem to play a central role for the interaction of drosomycin with its target organisms (Gao and Zhu, 2008). Indeed, a N. crassa mutant (Mut16) displaying clear differences in sphingolipid composition compared to the wild-type strain, was less susceptible to drosomycin (Ferket et al., 2003). Most interestingly, drosomycin closely interferes with the D. melanogaster voltage-gated sodium channel (DmNav1). This observation and the detection of drosomycin transcripts in the Drosophila brain and thoracic-abdominal ganglia (Jiggins and Kim, 2005) further support the assumption that drosomycin might act as neuropeptide in the fly's central nervous system (Cohen et al., 2009).

Tian et al. (2008) discovered for the first time that the antifungal drosomycin also exhibits antiparasitic activity. The observation that drosomycin inhibited the development of the ookinetes of Plasmodium berghei undoubtedly opens a new research field for the design of novel anti-malaria agents.

The termite antifungal defensin termicin inhibits the growth of the ubiquitous soil and entomopathogenic fungus M. anisopliae (Hamilton et al., 2011). The termite Reticulitermes produces termicin together with antifungal glucanases, e.g. the gram-negative bacteria binding protein 2 (GNBP2), which efficiently breaks down β-1,3-glucans of fungal cell walls (Bulmer et al., 2009). Termicin and GNBP2 are constitutively expressed in the salivary gland, stored in reservoirs and constitutively secreted and disseminated by self- and allo-grooming. The secretion of termicin and synergistically acting enzymes prevent M. anisopliae from attaching and penetrating the cuticle where it can effectively evade the insect immune defense (Hamilton and Bulmer, 2012). Furthermore, Reticulitermes incorporates termicin and GNBP2 in nest materials that may act as an external recognition system. The release of digested fungal cell wall material may trigger the immune defence of termites that are working in infested areas of their nests or support the accession of foraging grounds (Bulmer et al., 2009). Thus the secretion of antifungal proteins, such as termicin, could have been crucial in overcoming strong pathogenic pressures during social evolution in insects (Hamilton and Bulmer, 2012).

5. Plant AMPs

The plant defensin family is quite diverse in its primary structure as only eight structure-stabilizing cysteines are conserved among all plant defensins (Thevissen et al., 2003a). Reflecting the structural diversity, plant defensins display most diverse biological activities. Unlike the insect and mammalian defensins, which are mainly active against prokaryotes, only a few plant defensins have been shown to possess antibacterial activity (Stotz et al., 2009b; Koike et al., 2002; Segura et al., 1998). Most plant defensins are active against numerous filamentous ascomycetes (Carvalho Ade and Gomes, 2009), including phytopathogenic fungi such as Fusarium graminearum (Sagaram et al., 2011), B. cinerea (Stotz et al., 2009a), and Alternaria brassicicola (de Zelicourt et al., 2007). Some plant defensins inhibit the growth of baker's yeast and opportunistic human pathogenic fungi such as C. albicans (Thevissen et al., 2004; de Zelicourt et al., 2007). Plant defensins are divided into two groups: (i) antifungal activity is accompanied by changes in morphology such as hyperbranching of hyphae; (ii) antifungal activity occurs without hyphal distortions (Thomma et al., 2002). Consequently, it is assumed that differences exist in the mode-of-action of these closely related antifungal proteins.

Defensins are expressed by various organs and tissues, e.g. in seeds and in peripheral cell layers of fruits and floral organs, including the style, the ovary and the stamen filaments, of different plants like Arabidopsis (Penninckx et al., 1996), pea (Almeida et al., 2000), spinach (Segura et al., 1998), tobacco (Allen et al., 2008; Lay et al., 2003a,b) and radish (Terras et al., 1992). The peripheral location of plant defensins in different generative tissues reflects their defence role against microbial invasion, either as inducible antimicrobial agents within vegetative tissues or as part of the constitutive defence barriers, especially in storage and reproductive organs (Thevissen et al., 2003a). In addition some defensin genes are systemically induced upon fungal infection or wounding of vegetative tissues such as pI39 and pI230 in pea (Chiang and Hadwiger, 1991).

Among all antifungal defensins, the mode-of-action of plant defensins has been exceptionally well characterized at the molecular level. The following examples are given: the seed defensins DmAMP1 from dahlia (Dahlia merckii) (Osborn et al., 1995), RsAFP2 from radish (Raphanus sativus) (Terras et al., 1992), HsAFP1 from coral bells (Heuchera sanguinea) (Osborn et al., 1995), MsDef1 from alfalfa (Medicago sativa) (Gao et al., 2000), MtDef2 from barrel medic (Medicago truncatula) (Spelbrink et al., 2004) and Psd1 from pea (Pisum sativum) (Almeida et al., 2000). There is growing evidence that excludes a direct interaction of the cationic defensins with the negatively charged phospholipids of the plasma membrane as well as pore formation (Thevissen et al., 1996; Caaveiro et al., 1997). Instead, recent data underscore the existence of specific binding domains on the fungal plasma membrane, so called lipid rafts, which are specifically enriched in sphingolipids and sterols and which allow a high local concentration of membrane bound defensins (Thevissen et al., 2003a). For example, DmAMP1 and RsAFP2 interact with sphingolipids of the fungal plasma membrane, DmAMP1 with mannosyldiinositolphosphoryl-ceramide M(IP)2C, an acid complex sphingolipid (Thevissen et al., 2000, 2005), and RsAFP2 with a neutral sphingolipid class of glucosylceramides (GlcCer) (Thevissen et al., 2004). Notably, equimolar concentrations of ergosterol and sphingolipids enhance the interaction of DmAMP1, but not of RsAFP2 (Thevissen et al., 2004, 2003b). Similar binding activities as for RsAFP2 have been proposed for MsDef1 (Ramamoorthy et al., 2007a). HsAFP1 also interacts with specific, so far unidentified high affinity binding sites on plasma and microsomal membranes (Thevissen et al., 1997). The interaction of DmAMP1 and RsAFP2 with sphingolipid components of the fungal plasma membrane results in an increased Ca2+ influx and K+ efflux (Thevissen et al., 1996). The question whether fungal growth arrest is a direct consequence of increased membrane permeability or of the interaction of plant defensins with intracellular targets has not been fully answered yet. Nevertheless, existing data suggest that more complex mechanisms than membrane permeabilization are involved in cell death induction. RsAFP2 and HsAFP1 for example, increase the level of intracellular ROS in C. albicans (Aerts et al., 2011, 2007) which contributes to the induction of PCD (Aerts et al., 2008). Environmental stress signalling cascades, such as the mitogen-activated protein kinase (MAPK) pathway may also be activated (Aerts et al., 2011). The Gpmk1 and Mgv1 MAPK signalling were involved in the response to RsAFP2, MsDef1 and MtDef2 in F. graminearum (Aerts et al., 2008; Ramamoorthy et al., 2007b). Using a yeast two-hybrid screen the cell cycle control related protein cyclin F was found to interact with the P. sativum defensin Psd1 and further experiments proved the co-localization of Psd1 with the nucleus and the inhibition of cell division by Psd1 (Lobo et al., 2007).

Apart from their antimicrobial defence function, defensins also fulfil other important roles in plant physiology. An exceptionally well researched overview on the complex functions of plant defensins is presented by Carvalho Ade and Gomes (2009). For example, plant defensins mediate resistance against abiotic stress. The Arabidopsis halleri defensin AhPDF1.1 confers zinc- and a slight selenite-tolerance in Arabidopsis sp. (Oomen et al., 2011; Mirouze et al., 2006; Tamaoki et al., 2008).

Other features of plant defensins are related to the regulation of growth, development and fertilization. The supplementation of the culture medium with defensins MsDef1, MtDef2, RsAFP2 or AhPDF1.1 inhibits plant root growth (Allen et al., 2008; Oomen et al., 2011). The Solanum lycopersicon DEF2 is produced during early flower development and promotes meiosis. As such it plays a central role in regulating the growth of different plant organs. During later stages of flower differentiation the inactivation of DEF2 is crucial for survival and development of pollen grains. Consequently, antisense suppression or constitutive overexpression of DEF2 negatively interferes with pollen viability and seed production (Stotz et al., 2009a,b).

Fertilization in flowering plants involves distinct pollen–pistil interactions that include determinants of self-incompatibility, factors for pollen germination and tube growth, pollen tube attractants and bursting (Amien et al., 2010; Higashiyama, 2010; Okuda et al., 2009; Schopfer et al., 1999; Takayama et al., 2001). The Zea mays defensin ZmES4 plays a central role in inter-gametophyte signalling. It mediates pollen tube bursting by activating the K+ channel KZM1 (Amien et al., 2010).

Antifungal plant defensins are also involved in the defence against parasitic plants and insects. The sunflower defensin Ha-DEF1 specifically combats invasion of the parasitic plant Orobanche cumana. When Ha-DEF1 is exogenously applied to the parasitic plant embryo, it causes browning symptoms at the radicle apex of Orobanche seedlings during the germination step, but has necrotic effects neither on other parasitic plants, e.g. Striga hermonthica, nor on the host plant Arabidopsis thaliana (de Zelicourt et al., 2007). The defensin VrD1 from Vigna radiate seeds exhibits insecticidal activity by inhibiting the insect α-amylase that digests plant starch in the insect gut (Carvalho Ade and Gomes, 2009; Liu et al., 2006).

Finally, the inhibitory potential of antifungal plant defensins towards certain human cancer cell types and viruses has also been reported. Sesquin from the ground bean Vigna sesquipedalis shows anti-proliferative activity towards leukaemia M1 cells and breast cancer (MCF-7) cells and inhibits human immunodeficiency virus-type 1 reverse transcriptase (Wong and Ng, 2005). The Phaseolus limensis defensin Limyin has anti-cancer activity (Wang et al., 2009).

