Plan
Comptes Rendus

Innate immunity / Immunité innée
Defensins: antimicrobial peptides of vertebrates
Comptes Rendus. Biologies, Volume 327 (2004) no. 6, pp. 539-549.

Résumés

This review, based on my presentation at the French Academy of Sciences on May 19, 2003, describes recent progress in the study of antimicrobial peptides, mediators of innate immunity in plants and animals. The main focus is on vertebrate defensins, a family of cysteine-rich antimicrobial peptides abundantly represented in human cells and tissues.

Cette revue, basée sur ma présentation à l'Académie des sciences de Paris du 19 mai 2003, décrit les progrès récents dans l'étude de peptides, médiateurs de l'immunité innée chez les plantes et les animaux. Les défensines des vertébrés, une famille de peptides antimicrobiens riches en cystéine abondamment représentés dans les cellules et les tissus humains, sont plus particulièrement considérées.

Métadonnées
Reçu le :
Accepté le :
Publié le :
DOI : 10.1016/j.crvi.2003.12.007
Keywords: antimicrobial peptides, innate immunity, defensins, cysteine
Mots clés : peptides antimicrobiens, immunité innée, défensines, cystéine
Tomas Ganz 1

1 CHS 37–055, Department of Medicine, David Geffen School of Medicine at UCLA, Los Angeles, CA 90095-1690, USA
@article{CRBIOL_2004__327_6_539_0,
     author = {Tomas Ganz},
     title = {Defensins: antimicrobial peptides of vertebrates},
     journal = {Comptes Rendus. Biologies},
     pages = {539--549},
     publisher = {Elsevier},
     volume = {327},
     number = {6},
     year = {2004},
     doi = {10.1016/j.crvi.2003.12.007},
     language = {en},
}
TY  - JOUR
AU  - Tomas Ganz
TI  - Defensins: antimicrobial peptides of vertebrates
JO  - Comptes Rendus. Biologies
PY  - 2004
SP  - 539
EP  - 549
VL  - 327
IS  - 6
PB  - Elsevier
DO  - 10.1016/j.crvi.2003.12.007
LA  - en
ID  - CRBIOL_2004__327_6_539_0
ER  - 
%0 Journal Article
%A Tomas Ganz
%T Defensins: antimicrobial peptides of vertebrates
%J Comptes Rendus. Biologies
%D 2004
%P 539-549
%V 327
%N 6
%I Elsevier
%R 10.1016/j.crvi.2003.12.007
%G en
%F CRBIOL_2004__327_6_539_0
Tomas Ganz. Defensins: antimicrobial peptides of vertebrates. Comptes Rendus. Biologies, Volume 327 (2004) no. 6, pp. 539-549. doi : 10.1016/j.crvi.2003.12.007. https://comptes-rendus.academie-sciences.fr/biologies/articles/10.1016/j.crvi.2003.12.007/

Version originale du texte intégral

1 Introduction

Antimicrobial peptides are polypeptides of fewer than 100 amino acids, found in host defense settings, and exhibiting antimicrobial activity at physiologic ambient conditions and peptide concentrations. Two large families of antimicrobial peptides, defensins [1] and cathelicidins [2], are abundant and widely distributed in mammalian epithelia and phagocytes. Other mammalian antimicrobial peptides, including histatins [3], dermcidin [4], and ‘anionic peptides’ [5] have a more restricted tissue and animal species distribution. Many more families of antimicrobial peptides are found in invertebrates [6]. Individual antimicrobial peptides have been implicated in antimicrobial activity of phagocytes, inflammatory body fluids or epithelial secretions. Although this review is primarily focused on mammalian defensins, I will use key studies of other antimicrobial peptides to illustrate more general principles of antimicrobial peptide function.

2 Antimicrobial peptides: activities and their mechanisms

2.1 Activity

Several model antimicrobial peptides, including magainins from frog skin, tachyplesins and polyphemusins from horseshoe crab hemocytes and protegrins from pig leukocytes, have been subjected to detailed studies because of their structural simplicity, small size (16–22 amino acids) and potential pharmaceutical applications. These peptides exhibit a broad spectrum of antimicrobial activity that includes gram-negative and gram-positive bacteria [7,8] and fungi [3], at similar concentrations (in the range of 1–10 μg ml−1) and under similar testing conditions to those used for other antimicrobial pharmaceuticals. Protegrins and tachyplesins are also active against some enveloped viruses [9]. The antimicrobial activity of these peptides is remarkably specific, with little cytotoxicity to mammalian cells even at concentrations ten-fold or more higher than those required for antimicrobial activity [10].