6. AMPs of fungal origin

The number of identified small, cysteine-rich defensin-like AMPs in ascomycetes is constantly increasing, but only few have been characterized in more detail, e.g. AFP from Aspergillus giganteus strain lfGB0203 (Wnendt et al., 1994) and AFPNN5353 from A. giganteus strain A3274 (Binder et al., 2011), PAF from Penicillium chrysogenum strain Q176 (Marx et al., 1995), PgAFP from P. chrysogenum strain RP42C (Rodríguez-Martín et al., 2010), strain bubble protein (BP) from Penicillium brevicompactum Dierckx (Seibold et al., 2011), ANAFP from A. niger (Gun Lee et al., 1999), NAF from Penicillium nalgiovense (Geisen, 2000), AcAFP and AcAMP from Aspergillus clavatus (Skouri-Gargouri and Gargouri, 2008) and NFAP from Neosartorya fisheri (Kovacs et al., 2011). All these proteins exclusively exhibit antifungal activity against filamentous ascomycetes at micromolar concentrations, including opportunistic plant- and animal pathogens, such as Fusarium sp., Botrytis sp., and A. fumigatus (Marx, 2004; Meyer, 2008). Moreover, the P. chrysogenum PAF inhibits the growth of selected Zygomycete species (Galgoczy et al., 2005). Further exceptions are (i) the A. clavatus AcAMP with antibacterial activity (Hajji et al., 2010) and (ii) the A. niger AnAFP with anti-yeast activity (Gun Lee et al., 1999). (iii) The recently characterized P. brevicompactum BP only weakly affects yeast growth (Seibold et al., 2011). To date, PAF is one of the best characterized antifungal proteins of fungal origin besides its orthologous AFP from A. giganteus (Marx, 2004; Meyer, 2008; Marx et al., 2008), and both exhibit functional similarities with plant defensins to a certain extent.

Although direct evidence is lacking, it is believed that the toxicity of many of these fungal AMPs is based on their interaction with molecules/receptors situated in the cell wall and/or the plasma membrane of target fungi from where signalling cascades are activated. In the case of PAF, heterotrimeric G-protein and cAMP/protein kinase A signalling are involved, whereas AFP activates the cell wall integrity pathway (Leiter et al., 2005; Binder et al., 2011, 2010a; Hegedus et al., 2011b; Hagen et al., 2007; Ouedraogo et al., 2011). The interaction of PAF with the sensitive Aspergillus nidulans leads to an immediate hyperpolarization of the plasma membrane at hyphal tips accompanied by K+ efflux and Ca2+ influx (Leiter et al., 2005; Binder et al., 2010b; Kaiserer et al., 2003) which parallels the effect observed with plant defensins RsAFP2 and DmAMP1 (Thevissen et al., 1996). Changes in the membrane potential might influence the activity of ion channels/pumps/transporters or be a result of their activation. We could prove that the P. chrysogenum PAF and the A. giganteus AFPNN5353 trigger a rapid Ca2+ influx which leads to a sustained elevation of the cytoplasmic Ca2+ resting level in the sensitive fungi N. crassa and A. niger, respectively. This Ca2+ influx, however, does not result from unspecific pore formation in the plasma membrane (Binder et al., 2011, 2010b). It is well known that a sustained perturbation of Ca2+ homeostasis may trigger PCD. Intracellular ROS and apoptotic markers are indeed increased in PAF-treated sensitive fungi (Leiter et al., 2005). Interestingly, supplementation of the growth medium with Ca2+ ions, decreases the toxicity of antifungal proteins and counteracts the perturbation of the intracellular Ca2+ resting level. This points towards the induction of adaptive responses that may be induced by ion/Ca2+ signalling (Binder et al., 2011, 2010b; Ouedraogo et al., 2011). Even though PAF and AFP exhibit close similarities in their structure, function and antimicrobial spectrum, they also show significant differences in their mechanisms of action. AFP modulates the cell wall composition of sensitive fungi by enhancing α-1,3-glucan synthase A (agsA) and the chitin synthase D (chsD) gene expression (Hagen et al., 2007) which suggests an attempt by the fungus to counteract AFP attack. As evidenced by the close AFP relative AFPNN5353, agsA induction is mediated via protein kinase C (Pkc)/MAPK A. Instead, PAF fails to trigger the cell wall integrity pathway (CWIP) and PkcA/MAPK A signalling is not activated (Binder et al., 2010a).

Diverging data also exist about the localization of these two antifungal proteins. Whereas Theis et al. (2005, 2003) reported AFP to localize to the cell wall and/or the plasma membrane of sensitive fungi, a time- and concentration-dependent internalization and cytoplasmic localization of AFP was shown by Martin-Urdiroz et al. (2009). Another study showed the uptake of Alexa-labelled AFP and its co-localization with the nucleus (Moreno et al., 2006). Instead, we detected the energy-dependent endocytic uptake of the P. chrysogenum PAF into the cytoplasm of sensitive fungi, but no co-localization with the nucleus (Batta et al., 2009; Oberparleiter et al., 2003). A similar uptake mechanism and staining pattern in sensitive fungi was shown for the AFP relative AFPNN5353 (Binder et al., 2011).

Most interestingly, there is increasing evidence that antifungal proteins, similar to defensins from plants, insects and mammals, exhibit additional functions that go beyond antimicrobial activity. The P. chrysogenum paf and the A. giganteus afp exhibit a temporal and spatial regulation during asexual differentiation (Hegedus et al., 2011a; Meyer et al., 2002). Both genes are exclusively expressed in the vegetative mycelium at the developmental stage. No gene expression occurs in conidiophores or conidia (Hegedus et al., 2011a; Meyer, 2008; Meyer et al., 2002). Key regulators for asexual development and/or secondary metabolism such as StuA and VelA (Adams et al., 1998; Adams and Yu, 1998) were suggested to play a role in afp and paf gene expression. As for afp, StuA is positively involved in afp expression in A. giganteus (Meyer, 2008), whereas for paf, the P. chrysogenum VelA seems to play a central role in paf gene induction (Hegedus et al., 2011a; Hoff et al., 2010). Notably, depletion of PAF results in the down-regulation of brlA that encodes the central regulator for mitospore development and spore production is severely reduced in the Δpaf strain compared to the wild type.

Moreover, expression of the A. giganteus afp is inducible by alkaline growth conditions, carbon starvation, excess NaCl and ethanol and heat-shock all of which point towards a stress-related expression pattern (Meyer et al., 2002; Meyer and Stahl, 2002).

Taking advantage of a growing number of sequenced fungal genomes and using a computional approach, Zhu (2008) postulated the existence of 25 fungal CSα/β-defensins. We characterized one of these potential fungal defensins, Anisin1 from A. nidulans. Depletion of Anisin1 negatively affects brlA expression and asexual spore production. Furthermore, the deregulation of the major scavenger against H2O2, CatB, in a Δanisin1 mutant suggests a role of Anisin1 in sensing/signalling elevated H2O2 levels (Eigentler et al., 2012). Although the antimicrobial activity of Anisin1 has not been investigated to date, we provided evidence that Anisin1 has an important function in improving the fitness of A. nidulans by linking stress signalling with developmental regulation (Eigentler et al., 2012).

7. Conclusion

As summarized in Table 1, antifungal defensins and defensin-like AMPs exhibit complex biological functions in addition to antimicrobial toxicity. The multitude of biological functions undoubtedly underscores the great impact of these proteins on our life. They gain great medical and biotechnological significance as attractive candidates for drug development and combinatorial therapies. In addition, they allow crop improvement by the generation of transgenic plants with enhanced resistance against phytopathogenic fungi, insects and plant parasites (Terras et al., 1995). Notably, the recent discovery that antifungal AMPs of fungal origin also cover sensing/signalling functions contributes to our knowledge on the regulation of fungal development and may be of extraordinary importance for the optimization of fermentation processes and for the production of secondary metabolites and other fungal products.

Table 1.

Summary of the current knowledge regarding the mechanisms of antifungal action and additional biological functions of selected antifungal defensins and defensin-like AMPs from fungi, plants, insects and mammals.

AMP producing organism Interaction molecules of the fungal envelope Cellular uptake Mechanism of action and signalling pathways involved in fungi–AMP interaction Biological functions beyond antimicrobial activity References
Fungi PAF Penicillium chrysogenum Q176 n.d. + Plasma membrane hyperpolarization, K+ efflux, Ca2+ influx, perturbation of the intracellular Ca2+ homeostasis, cAMP/PkA signalling, generation of intracellular ROS, apoptosis Signalling, supports asexual development (Hegedus et al., 2011a; Leiter et al., 2005; Binder et al., 2010a; Binder et al., 2010b; Kaiserer et al., 2003; Oberparleiter et al., 2003)
AFP Aspergillus giganteus lfGB0203 n.d. ± Plasma membrane permeabilization, activation of the cell wall integrity pathway n.d. (Hagen et al., 2007; Ouedraogo et al., 2011; Theis et al., 2005, 2003)
AFPNN5353Aspergillus giganteus A3274 n.d. + Ca2+ influx, perturbation of the intracellular Ca2+ homeostasis, MAPK signalling, activation of the cell wall integrity pathway n.d. (Binder et al., 2011)
Anisin1 Aspergillus nidulans n.d. n.d. n.d. Support of asexual development, oxidative stress signalling (Eigentler et al., 2012)
Plants DmAMP1 Dahlia merckii Sphingolipid M(IP)2C n.d. K+ efflux, Ca2+ uptake, alkalinization of the medium, membrane potential changes, membrane permeabilization n.d. (Thevissen et al., 1996, 2000, 2005, 2003b, 1999)
RsAFP2 Raphanus sativus Sphingolipid GlcCer n.d. K+ efflux, Ca2+ uptake, alkalinization of the medium, membrane potential changes, plasma membrane permeabilization, induction of ROS accumulation, MAPK signalling Plant root growth inhibition (Allen et al., 2008; Thevissen et al., 1996, 2004, 1999; Aerts et al., 2007; Ramamoorthy et al., 2007b)
HsAFP1 Heuchera sanguinea Interaction with plasma and microsomal membranes n.d. Plasma membrane permeabilization, ROS accumulation, apoptosis, MAPK signalling n.d. (Aerts et al., 2011; Thevissen et al., 1997, 1999)
MsDef1 Medicago sativa Sphingolipid GlcCer n.d. MAPK signalling Mammalian Ca2+ channel blocker, plant root growth inhibition (Allen et al., 2008; Spelbrink et al., 2004; Ramamoorthy et al., 2007a,b)
MtDef2 Medicago truncatula n.d. n.d. MAPK signalling Plant root growth inhibition (Allen et al., 2008; Ramamoorthy et al., 2007b)
Psd1 Pisum sativum n.d. + Interaction with cyclin F and interferes with cell division n.d. (Lobo et al., 2007)
AhPDF1.1 Arabidopsis halleri n.d. n.d. n.d. Tolerance against zinc and selenite, inhibition of plant root growth (Oomen et al., 2011; Mirouze et al., 2006; Tamaoki et al., 2008)
DEF2 Solanum lycopersicon n.d. n.d. n.d. Influence of pollen viability, seed production, and the growth of various organs of the producing plant (Stotz et al., 2009a)
ZmES4 Zea mais n.d. n.d. n.d. Inter-gametophype signalling, induction of pollen tube burst (Amien et al., 2010)
Ha-DEF1 Helianthus inbred line (LR1) n.d. n.d. Plasma membrane permeabilization Parasitic plant defence (de Zelicourt et al., 2007)
VrD1 Vigna radiata n.d. n.d. n.d. Insect inhibitory activity (Liu et al., 2006; Chen et al., 2004, 2002)
Sesquin Vigna sesquipedalis n.d. n.d. n.d. Inhibitory activity against cancer cells, inhibitory effect towards HIV reverse transcriptase (Wong and Ng, 2005)
Limyin Phaseolus limensis n.d. n.d. n.d. Inhibitory activity against cancer cells (Wang et al., 2009)
Insects Drosomycin Drosophila melanogaster Sphingolipids n.d. Plasma membrane permeabilization, pore formation Interaction with voltage-gated sodium channel (DmNav1), neuropeptide activity? (Gao and Zhu, 2008; Ferket et al., 2003; Jiggins and Kim, 2005)
Heliomicin Heliothis virescens Sphingolipid GlcCer n.d. n.d. n.d. (Lamberty et al., 1999; Thevissen et al., 2004)
Termicin Pseudocanthothermes spiniger, Reticulitermes flavipes Cell wall n.d. Acts synergistically with glucanases to break down β-glucans of the fungal cell wall Immune defence, external recognition system, accession of foraging grounds (Bulmer et al., 2009; Hamilton and Bulmer, 2012)
Mammals hBD1 Homo sapiens n.d. n.d. Plasma membrane permeabilization, pore formation Chemotactic for immature dendritic cells, suppression of cancer development (Sun et al., 2005; Yang et al., 1999, 2000)
hBD2 Homo sapiens n.d. n.d. Plasma membrane permeabilization, pore formation Chemotactic for CD45R0+ memory T cells, chemotactic for immature dendritic cells, degranulation of mast cells, migration and proliferation of keratinocytes, promotion of wound healing (Baroni et al., 2009; Yang et al., 1999, 2000; Befus et al., 1999; Niyonsaba et al., 2001; Niyonsaba et al., 2006)
hBD3 Homo sapiens n.d. n.d. Plasma membrane permeabilization, pore formation Chemotactic for immature dendritic cells, mast cell activation, migration and proliferation of keratinocytes, promotion of wound healing (Harder et al., 2001; Niyonsaba et al., 2006; Chen et al., 2007)
hBD4 Homo sapiens n.d. n.d. Plasma membrane permeabilization, pore formation Mast cell activation, migration and proliferation of keratinocytes, promotion of wound healing (Niyonsaba et al., 2006; Chen et al., 2007)
mDF2beta Mus musculus n.d. n.d. Plasma membrane permeabilization, pore formation Maturation of dendritic cells (Biragyn et al., 2002)

n.d. – not determined.

cellular uptake: − no uptake; + intracellular localization.

It is, however, most puzzling how one protein may display such functional diversity. One explanation could be its specific protein structure. The disulfide bonds between the cysteine residues not only stabilize a compact structure, but they may be sensitive to specific environmental or cellular redox conditions. These may influence the redox status of the protein itself and finally determine its conformation and binding affinity with target molecules and even promote protein dimerization or oligomerization (Batta et al., 2009; Wu et al., 2003; Hoover et al., 2000, 2001). In this respect it may be assumed that the increase of ROS concentrations in the cytoplasm of sensitive fungi is triggered by the redox activity of antifungal defensins and defensin-like AMPs (Hegedus et al., 2011b).

Especially the occurrence of antifungal AMPs in biotechnologically important fungi suggests a simple protein preparation from easy-to-handle microorganisms that are partially considered as “safe” by the Food and Drug Administration. However, increasing knowledge about the structure–function relation of defensins and defensin-like AMPs opens new opportunities to improve their function and their (selective) antimicrobial activity, or to introduce new activities (Francois et al., 2005), for example, by the substitution of single amino acids or by the de novo synthesis of non-natural peptides. The latter uncouples protein production from biological systems and a nice example is given by characterization of the synthetic antifungal hexapeptide PAF26 (RKKWFW) (Munoz et al., 2006, 2007, 2012; Lopez-Garcia et al., 2007).

AMPs belong to a fast growing scientific field and the identification of novel antifungal proteins may be expected in the future. By a detailed characterization of the biological functions of antifungal AMPs, we will undoubtedly gain deeper insight into their modes-of-action. This will not only complete our understanding of specific cellular/biological processes but also provide an excellent basis for new approaches in medicine, agriculture and biotechnology.

Acknowledgements

We thank Doris Bratschun for technical assistance in the research of the P. chrysogenum antifungal protein PAF, the A. giganteus AFPNN5353 and the A. nidulans Anisin1 protein. We thank the Austrian Science Fund (FWF 19970), the D. Swarowski Forschungsförderungsfonds (FB2/06) and the program of the Österreichischer Austauschdienst ÖAD (79öu1 and SI15/2009) for financial support. Nikoletta Hegedüs was financially supported by the Tiroler Wissenschaftsfonds (TWF UNI-0404/966).

Biographies

graphic file with name fx1.jpg

Nikoletta Hegedüs graduated in June 2005 as Environmental Researcher (MSc) at the Faculty of Science, University of Debrecen, Hungary. She completed her PhD in Biology in September 2011 at the Department of Microbial Biotechnology and Cell Biology, University of Debrecen, Hungary. Since October 2011 she holds a postdoctoral position in the research group of Florentine Marx at the Division of Molecular Biology, Biocenter, Innsbruck Medical University, Austria. Main area of research: the molecular mechanism of action of small cysteine-rich antifungal proteins produced by fungi with special emphasis on their sensing/signalling function.

graphic file with name fx2.jpg

Florentine Marx habilitated in Microbiology at the Innsbruck Medical University, Austria. Since 2006 she is Associate Professor at the Division of Molecular Biology, Biocenter, Innsbruck Medical University, Austria, and head of the Applied Mycology group. Her main research area covers antimicrobial peptides and the characterization of their mechanistic function at the molecular level. Her special interest lies in studying the function of small, cationic and cysteine-rich antifungal proteins in cell signalling, induction of oxidative stress and apoptosis and regulation of fungal differentiation.