The structurally more complex mammalian defensins (29–50 amino acids) are also active against bacteria and fungi, especially when tested under low ionic strength conditions [11–13] and with low concentrations of divalent cations, plasma proteins or other interfering substances. Under these optimal conditions, antimicrobial activity is observed at concentrations as low as 1–10 μg ml−1 (low μM). Increasing concentrations of salts and plasma proteins competitively inhibit the antimicrobial activity of defensins, in a manner that is dependent on both the specific defensin and its microbial target. At higher concentrations, some defensins are cytotoxic to mammalian cells [14–16]. Certain enveloped viruses are also inactivated by defensins [17,18]. In general, metabolically active bacteria are much more sensitive to defensins then bacteria made inactive by nutrient deprivation or metabolic inhibitors.

2.2 Mechanisms of antimicrobial activity

Antimicrobial peptides are almost always cationic and amphipathic (i.e., they are positively-charged and contain both hydrophobic and hydrophilic domains). This allows them to interact with biological membranes in such a way that the cationic domains are near the negatively charged phospholipid headgroups, while the hydrophobic portions of the peptide are submerged within the hydrophobic interior of the membrane composed of fatty acid chains. The simplest antimicrobial peptides, typified by magainins, form an alpha helix, with cationic and hydrophobic side chains radially arranged on opposite surfaces of the helix. Another simple structure is the beta-sheet hairpin (e.g., protegrins, tachyplesins, polyphemusins) containing positively charged clusters separated by hydrophobic regions. The interactions between simple antimicrobial peptides and model membranes have been extensively explored [19–21]. In these systems, there is strong evidence of a two-stage interaction between the peptides and the membranes. During the first stage, the peptides, attracted to the membrane by electrostatic forces, form a carpet within but near the surface of the membrane, with the long axes of the peptides parallel to the membrane. As more and more peptide molecules accumulate, the membrane becomes distorted and strained, favoring a transition to an energetically more favorable state where the peptides are oriented with their long axes across the membrane, creating toroidal wormholes in the membrane or otherwise disrupting membrane integrity. These interactions are favored by the presence of anionic phospholipids in the membrane and inhibited by neutral phospholipids or cholesterol. Anionic phospholipids are characteristic of bacteria but neutral phospholipids and cholesterol are found in animal cell membranes, explaining the preferential effect of antimicrobial peptides on bacterial targets.

Structurally more complex antimicrobial peptides are thought to act by similar mechanisms. Model bacteria (E. coli ML-35) [22] and mammalian cell line K562 [16] treated by defensins become permeable to small molecules (small sugars and trypan respectively). In bacteria, permeabilization coincides with inhibition of RNA, DNA and protein synthesis and decreased bacterial viability, as assessed by the colony-forming assay. In the model cell line, the permeabilized cells can be rescued for up to 1 h by removing the defensin, and there is evidence that additional intracellular sites of action contribute to cell death [16].

In experiments with artificial phospholipid membranes, defensins NP-1 (rabbit) and HNP-1 (human) formed voltage-dependent channels, requiring negative potential on the membrane side opposite to where defensins where applied [23]. This is consistent with the idea that the insertion of defensin molecules into the membrane is dependent on electrical forces acting on the positively charged defensin molecule. The effect of these forces is evident even when no transmembrane potential is applied externally. Unlike another cationic peptide, melittin, that indiscriminately permeabilized vesicles composed of neutral or anionic phospholipids, defensins were much more active against vesicles that included negatively charged phospholipids [24]. In general, the activity of defensins against vesicles was diminished in the presence of increased salt concentrations, supporting the importance of electrostatic forces between the anionic phospholipids headgroups and the cationic defensins. Interestingly, the permeabilizing activity of the most highly cationic defensins was not inhibited by moderate salt concentrations, indicating that electrostatic screening by salt ions may not have been complete. In other experiments, large unilamellar vesicles composed of the negatively charged phospholipid palmitoyloleoylphosphatidylglycerol were permeabilized by human defensin HNP-2 but the addition of neutral phospholipids to the lipid mix inhibited both defensin binding and permeabilization [25]. Using a very different methodology, the importance of anionic phospholipids for the membrane interactions with defensins was clearly shown by calorimetric measurements of the effects of defensins on phase transitions in membranes [26]. These studies suggest that defensin molecules enter into the membrane under the influence of both externally applied and local electric fields.

It is much less certain what happens once the defensin molecules are in the membrane. The observed leakage of dye markers from liposomes implies that pores (we use this term to refer to any ion or water-permeable structure within the membrane) form either stably or transiently. For some defensins, the release of internal markers from each vesicle occurred in all or none fashion [25], indicating that the pores formed were stable. By measuring the ability of pores to allow the passage of marker molecules of various sizes, the pore diameter was estimated at 25 Å. The authors proposed a model of a defensin pore – a hexamer of dimers – that generates an opening of the observed size. However, stable pore formation is not the only mechanism of defensin interaction with membranes. The more cationic rabbit defensins induced a partial release of markers from individual vesicles indicating that the pores formed were not stable. It is possible that electrostatic repulsion between the highly cationic rabbit defensin molecules destabilizes the pores.