References

  1. Adams T.H., Yu J.H. Coordinate control of secondary metabolite production and asexual sporulation in Aspergillus nidulans. Curr. Opin. Microbiol. 1998;1:674–677. doi: 10.1016/s1369-5274(98)80114-8. [DOI] [PubMed] [Google Scholar]
  2. Adams T., Wieser J., Yu J. Asexual sporulation in Aspergillus nidulans. Microbiol. Mol. Biol. Rev. 1998;62:35–54. doi: 10.1128/mmbr.62.1.35-54.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Aerts A.M., Francois I.E., Meert E.M., Li Q.T., Cammue B.P., Thevissen K. The antifungal activity of RsAFP2, a plant defensin from raphanus sativus, involves the induction of reactive oxygen species in Candida albicans. J. Mol. Microbiol. Biotechnol. 2007;13:243–247. doi: 10.1159/000104753. [DOI] [PubMed] [Google Scholar]
  4. Aerts A.M., Francois I.E., Cammue B.P., Thevissen K. The mode of antifungal action of plant, insect and human defensins. Cell. Mol. Life Sci. 2008;65:2069–2079. doi: 10.1007/s00018-008-8035-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Aerts A.M., Carmona-Gutierrez D., Lefevre S., Govaert G., Francois I.E., Madeo F., Santos R., Cammue B.P., Thevissen K. The antifungal plant defensin RsAFP2 from radish induces apoptosis in a metacaspase independent way in Candida albicans. FEBS Lett. 2009;583:2513–2516. doi: 10.1016/j.febslet.2009.07.004. [DOI] [PubMed] [Google Scholar]
  6. Aerts A.M., Bammens L., Govaert G., Carmona-Gutierrez D., Madeo F., Cammue B.P., Thevissen K. The antifungal plant defensin HsAFP1 from Heuchera sanguinea induces apoptosis in Candida albicans. Front. Microbiol. 2011;2:47. doi: 10.3389/fmicb.2011.00047. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Ali R.S., Falconer A., Ikram M., Bissett C.E., Cerio R., Quinn A.G. Expression of the peptide antibiotics human beta defensin-1 and human beta defensin-2 in normal human skin. J. Invest. Dermatol. 2001;117:106–111. doi: 10.1046/j.0022-202x.2001.01401.x. [DOI] [PubMed] [Google Scholar]
  8. Allen A., Snyder A.K., Preuss M., Nielsen E.E., Shah D.M., Smith T.J. Plant defensins and virally encoded fungal toxin KP4 inhibit plant root growth. Planta. 2008;227:331–339. doi: 10.1007/s00425-007-0620-1. [DOI] [PubMed] [Google Scholar]
  9. Almeida M.S., Cabral K.M., Zingali R.B., Kurtenbach E. Characterization of two novel defense peptides from pea (Pisum sativum) seeds. Arch. Biochem. Biophys. 2000;378:278–286. doi: 10.1006/abbi.2000.1824. [DOI] [PubMed] [Google Scholar]
  10. Amien S., Kliwer I., Marton M.L., Debener T., Geiger D., Becker D., Dresselhaus T. Defensin-like ZmES4 mediates pollen tube burst in maize via opening of the potassium channel KZM1. PLoS Biol. 2010;8:e1000388. doi: 10.1371/journal.pbio.1000388. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Amlie-Lefond C., Paz D.A., Connelly M.P., Huffnagle G.B., Whelan N.T., Whelan H.T. Innate immunity for biodefense: a strategy whose time has come. J. Allergy Clin. Immunol. 2005;116:1334–1342. doi: 10.1016/j.jaci.2005.08.048. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Baroni A., Donnarumma G., Paoletti I., Longanesi-Cattani I., Bifulco K., Tufano M.A., Carriero M.V. Antimicrobial human beta-defensin-2 stimulates migration, proliferation and tube formation of human umbilical vein endothelial cells. Peptides. 2009;30:267–272. doi: 10.1016/j.peptides.2008.11.001. [DOI] [PubMed] [Google Scholar]
  13. Batta G., Barna T., Gaspari Z., Sandor S., Kover K.E., Binder U., Sarg B., Kaiserer L., Chhillar A.K., Eigentler A., Leiter E., Hegedus N., Pocsi I., Lindner H., Marx F. Functional aspects of the solution structure and dynamics of PAF – a highly-stable antifungal protein from Penicillium chrysogenum. FEBS J. 2009;276:2875–2890. doi: 10.1111/j.1742-4658.2009.07011.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Befus A.D., Mowat C., Gilchrist M., Hu J., Solomon S., Bateman A. Neutrophil defensins induce histamine secretion from mast cells: mechanisms of action. J. Immunol. 1999;163:947–953. [PubMed] [Google Scholar]
  15. 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. doi: 10.1111/j.1365-2958.2009.06936.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. 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. doi: 10.1128/EC.00050-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. 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. doi: 10.1186/1471-2180-11-209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Biragyn A., Ruffini P.A., Leifer C.A., Klyushnenkova E., Shakhov A., Chertov O., Shirakawa A.K., Farber J.M., Segal D.M., Oppenheim J.J., Kwak L.W. Toll-like receptor 4-dependent activation of dendritic cells by beta-defensin 2. Science. 2002;298:1025–1029. doi: 10.1126/science.1075565. [DOI] [PubMed] [Google Scholar]
  19. Boman H.G. Antibacterial peptides: basic facts and emerging concepts. J. Intern. Med. 2003;254:197–215. doi: 10.1046/j.1365-2796.2003.01228.x. [DOI] [PubMed] [Google Scholar]
  20. Brogden K.A. Antimicrobial peptides: pore formers or metabolic inhibitors in bacteria? Nat. Rev. Microbiol. 2005;3:238–250. doi: 10.1038/nrmicro1098. [DOI] [PubMed] [Google Scholar]
  21. Bulet P., Stocklin R. Insect antimicrobial peptides: structures, properties and gene regulation. Protein Pept. Lett. 2005;12:3–11. doi: 10.2174/0929866053406011. [DOI] [PubMed] [Google Scholar]
  22. Bulet P., Hetru C., Dimarcq J.L., Hoffmann D. Antimicrobial peptides in insects; structure and function. Dev. Comp. Immunol. 1999;23:329–344. doi: 10.1016/s0145-305x(99)00015-4. [DOI] [PubMed] [Google Scholar]
  23. Bulmer M.S., Bachelet I., Raman R., Rosengaus R.B., Sasisekharan R. Targeting an antimicrobial effector function in insect immunity as a pest control strategy. Proc. Natl. Acad. Sci. U.S.A. 2009;106:12652–12657. doi: 10.1073/pnas.0904063106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Caaveiro J.M., Molina A., Gonzalez-Manas J.M., Rodriguez-Palenzuela P., Garcia-Olmedo F., Goni F.M. Differential effects of five types of antipathogenic plant peptides on model membranes. FEBS Lett. 1997;410:338–342. doi: 10.1016/s0014-5793(97)00613-3. [DOI] [PubMed] [Google Scholar]
  25. 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. doi: 10.1021/bi00009a032. [DOI] [PubMed] [Google Scholar]
  26. Carvalho Ade O., Gomes V.M. Plant defensins – prospects for the biological functions and biotechnological properties. Peptides. 2009;30:1007–1020. doi: 10.1016/j.peptides.2009.01.018. [DOI] [PubMed] [Google Scholar]
  27. Chen K.C., Lin C.Y., Kuan C.C., Sung H.Y., Chen C.S. A novel defensin encoded by a mungbean cDNA exhibits insecticidal activity against bruchid. J. Agric. Food Chem. 2002;50:7258–7263. doi: 10.1021/jf020527q. [DOI] [PubMed] [Google Scholar]
  28. Chen J.J., Chen G.H., Hsu H.C., Li S.S., Chen C.S. Cloning and functional expression of a mungbean defensin VrD1 in Pichia pastoris. J. Agric. Food Chem. 2004;52:2256–2261. doi: 10.1021/jf030662i. [DOI] [PubMed] [Google Scholar]
  29. Chen X., Niyonsaba F., Ushio H., Hara M., Yokoi H., Matsumoto K., Saito H., Nagaoka I., Ikeda S., Okumura K., Ogawa H. Antimicrobial peptides human beta-defensin (hBD)-3 and hBD-4 activate mast cells and increase skin vascular permeability. Eur. J. Immunol. 2007;37:434–444. doi: 10.1002/eji.200636379. [DOI] [PubMed] [Google Scholar]
  30. Chiang C., Hadwiger L. The Fusarium solani-induced expression of a pea gene family encoding high cysteine content proteins. Mol. Plant Microbe Interact. 1991;4:324–331. doi: 10.1094/mpmi-4-324. [DOI] [PubMed] [Google Scholar]
  31. Cociancich S., Ghazi A., Hetru C., Hoffmann J.A., Letellier L. Insect defensin, an inducible antibacterial peptide, forms voltage-dependent channels in Micrococcus luteus. J. Biol. Chem. 1993;268:19239–19245. [PubMed] [Google Scholar]
  32. Cohen L., Moran Y., Sharon A., Segal D., Gordon D., Gurevitz M. Drosomycin, an innate immunity peptide of Drosophila melanogaster, interacts with the fly voltage-gated sodium channel. J. Biol. Chem. 2009;284:23558–23563. doi: 10.1074/jbc.M109.023358. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. de Zelicourt A., Letousey P., Thoiron S., Campion C., Simoneau P., Elmorjani K., Marion D., Simier P., Delavault P. Ha-DEF1, a sunflower defensin, induces cell death in Orobanche parasitic plants. Planta. 2007;226:591–600. doi: 10.1007/s00425-007-0507-1. [DOI] [PubMed] [Google Scholar]
  34. Dimarcq J.L., Bulet P., Hetru C., Hoffmann J. Cysteine-rich antimicrobial peptides in invertebrates. Biopolymers. 1998;47:465–477. doi: 10.1002/(SICI)1097-0282(1998)47:6<465::AID-BIP5>3.0.CO;2-#. [DOI] [PubMed] [Google Scholar]
  35. Eigentler A., Pocsi I., Marx F. The anisin1 gene encodes a defensin-like protein and supports the fitness of Aspergillus nidulans. Arch. Microbiol. 2012;194:427–437. doi: 10.1007/s00203-011-0773-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Fehlbaum P., Bulet P., Michaut L., Lagueux M., Broekaert W.F., Hetru C., Hoffmann J.A. Insect immunity. Septic injury of Drosophila induces the synthesis of a potent antifungal peptide with sequence homology to plant antifungal peptides. J. Biol. Chem. 1994;269:33159–33163. [PubMed] [Google Scholar]
  37. Fellermann K., Stange D.E., Schaeffeler E., Schmalzl H., Wehkamp J., Bevins C.L., Reinisch W., Teml A., Schwab M., Lichter P., Radlwimmer B., Stange E.F. A chromosome 8 gene-cluster polymorphism with low human beta-defensin 2 gene copy number predisposes to Crohn disease of the colon. Am. J. Hum. Genet. 2006;79:439–448. doi: 10.1086/505915. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Feng Z., Jiang B., Chandra J., Ghannoum M., Nelson S., Weinberg A. Human beta-defensins: differential activity against candidal species and regulation by Candida albicans. J. Dent. Res. 2005;84:445–450. doi: 10.1177/154405910508400509. [DOI] [PubMed] [Google Scholar]
  39. Ferket K.K.A., Levery S.B., Park C., Cammue B.P.A., Thevissen K. Isolation and characterization of Neurospora crassa mutants resistant to antifungal plant defensins. Fungal Genet. Biol. 2003;40:176–185. doi: 10.1016/s1087-1845(03)00085-9. [DOI] [PubMed] [Google Scholar]
  40. Ferrandon D., Jung A.C., Criqui M., Lemaitre B., Uttenweiler-Joseph S., Michaut L., Reichhart J., Hoffmann J.A. A drosomycin-GFP reporter transgene reveals a local immune response in Drosophila that is not dependent on the Toll pathway. EMBO J. 1998;17:1217–1227. doi: 10.1093/emboj/17.5.1217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Francois I.E., Aerts A.M., Cammue B.P., Thevissen K. Currently used antimycotics: spectrum, mode of action and resistance occurrence. Curr. Drug Targets. 2005;6:895–907. doi: 10.2174/138945005774912744. [DOI] [PubMed] [Google Scholar]
  42. Galgoczy L., Papp T., Leiter E., Marx F., Pocsi I., Vagvolgyi C. Sensitivity of different zygomycetes to the Penicillium chrysogenum antifungal protein (PAF) J. Basic Microbiol. 2005;45:136–141. doi: 10.1002/jobm.200410512. [DOI] [PubMed] [Google Scholar]
  43. Ganz T., Selsted M.E., Szklarek D., Harwig S.S., Daher K., Bainton D.F., Lehrer R.I. Defensins. Natural peptide antibiotics of human neutrophils. J. Clin. Invest. 1985;76:1427–1435. doi: 10.1172/JCI112120. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Gao B., Zhu S.Y. Differential potency of drosomycin to Neurospora crassa and its mutant: implications for evolutionary relationship between defensins from insects and plants. Insect Mol. Biol. 2008;17:405–411. doi: 10.1111/j.1365-2583.2008.00810.x. [DOI] [PubMed] [Google Scholar]
  45. Gao A.G., Hakimi S.M., Mittanck C.A., Wu Y., Woerner B.M., Stark D.M., Shah D.M., Liang J., Rommens C.M. Fungal pathogen protection in potato by expression of a plant defensin peptide. Nat. Biotechnol. 2000;18:1307–1310. doi: 10.1038/82436. [DOI] [PubMed] [Google Scholar]
  46. Geisen R. P. nalgiovense carries a gene which is homologous to the paf gene of P. chrysogenum which codes for an antifungal peptide. Int. J. Food Microbiol. 2000;62:95–101. doi: 10.1016/s0168-1605(00)00367-6. [DOI] [PubMed] [Google Scholar]
  47. Giangaspero A., Sandri L., Tossi A. Amphipathic alpha helical antimicrobial peptides. Eur. J. Biochem. 2001;268:5589–5600. doi: 10.1046/j.1432-1033.2001.02494.x. [DOI] [PubMed] [Google Scholar]
  48. Guani-Guerra E., Santos-Mendoza T., Lugo-Reyes S.O., Teran L.M. Antimicrobial peptides: general overview and clinical implications in human health and disease. Clin. Immunol. 2010;135:1–11. doi: 10.1016/j.clim.2009.12.004. [DOI] [PubMed] [Google Scholar]
  49. Gun Lee D., Shin S., Maeng C., Jin Z., Kim K., Hahm K. Isolation and characterization of a novel antifungal peptide from Aspergillus niger. Biochem. Biophys. Res. Commun. 1999;263:646–651. doi: 10.1006/bbrc.1999.1428. [DOI] [PubMed] [Google Scholar]
  50. Hagen S., Marx F., Ram A.F., Meyer V. The antifungal protein AFP from Aspergillus giganteus inhibits chitin synthesis in sensitive fungi. Appl. Environ. Microbiol. 2007;73:2128–2134. doi: 10.1128/AEM.02497-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Hajji M., Jellouli K., Hmidet N., Balti R., Sellami-Kamoun A., Nasri M. A highly thermostable antimicrobial peptide from Aspergillus clavatus ES1: biochemical and molecular characterization. J. Ind. Microbiol. Biotechnol. 2010;37:805–813. doi: 10.1007/s10295-010-0725-6. [DOI] [PubMed] [Google Scholar]
  52. Hamilton C., Bulmer M.S. Molecular antifungal defenses in subterranean termites: RNA interference reveals in vivo roles of termicins and GNBPs against a naturally encountered pathogen. Dev. Comp. Immunol. 2012;36:372–377. doi: 10.1016/j.dci.2011.07.008. [DOI] [PubMed] [Google Scholar]
  53. Hamilton C., Lay F., Bulmer M.S. Subterranean termite prophylactic secretions and external antifungal defenses. J. Insect. Physiol. 2011;57:1259–1266. doi: 10.1016/j.jinsphys.2011.05.016. [DOI] [PubMed] [Google Scholar]
  54. Harder J., Bartels J., Christophers E., Schroder J.M. Isolation and characterization of human beta-defensin-3, a novel human inducible peptide antibiotic. J. Biol. Chem. 2001;276:5707–5713. doi: 10.1074/jbc.M008557200. [DOI] [PubMed] [Google Scholar]
  55. Hazrati E., Galen B., Lu W., Wang W., Ouyang Y., Keller M.J., Lehrer R.I., Herold B.C. Human alpha- and beta-defensins block multiple steps in herpes simplex virus infection. J. Immunol. 2006;177:8658–8666. doi: 10.4049/jimmunol.177.12.8658. [DOI] [PubMed] [Google Scholar]
  56. Hegedus N., Sigl C., Zadra I., Pocsi I., Marx F. The paf gene product modulates asexual development in Penicillium chrysogenum. J. Basic Microbiol. 2011;51:253–262. doi: 10.1002/jobm.201000321. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Hegedus N., Leiter E., Kovacs B., Tomori V., Kwon N.J., Emri T., Marx F., Batta G., Csernoch L., Haas H., Yu J.H., Pocsi I. The small molecular mass antifungal protein of Penicillium chrysogenum – a mechanism of action oriented review. J. Basic Microbiol. 2011;51:561–571. doi: 10.1002/jobm.201100041. [DOI] [PubMed] [Google Scholar]
  58. Higashiyama T. Peptide signaling in pollen–pistil interactions. Plant Cell Physiol. 2010;51:177–189. doi: 10.1093/pcp/pcq008. [DOI] [PubMed] [Google Scholar]
  59. Hoff B., Kamerewerd J., Sigl C., Mitterbauer R., Zadra I., Kürnsteiner H., Kück U. Two components of a velvet-like complex control hyphal morphogenesis, conidiophore development, and penicillin biosynthesis in Penicillium chrysogenum. Eukaryot. Cell. 2010;9:1236–1250. doi: 10.1128/EC.00077-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Hoover D.M., Rajashankar K.R., Blumenthal R., Puri A., Oppenheim J.J., Chertov O., Lubkowski J. The structure of human beta-defensin-2 shows evidence of higher order oligomerization. J. Biol. Chem. 2000;275:32911–32918. doi: 10.1074/jbc.M006098200. [DOI] [PubMed] [Google Scholar]
  61. Hoover D.M., Chertov O., Lubkowski J. The structure of human beta-defensin-1: new insights into structural properties of beta-defensins. J. Biol. Chem. 2001;276:39021–39026. doi: 10.1074/jbc.M103830200. [DOI] [PubMed] [Google Scholar]
  62. Hoover D.M., Wu Z., Tucker K., Lu W., Lubkowski J. Antimicrobial characterization of human beta-defensin 3 derivatives. Antimicrob. Agents Chemother. 2003;47:2804–2809. doi: 10.1128/AAC.47.9.2804-2809.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Jiggins F.M., Kim K.W. The evolution of antifungal peptides in Drosophila. Genetics. 2005;171:1847–1859. doi: 10.1534/genetics.105.045435. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Kaiserer L., Oberparleiter C., Weiler-Gorz R., Burgstaller W., Leiter E., Marx F. Characterization of the Penicillium chrysogenum antifungal protein PAF. Arch. Microbiol. 2003;180:204–210. doi: 10.1007/s00203-003-0578-8. [DOI] [PubMed] [Google Scholar]
  65. Koike M., Okamoto T., Tsuda S., Imai R. A novel plant defensin-like gene of winter wheat is specifically induced during cold acclimation. Biochem. Biophys. Res. Commun. 2002;298:46–53. doi: 10.1016/s0006-291x(02)02391-4. [DOI] [PubMed] [Google Scholar]
  66. Kovacs L., Viragh M., Tako M., Papp T., Vagvolgyi C., Galgoczy L. Isolation and characterization of Neosartorya fischeri antifungal protein (NFAP) Peptides. 2011;32:1724–1731. doi: 10.1016/j.peptides.2011.06.022. [DOI] [PubMed] [Google Scholar]
  67. Lai Y., Gallo R.L. AMPed up immunity: how antimicrobial peptides have multiple roles in immune defense. Trends Immunol. 2009;30:131–141. doi: 10.1016/j.it.2008.12.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Lamberty M., Ades S., Uttenweiler-Joseph S., Brookhart G., Bushey D., Hoffmann J.A., Bulet P. Insect immunity. Isolation from the lepidopteran Heliothis virescens of a novel insect defensin with potent antifungal activity. J. Biol. Chem. 1999;274:9320–9326. doi: 10.1074/jbc.274.14.9320. [DOI] [PubMed] [Google Scholar]
  69. Lamberty M., Zachary D., Lanot R., Bordereau C., Robert A., Hoffmann J.A., Bulet P. Insect immunity. Constitutive expression of a cysteine-rich antifungal and a linear antibacterial peptide in a termite insect. J. Biol. Chem. 2001;276:4085–4092. doi: 10.1074/jbc.M002998200. [DOI] [PubMed] [Google Scholar]
  70. Landon C., Sodano P., Hetru C., Hoffmann J., Ptak M. Solution structure of drosomycin, the first inducible antifungal protein from insects. Protein Sci. 1997;6:1878–1884. doi: 10.1002/pro.5560060908. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Lay F.T., Anderson M.A. Defensins – components of the innate immune system in plants. Curr. Protein Pept. Sci. 2005;6:85–101. doi: 10.2174/1389203053027575. [DOI] [PubMed] [Google Scholar]
  72. Lay F.T., Brugliera F., Anderson M.A. Isolation and properties of floral defensins from ornamental tobacco and petunia. Plant Physiol. 2003;131:1283–1293. doi: 10.1104/pp.102.016626. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Lay F.T., Schirra H.J., Scanlon M.J., Anderson M.A., Craik D.J. The three-dimensional solution structure of NaD1, a new floral defensin from Nicotiana alata and its application to a homology model of the crop defense protein alfAFP. J. Mol. Biol. 2003;325:175–188. doi: 10.1016/s0022-2836(02)01103-8. [DOI] [PubMed] [Google Scholar]
  74. Lee Y.S., Yun E.K., Jang W.S., Kim I., Lee J.H., Park S.Y., Ryu K.S., Seo S.J., Kim C.H., Lee I.H. Purification, cDNA cloning and expression of an insect defensin from the great wax moth, Galleria mellonella. Insect Mol. Biol. 2004;13:65–72. doi: 10.1111/j.1365-2583.2004.00462.x. [DOI] [PubMed] [Google Scholar]
  75. Lehrer R.I., Ganz T. Endogenous vertebrate antibiotics. Defensins, protegrins, and other cysteine-rich antimicrobial peptides. Ann. N.Y.Acad. Sci. 1996;797:228–239. doi: 10.1111/j.1749-6632.1996.tb52963.x. [DOI] [PubMed] [Google Scholar]
  76. Lehrer R.I., Ganz T. Antimicrobial peptides in mammalian and insect host defence. Curr. Opin. Immunol. 1999;11:23–27. doi: 10.1016/s0952-7915(99)80005-3. [DOI] [PubMed] [Google Scholar]
  77. Lehrer R.I. Multispecific myeloid defensins. Curr. Opin. Hematol. 2007;14:16–21. doi: 10.1097/00062752-200701000-00005. [DOI] [PubMed] [Google Scholar]
  78. Leiter E., Szappanos H., Oberparleiter C., Kaiserer L., Csernoch L., Pusztahelyi T., Emri T., Pocsi I., Salvenmoser W., Marx F. Antifungal protein PAF severely affects the integrity of the plasma membrane of Aspergillus nidulans and induces an apoptosis-like phenotype. Antimicrob. Agents Chemother. 2005;49:2445–2453. doi: 10.1128/AAC.49.6.2445-2453.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Liu A.Y., Destoumieux D., Wong A.V., Park C.H., Valore E.V., Liu L., Ganz T. Human beta-defensin-2 production in keratinocytes is regulated by interleukin-1, bacteria, and the state of differentiation. J. Invest. Dermatol. 2002;118:275–281. doi: 10.1046/j.0022-202x.2001.01651.x. [DOI] [PubMed] [Google Scholar]
  80. Liu Y.J., Cheng C.S., Lai S.M., Hsu M.P., Chen C.S., Lyu P.C. Solution structure of the plant defensin VrD1 from mung bean and its possible role in insecticidal activity against bruchids. Proteins. 2006;63:777–786. doi: 10.1002/prot.20962. [DOI] [PubMed] [Google Scholar]
  81. Lobo D.S., Pereira I.B., Fragel-Madeira L., Medeiros L.N., Cabral L.M., Faria J., Bellio M., Campos R.C., Linden R., Kurtenbach E. Antifungal Pisum sativum defensin 1 interacts with Neurospora crassa cyclin F related to the cell cycle. Biochemistry. 2007;46:987–996. doi: 10.1021/bi061441j. [DOI] [PubMed] [Google Scholar]
  82. Lopez-Garcia B., Ubhayasekera W., Gallo R.L., Marcos J.F. Parallel evaluation of antimicrobial peptides derived from the synthetic PAF26 and the human LL37. Biochem. Biophys. Res. Commun. 2007;356:107–113. doi: 10.1016/j.bbrc.2007.02.093. [DOI] [PubMed] [Google Scholar]
  83. Mak P., Zdybicka-Barabas A., Cytrynska M. A different repertoire of Galleria mellonella antimicrobial peptides in larvae challenged with bacteria and fungi. Dev. Comp. Immunol. 2010;34:1129–1136. doi: 10.1016/j.dci.2010.06.005. [DOI] [PubMed] [Google Scholar]
  84. Mallow E.B., Harris A., Salzman N., Russell J.P., DeBerardinis R.J., Ruchelli E., Bevins C.L. Human enteric defensins. Gene structure and developmental expression. J. Biol. Chem. 1996;271:4038–4045. doi: 10.1074/jbc.271.8.4038. [DOI] [PubMed] [Google Scholar]
  85. Maroti G., Kereszt A., Kondorosi E., Mergaert P. Natural roles of antimicrobial peptides in microbes, plants and animals. Res. Microbiol. 2011;162:363–374. doi: 10.1016/j.resmic.2011.02.005. [DOI] [PubMed] [Google Scholar]
  86. Martin-Urdiroz M., Martinez-Rocha A.L., Di Pietro A., Martinez-del-Pozo A., Roncero M.I. Differential toxicity of antifungal protein AFP against mutants of Fusarium oxysporum. Int. Microbiol. 2009;12:115–121. [PubMed] [Google Scholar]
  87. Marx F., Haas H., Reindl M., Stöffler G., Lottspeich F., Redl B. Cloning, structural organization and regulation of expression of the Penicillium chrysogenum paf gene encoding an abundantly secreted protein with antifungal activity. Gene. 1995;167:167–171. doi: 10.1016/0378-1119(95)00701-6. [DOI] [PubMed] [Google Scholar]
  88. 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. doi: 10.1007/s00018-007-7364-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. 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. doi: 10.1007/s00253-004-1600-z. [DOI] [PubMed] [Google Scholar]
  90. Mendez-Samperio P., Alba L., Trejo A. Mycobacterium bovis-mediated induction of human beta-defensin-2 in epithelial cells is controlled by intracellular calcium and p38MAPK. J. Infect. 2007;54:469–474. doi: 10.1016/j.jinf.2006.08.009. [DOI] [PubMed] [Google Scholar]
  91. Meyer V., Stahl U. New insights in the regulation of the afp gene encoding the antifungal protein of Aspergillus giganteus. Curr. Genet. 2002;42:36–42. doi: 10.1007/s00294-002-0324-9. [DOI] [PubMed] [Google Scholar]
  92. Meyer V., Wedde M., Stahl U. Transcriptional regulation of the antifungal protein in Aspergillus giganteus. Mol. Genet. Genomics. 2002;266:747–757. doi: 10.1007/s00438-001-0609-6. [DOI] [PubMed] [Google Scholar]
  93. 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. doi: 10.1007/s00253-007-1291-3. [DOI] [PubMed] [Google Scholar]
  94. Mirouze M., Sels J., Richard O., Czernic P., Loubet S., Jacquier A., Francois I.E., Cammue B.P., Lebrun M., Berthomieu P., Marques L. A putative novel role for plant defensins: a defensin from the zinc hyper-accumulating plant, Arabidopsis halleri, confers zinc tolerance. Plant J. 2006;47:329–342. doi: 10.1111/j.1365-313X.2006.02788.x. [DOI] [PubMed] [Google Scholar]
  95. Moreno A.B., Martinez 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. doi: 10.1007/s00253-006-0362-1. [DOI] [PubMed] [Google Scholar]
  96. Munoz A., Lopez-Garcia B., Marcos J.F. Studies on the mode of action of the antifungal hexapeptide PAF26. Antimicrob. Agents Chemother. 2006;50:3847–3855. doi: 10.1128/AAC.00650-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Munoz A., Lopez-Garcia B., Perez-Paya E., Marcos J.F. Antimicrobial properties of derivatives of the cationic tryptophan-rich hexapeptide PAF26. Biochem. Biophys. Res. Commun. 2007;354:172–177. doi: 10.1016/j.bbrc.2006.12.173. [DOI] [PubMed] [Google Scholar]
  98. Munoz A., Marcos J.F., Read N.D. Concentration-dependent mechanisms of cell penetration and killing by the de novo designed antifungal hexapeptide PAF26. Mol. Microbiol. 2012;85:89–106. doi: 10.1111/j.1365-2958.2012.08091.x. [DOI] [PubMed] [Google Scholar]
  99. Murakami M., Ohtake T., Dorschner R.A., Schittek B., Garbe C., Gallo R.L. Cathelicidin anti-microbial peptide expression in sweat, an innate defense system for the skin. J. Invest. Dermatol. 2002;119:1090–1095. doi: 10.1046/j.1523-1747.2002.19507.x. [DOI] [PubMed] [Google Scholar]
  100. Mygind P.H., Fischer R.L., Schnorr K.M., Hansen M.T., Sonksen C.P., Ludvigsen S., Raventos D., Buskov S., Christensen B., De Maria L., Taboureau O., Yaver D., Elvig-Jorgensen S.G., Sorensen M.V., Christensen B.E., Kjaerulff S., Frimodt-Moller N., Lehrer R.I., Zasloff M., Kristensen H.H. Plectasin is a peptide antibiotic with therapeutic potential from a saprophytic fungus. Nature. 2005;437:975–980. doi: 10.1038/nature04051. [DOI] [PubMed] [Google Scholar]
  101. Niyonsaba F., Someya A., Hirata M., Ogawa H., Nagaoka I. Evaluation of the effects of peptide antibiotics human beta-defensins-1/-2 and LL-37 on histamine release and prostaglandin D(2) production from mast cells. Eur. J. Immunol. 2001;31:1066–1075. doi: 10.1002/1521-4141(200104)31:4<1066::aid-immu1066>3.0.co;2-#. [DOI] [PubMed] [Google Scholar]
  102. Niyonsaba F., Nagaoka I., Ogawa H. Human defensins and cathelicidins in the skin: beyond direct antimicrobial properties. Crit. Rev. Immunol. 2006;26:545–576. doi: 10.1615/critrevimmunol.v26.i6.60. [DOI] [PubMed] [Google Scholar]
  103. Niyonsaba F., Ushio H., Nakano N., Ng W., Sayama K., Hashimoto K., Nagaoka I., Okumura K., Ogawa H. Antimicrobial peptides human beta-defensins stimulate epidermal keratinocyte migration, proliferation and production of proinflammatory cytokines and chemokines. J. Invest. Dermatol. 2007;127:594–604. doi: 10.1038/sj.jid.5700599. [DOI] [PubMed] [Google Scholar]
  104. Oberparleiter C., Kaiserer L., Haas H., Ladurner P., Andratsch M., Marx F. Active internalization of the Penicillium chrysogenum antifungal protein PAF in sensitive aspergilli. Antimicrob. Agents Chemother. 2003;47:3598–3601. doi: 10.1128/AAC.47.11.3598-3601.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Okuda S., Tsutsui H., Shiina K., Sprunck S., Takeuchi H., Yui R., Kasahara R.D., Hamamura Y., Mizukami A., Susaki D., Kawano N., Sakakibara T., Namiki S., Itoh K., Otsuka K., Matsuzaki M., Nozaki H., Kuroiwa T., Nakano A., Kanaoka M.M., Dresselhaus T., Sasaki N., Higashiyama T. Defensin-like polypeptide LUREs are pollen tube attractants secreted from synergid cells. Nature. 2009;458:357–361. doi: 10.1038/nature07882. [DOI] [PubMed] [Google Scholar]
  106. Oomen R.J., Seveno-Carpentier E., Ricodeau N., Bournaud C., Conejero G., Paris N., Berthomieu P., Marques L. Plant defensin AhPDF1.1 is not secreted in leaves but it accumulates in intracellular compartments. New Phytol. 2011;192:140–150. doi: 10.1111/j.1469-8137.2011.03792.x. [DOI] [PubMed] [Google Scholar]
  107. Osborn R., De Samblanx G., Thevissen K., Goderis I., Torrekens S., Van Leuven F., Attenborough S., Rees S., Broekaert W. Isolation and characterisation of plant defensins from seeds of Asteraceae, Fabaceae, Hippocastanaceae and Saxifragaceae. FEBS Lett. 1995;368:257–262. doi: 10.1016/0014-5793(95)00666-w. [DOI] [PubMed] [Google Scholar]
  108. Otvos L., Jr. Antibacterial peptides isolated from insects. J. Pept. Sci. 2000;6:497–511. doi: 10.1002/1099-1387(200010)6:10<497::AID-PSC277>3.0.CO;2-W. [DOI] [PubMed] [Google Scholar]
  109. Ouedraogo J.P., Hagen S., Spielvogel A., Engelhardt S., Meyer V. Survival strategies of yeast and filamentous fungi against the antifungal protein AFP. J. Biol. Chem. 2011;286:13859–13868. doi: 10.1074/jbc.M110.203588. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Papagianni M. Ribosomally synthesized peptides with antimicrobial properties: biosynthesis, structure, function, and applications. Biotechnol. Adv. 2003;21:465–499. doi: 10.1016/s0734-9750(03)00077-6. [DOI] [PubMed] [Google Scholar]
  111. Pazgier M., Hoover D.M., Yang D., Lu W., Lubkowski J. Human beta-defensins. Cell. Mol. Life Sci. 2006;63:1294–1313. doi: 10.1007/s00018-005-5540-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Penninckx I., Eggermont K., Terras F., Thomma B., De Samblanx G., Buchala A., Métraux J., Manners J., Broekaert W. Pathogen-induced systemic activation of a plant defensin gene in Arabidopsis follows a salicylic acid-independent pathway. Plant Cell. 1996;8:2309–2323. doi: 10.1105/tpc.8.12.2309. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Raina S., De Vizio D., Odell M., Clements M., Vanhulle S., Keshavarz T. Microbial quorum sensing: a tool or a target for antimicrobial therapy? Biotechnol. Appl. Biochem. 2009;54:65–84. doi: 10.1042/BA20090072. [DOI] [PubMed] [Google Scholar]
  114. Ramamoorthy V., Cahoon E.B., Li J., Thokala M., Minto R.E., Shah D.M. Glucosylceramide synthase is essential for alfalfa defensin-mediated growth inhibition but not for pathogenicity of Fusarium graminearum. Mol. Microbiol. 2007;66:771–786. doi: 10.1111/j.1365-2958.2007.05955.x. [DOI] [PubMed] [Google Scholar]
  115. Ramamoorthy V., Zhao X., Snyder A.K., Xu J.R., Shah D.M. Two mitogen-activated protein kinase signalling cascades mediate basal resistance to antifungal plant defensins in Fusarium graminearum. Cell. Microbiol. 2007;9:1491–1506. doi: 10.1111/j.1462-5822.2006.00887.x. [DOI] [PubMed] [Google Scholar]
  116. Reddy K.V., Yedery R.D., Aranha C. Antimicrobial peptides: premises and promises. Int. J. Antimicrob. Agents. 2004;24:536–547. doi: 10.1016/j.ijantimicag.2004.09.005. [DOI] [PubMed] [Google Scholar]
  117. Rodríguez-Martín A., Acosta R., Liddell S., Núñez F., Benito M., Asensio M. Characterization of the novel antifungal protein PgAFP and the encoding gene of Penicillium chrysogenum. Peptides. 2010;31:541–547. doi: 10.1016/j.peptides.2009.11.002. [DOI] [PubMed] [Google Scholar]
  118. Rosenfeld Y., Lev N., Shai Y. Effect of the hydrophobicity to net positive charge ratio on antibacterial and anti-endotoxin activities of structurally similar antimicrobial peptides. Biochemistry. 2010;49:853–861. doi: 10.1021/bi900724x. [DOI] [PubMed] [Google Scholar]
  119. Sagaram U.S., Pandurangi R., Kaur J., Smith T.J., Shah D.M. Structure-activity determinants in antifungal plant defensins MsDef1 and MtDef4 with different modes of action against Fusarium graminearum. PLoS One. 2011;6:e18550. doi: 10.1371/journal.pone.0018550. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Salzet M. Antimicrobial peptides are signaling molecules. Trends Immunol. 2002;23:283–284. doi: 10.1016/s1471-4906(02)02236-6. [DOI] [PubMed] [Google Scholar]
  121. Schauber J., Dorschner R.A., Yamasaki K., Brouha B., Gallo R.L. Control of the innate epithelial antimicrobial response is cell-type specific and dependent on relevant microenvironmental stimuli. Immunology. 2006;118:509–519. doi: 10.1111/j.1365-2567.2006.02399.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Schneider J.J., Unholzer A., Schaller M., Schafer-Korting M., Korting H.C. Human defensins. J. Mol. Med. (Berl) 2005;83:587–595. doi: 10.1007/s00109-005-0657-1. [DOI] [PubMed] [Google Scholar]
  123. Schopfer C.R., Nasrallah M.E., Nasrallah J.B. The male determinant of self-incompatibility in Brassica. Science. 1999;286:1697–1700. doi: 10.1126/science.286.5445.1697. [DOI] [PubMed] [Google Scholar]
  124. Schuhmann B., Seitz V., Vilcinskas A., Podsiadlowski L. Cloning and expression of gallerimycin, an antifungal peptide expressed in immune response of greater wax moth larvae, Galleria mellonella. Arch. Insect Biochem. Physiol. 2003;53:125–133. doi: 10.1002/arch.10091. [DOI] [PubMed] [Google Scholar]
  125. Schutte B.C., Mitros J.P., Bartlett J.A., Walters J.D., Jia H.P., Welsh M.J., Casavant T.L., McCray P.B., Jr. Discovery of five conserved beta-defensin gene clusters using a computational search strategy. Proc. Natl. Acad. Sci. U.S.A. 2002;99:2129–2133. doi: 10.1073/pnas.042692699. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Segura A., Moreno M., Molina A., García-Olmedo F. Novel defensin subfamily from spinach (Spinacia oleracea) FEBS Lett. 1998;435:159–162. doi: 10.1016/s0014-5793(98)01060-6. [DOI] [PubMed] [Google Scholar]
  127. Seibold M., Wolschann P., Bodevin S., Olsen O. Properties of the bubble protein, a defensin and an abundant component of a fungal exudate. Peptides. 2011;32:1989–1995. doi: 10.1016/j.peptides.2011.08.022. [DOI] [PubMed] [Google Scholar]
  128. Selsted M.E., Ouellette A.J. Mammalian defensins in the antimicrobial immune response. Nat. Immunol. 2005;6:551–557. doi: 10.1038/ni1206. [DOI] [PubMed] [Google Scholar]
  129. Skouri-Gargouri H., Gargouri A. First isolation of a novel thermostable antifungal peptide secreted by Aspergillus clavatus. Peptides. 2008;29:1871–1877. doi: 10.1016/j.peptides.2008.07.005. [DOI] [PubMed] [Google Scholar]
  130. Spelbrink R.G., Dilmac N., Allen A., Smith T.J., Shah D.M., Hockerman G.H. Differential antifungal and calcium channel-blocking activity among structurally related plant defensins. Plant Physiol. 2004;135:2055–2067. doi: 10.1104/pp.104.040873. [DOI] [PMC free article] [PubMed] [Google Scholar]
  131. Stotz H.U., Spence B., Wang Y. A defensin from tomato with dual function in defense and development. Plant Mol. Biol. 2009;71:131–143. doi: 10.1007/s11103-009-9512-z. [DOI] [PubMed] [Google Scholar]
  132. Stotz H.U., Thomson J.G., Wang Y. Plant defensins: defense, development and application. Plant Signal. Behav. 2009;4:1010–1012. doi: 10.4161/psb.4.11.9755. [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Sun L., Finnegan C.M., Kish-Catalone T., Blumenthal R., Garzino-Demo P., La Terra Maggiore G.M., Berrone S., Kleinman C., Wu Z., Abdelwahab S., Lu W., Garzino-Demo A. Human beta-defensins suppress human immunodeficiency virus infection: potential role in mucosal protection. J. Virol. 2005;79:14318–14329. doi: 10.1128/JVI.79.22.14318-14329.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. Sun C.Q., Arnold R., Fernandez-Golarz C., Parrish A.B., Almekinder T., He J., Ho S.M., Svoboda P., Pohl J., Marshall F.F., Petros J.A. Human beta-defensin-1, a potential chromosome 8p tumor suppressor: control of transcription and induction of apoptosis in renal cell carcinoma. Cancer Res. 2006;66:8542–8549. doi: 10.1158/0008-5472.CAN-06-0294. [DOI] [PubMed] [Google Scholar]
  135. Takayama S., Shimosato H., Shiba H., Funato M., Che F.S., Watanabe M., Iwano M., Isogai A. Direct ligand–receptor complex interaction controls Brassica self-incompatibility. Nature. 2001;413:534–538. doi: 10.1038/35097104. [DOI] [PubMed] [Google Scholar]
  136. Tamaoki M., Freeman J., Marquse L., Pilon-Smits E. New insights into the roles of ethylene and jasmonic acid in the acquisition of selenium resistance in plants. Plant Signal. Behav. 2008;3:865–867. doi: 10.4161/psb.3.10.6050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Taylor K., Barran P.E., Dorin J.R. Structure–activity relationships in beta-defensin peptides. Biopolymers. 2008;90:1–7. doi: 10.1002/bip.20900. [DOI] [PubMed] [Google Scholar]
  138. Terras F., Schoofs H., De Bolle M., Van Leuven F., Rees S., Vanderleyden J., Cammue B., Broekaert W. Analysis of two novel classes of plant antifungal proteins from radish (Raphanus sativus L.) seeds. J. Biol. Chem. 1992;267:15301–15309. [PubMed] [Google Scholar]
  139. Terras F., Eggermont K., Kovaleva V., Raikhel N., Osborn R., Kester A., Rees S., Torrekens S., Van Leuven F., Vanderleyden J. Small cysteine-rich antifungal proteins from radish: their role in host defense. Plant Cell. 1995;7:573–588. doi: 10.1105/tpc.7.5.573. [DOI] [PMC free article] [PubMed] [Google Scholar]
  140. Theis T., Wedde M., Meyer V., Stahl U. The antifungal protein from Aspergillus giganteus causes membrane permeabilization. Antimicrob. Agents Chemother. 2003;47:588–593. doi: 10.1128/AAC.47.2.588-593.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  141. 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. doi: 10.1016/j.resmic.2004.08.006. [DOI] [PubMed] [Google Scholar]
  142. Thevissen K., Ghazi A., De Samblanx G.W., Brownlee C., Osborn R.W., Broekaert W.F. Fungal membrane responses induced by plant defensins and thionins. J. Biol. Chem. 1996;271:15018–15025. doi: 10.1074/jbc.271.25.15018. [DOI] [PubMed] [Google Scholar]
  143. Thevissen K., Osborn R.W., Acland D.P., Broekaert W.F. Specific, high affinity binding sites for an antifungal plant defensin on Neurospora crassa hyphae and microsomal membranes. J. Biol. Chem. 1997;272:32176–32181. doi: 10.1074/jbc.272.51.32176. [DOI] [PubMed] [Google Scholar]
  144. Thevissen K., Terras F.R., Broekaert W.F. Permeabilization of fungal membranes by plant defensins inhibits fungal growth. Appl. Environ. Microbiol. 1999;65:5451–5458. doi: 10.1128/aem.65.12.5451-5458.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Thevissen K., Cammue B.P., Lemaire K., Winderickx J., Dickson R.C., Lester R.L., Ferket K.K., Van Even F., Parret A.H., Broekaert W.F. A gene encoding a sphingolipid biosynthesis enzyme determines the sensitivity of Saccharomyces cerevisiae to an antifungal plant defensin from dahlia (Dahlia merckii) Proc. Natl. Acad. Sci. U.S.A. 2000;97:9531–9536. doi: 10.1073/pnas.160077797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Thevissen K., Ferket K.K., Francois I.E., Cammue B.P. Interactions of antifungal plant defensins with fungal membrane components. Peptides. 2003;24:1705–1712. doi: 10.1016/j.peptides.2003.09.014. [DOI] [PubMed] [Google Scholar]
  147. Thevissen K., Francois I.E., Takemoto J.Y., Ferket K.K., Meert E.M., Cammue B.P. DmAMP1, an antifungal plant defensin from dahlia (Dahlia merckii), interacts with sphingolipids from Saccharomyces cerevisiae. FEMS Microbiol. Lett. 2003;226:169–173. doi: 10.1016/S0378-1097(03)00590-1. [DOI] [PubMed] [Google Scholar]
  148. Thevissen K., Warnecke D.C., Francois I.E., Leipelt M., Heinz E., Ott C., Zahringer U., Thomma B.P., Ferket K.K., Cammue B.P. Defensins from insects and plants interact with fungal glucosylceramides. J. Biol. Chem. 2004;279:3900–3905. doi: 10.1074/jbc.M311165200. [DOI] [PubMed] [Google Scholar]
  149. Thevissen K., Idkowiak-Baldys J., Im Y.J., Takemoto J., Francois I.E., Ferket K.K., Aerts A.M., Meert E.M., Winderickx J., Roosen J., Cammue B.P. SKN1, a novel plant defensin-sensitivity gene in Saccharomyces cerevisiae, is implicated in sphingolipid biosynthesis. FEBS Lett. 2005;579:1973–1977. doi: 10.1016/j.febslet.2005.02.043. [DOI] [PubMed] [Google Scholar]
  150. Thomma B., Cammue B., Thevissen K. Plant defensins. Planta. 2002;216:193–202. doi: 10.1007/s00425-002-0902-6. [DOI] [PubMed] [Google Scholar]
  151. Tian C., Gao B., Rodriguez Mdel C., Lanz-Mendoza H., Ma B., Zhu S. Gene expression, antiparasitic activity, and functional evolution of the drosomycin family. Mol. Immunol. 2008;45:3909–3916. doi: 10.1016/j.molimm.2008.06.025. [DOI] [PubMed] [Google Scholar]
  152. Tollner T.L., Yudin A.I., Treece C.A., Overstreet J.W., Cherr G.N. Macaque sperm release ESP13.2 and PSP94 during capacitation: the absence of ESP13.2 is linked to sperm-zona recognition and binding. Mol. Reprod. Dev. 2004;69:325–337. doi: 10.1002/mrd.20132. [DOI] [PubMed] [Google Scholar]
  153. Tossi A., Sandri L., Giangaspero A. Amphipathic, alpha-helical antimicrobial peptides. Biopolymers. 2000;55:4–30. doi: 10.1002/1097-0282(2000)55:1<4::AID-BIP30>3.0.CO;2-M. [DOI] [PubMed] [Google Scholar]
  154. Tzou P., Ohresser S., Ferrandon D., Capovilla M., Reichhart J.M., Lemaitre B., Hoffmann J.A., Imler J.L. Tissue-specific inducible expression of antimicrobial peptide genes in Drosophila surface epithelia. Immunity. 2000;13:737–748. doi: 10.1016/s1074-7613(00)00072-8. [DOI] [PubMed] [Google Scholar]
  155. Varoga D., Paulsen F.P., Kohrs S., Grohmann S., Lippross S., Mentlein R., Tillmann B.N., Goldring M.B., Besch L., Pufe T. Expression and regulation of human beta-defensin-2 in osteoarthritic cartilage. J. Pathol. 2006;209:166–173. doi: 10.1002/path.1974. [DOI] [PubMed] [Google Scholar]
  156. Wang S., Rao P., Ye X. Isolation and biochemical characterization of a novel leguminous defense peptide with antifungal and antiproliferative potency. Appl. Microbiol. Biotechnol. 2009;82:79–86. doi: 10.1007/s00253-008-1729-2. [DOI] [PubMed] [Google Scholar]
  157. Wiesner J., Vilcinskas A. Antimicrobial peptides: the ancient arm of the human immune system. Virulence. 2010;1:440–464. doi: 10.4161/viru.1.5.12983. [DOI] [PubMed] [Google Scholar]
  158. Wilson N.J., Boniface K., Chan J.R., McKenzie B.S., Blumenschein W.M., Mattson J.D., Basham B., Smith K., Chen T., Morel F., Lecron J.C., Kastelein R.A., Cua D.J., McClanahan T.K., Bowman E.P., de Waal Malefyt R. Development, cytokine profile and function of human interleukin 17-producing helper T cells. Nat. Immunol. 2007;8:950–957. doi: 10.1038/ni1497. [DOI] [PubMed] [Google Scholar]
  159. Wnendt S., Ulbrich N., Stahl U. Molecular cloning, sequence analysis and expression of the gene encoding an antifungal-protein from Aspergillus giganteus. Curr. Genet. 1994;25:519–523. doi: 10.1007/BF00351672. [DOI] [PubMed] [Google Scholar]
  160. Wolk K., Kunz S., Witte E., Friedrich M., Asadullah K., Sabat R. IL-22 increases the innate immunity of tissues. Immunity. 2004;21:241–254. doi: 10.1016/j.immuni.2004.07.007. [DOI] [PubMed] [Google Scholar]
  161. Wong J.H., Ng T.B. Sesquin, a potent defensin-like antimicrobial peptide from ground beans with inhibitory activities toward tumor cells and HIV-1 reverse transcriptase. Peptides. 2005;26:1120–1126. doi: 10.1016/j.peptides.2005.01.003. [DOI] [PubMed] [Google Scholar]
  162. Wong J.H., Xia L., Ng T.B. A review of defensins of diverse origins. Curr. Protein Pept. Sci. 2007;8:446–459. doi: 10.2174/138920307782411446. [DOI] [PubMed] [Google Scholar]
  163. Wu Z., Hoover D.M., Yang D., Boulegue C., Santamaria F., Oppenheim J.J., Lubkowski J., Lu W. Engineering disulfide bridges to dissect antimicrobial and chemotactic activities of human beta-defensin 3. Proc. Natl. Acad. Sci. U.S.A. 2003;100:8880–8885. doi: 10.1073/pnas.1533186100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Yang D., Chertov O., Bykovskaia S.N., Chen Q., Buffo M.J., Shogan J., Anderson M., Schroder J.M., Wang J.M., Howard O.M., Oppenheim J.J. Beta-defensins: linking innate and adaptive immunity through dendritic and T cell CCR6. Science. 1999;286:525–528. doi: 10.1126/science.286.5439.525. [DOI] [PubMed] [Google Scholar]
  165. Yang D., Chen Q., Chertov O., Oppenheim J.J. Human neutrophil defensins selectively chemoattract naive T and immature dendritic cells. J. Leukoc. Biol. 2000;68:9–14. [PubMed] [Google Scholar]
  166. Yang D., Biragyn A., Kwak L.W., Oppenheim J.J. Mammalian defensins in immunity: more than just microbicidal. Trends Immunol. 2002;23:291–296. doi: 10.1016/s1471-4906(02)02246-9. [DOI] [PubMed] [Google Scholar]
  167. Yang D., Biragyn A., Hoover D.M., Lubkowski J., Oppenheim J.J. Multiple roles of antimicrobial defensins, cathelicidins, and eosinophil-derived neurotoxin in host defense. Annu. Rev. Immunol. 2004;22:181–215. doi: 10.1146/annurev.immunol.22.012703.104603. [DOI] [PubMed] [Google Scholar]
  168. Yudin A.I., Treece C.A., Tollner T.L., Overstreet J.W., Cherr G.N. The carbohydrate structure of DEFB126, the major component of the cynomolgus Macaque sperm plasma membrane glycocalyx. J. Membr. Biol. 2005;207:119–129. doi: 10.1007/s00232-005-0806-z. [DOI] [PubMed] [Google Scholar]
  169. Zhang Z.T., Zhu S.Y. Drosomycin, an essential component of antifungal defence in Drosophila. Insect Mol. Biol. 2009;18:549–556. doi: 10.1111/j.1365-2583.2009.00907.x. [DOI] [PubMed] [Google Scholar]
  170. Zhao C., Wang I., Lehrer R.I. Widespread expression of beta-defensin hBD-1 in human secretory glands and epithelial cells. FEBS Lett. 1996;396:319–322. doi: 10.1016/0014-5793(96)01123-4. [DOI] [PubMed] [Google Scholar]
  171. Zhou C.X., Zhang Y.L., Xiao L., Zheng M., Leung K.M., Chan M.Y., Lo P.S., Tsang L.L., Wong H.Y., Ho L.S., Chung Y.W., Chan H.C. An epididymis-specific beta-defensin is important for the initiation of sperm maturation. Nat. Cell Biol. 2004;6:458–464. doi: 10.1038/ncb1127. [DOI] [PubMed] [Google Scholar]
  172. Zhu S. Discovery of six families of fungal defensin-like peptides provides insights into origin and evolution of the CSalphabeta defensins. Mol. Immunol. 2008;45:828–838. doi: 10.1016/j.molimm.2007.06.354. [DOI] [PubMed] [Google Scholar]

RESOURCES