3 Vertebrate defensins

3.1 The structure of defensins

Defensins [1,27] are a family of vertebrate antimicrobial peptides with a characteristic β-sheet rich fold and a framework of six disulfide-linked cysteines [1,27]. The two major defensin subfamilies, α- and β-defensins, differ somewhat in cysteine spacing and connectivity (Fig. 1). Several structures representative of these two families have been solved by 2D-NMR and by X-ray crystallography [28–35]. Both α- and β-defensins consist of a triple stranded β-sheet with a distinctive ‘defensin’ fold (Fig. 2). Whereas in α-defensins the six cysteines are linked in the 1–6, 2–4, 3–5 pattern [36], in β-defensins the pattern is 1–5, 2–4, 3–6 [37]. Because cysteines 5 and 6 are adjacent in both types of defensins, this difference in connectivity does not substantially alter the structure [33]. More recently, another structurally very distinct subfamily of θ-defensins [38] has been identified in the rhesus macaque monkey leukocytes. The mature θ-defensin peptides arise by an as yet uncharacterized process that generates a cyclic peptide by splicing and cyclization from two 9-amino acid segments of α-defensin-like precursor peptides. Based on their adjacent chromosomal location and similar peptide precursor and gene structure it is highly likely that all vertebrate defensins arose from a common gene precursor [39]. Antimicrobial peptides from invertebrates and plants containing six or eight cysteines in disulfide linkage have also been called defensins (e.g., insect and plant defensins). Their evolutionary relationship to vertebrate defensins is uncertain.

Fig. 1

Amino acid sequences and connectivities of human defensin peptides.

Fig. 2

Cartoon diagrams of a human α-defensin and a human β-defensin. Note the similarity of the folding patterns of the monomers.

3.2 Distribution of defensins

During studies of the antimicrobial activity of rabbit and guinea pig leukocyte lysates in the 1960s, the peptides originally attracted attention because of their abundance and broad spectrum of antimicrobial activity [40]. Subsequent technical developments facilitated their isolation and detailed chemical characterization [11,41,42]. Their discovery in human leukocytes [1,27] suggested that the peptides were widely distributed in nature. After their isolation from leukocytes, defensins were found in other host defense settings where they were produced by epithelial cells [43,44]. Typical defensin peptides have been found in all mammals that have been carefully examined, as well as in chickens and turkeys [45–48]. Defensin-like peptides (growth arresting peptide [49] and crotamines) have been also isolated from snake venom where they may represent an adaptation of epithelial host defense peptides for efficacy against larger predators.

Defensins are found predominantly in cells and tissues involved in host defense against microbial infections. The highest concentrations of defensins (>10 mg ml−1) are found in granules, the storage organelles of leukocytes [1,50]. When leukocytes ingest microbes into phagocytic vacuoles, the granules fuse to these vacuoles and deliver their contents onto the target microbe. Since there is little free space in phagocytic vacuoles the microbe is exposed to minimally diluted granule material. Similarly, Paneth cells, specialized host defense cells of the small intestine, contain secretory granules that they release into narrow intestinal pits, called crypts. The concentration of defensins in the crypts may also reach >10 mg ml−1 [51]. Various epithelia produce defensins, in some cases constitutively [52], in others in response to infection [53]. The average concentration of defensins in these epithelia is in the 10–100-μg ml−1 range [53,54], but because the peptides are not evenly distributed the local concentrations could be much higher.

Patterns of tissue distribution are quite variable even when closely related species are compared. Among rodents, mice lack leukocyte defensins [55], rats have them [56] and both species have numerous Paneth-cell defensins and epithelial β-defensins. In some cases, defensin expression appears to be induced by a combination of a specific cell type and tissue environment. Inflammatory macrophages are leukocytes that arise by differentiation from circulating blood monocytes, under the influence of local tissue signals. In the rabbit, alveolar (lung) macrophages have abundant α-defensins in amounts comparable to rabbit neutrophils but defensins are absent from their peritoneal macrophages [57]. Although defensin expression in monocytes, macrophages and lymphocytes of some mammals can be detected by highly sensitive techniques [58–60], high levels of defensins in macrophages have only been documented in rabbits. We suspect that these peculiarities of the pattern of expression of defensins in certain animal species could be related to the evolutionary pressure from species-specific pathogens.

3.3 Microbial resistance to defensins

Specific mechanisms that confer increased bacterial resistance to defensins have been identified by insertional mutagenesis. Disruption of the two-component transcriptional regulator phoP–phoQ increases the sensitivity of Salmonella to defensins and other cationic peptides [61–64]. PhoP–phoQ directly regulates multiple genes involved in resistance to cationic peptides and also exerts some of its activity by modulating a second two-component regulator, PmrA–PmrB. The function of the downstream genes includes covalent modification of lipopolysaccharides that decreases their affinity for cationic peptides [65] and expression of membrane proteases that degrade cationic peptides [66]. In Neisseria gonorrhoeae, a bacterium naturally quite resistant to defensins, the energy-dependent efflux system mtr increases the resistance to protegrins, potent mini defensin-like peptides of pig neutrophils [67]. In Staphylococci, the disruption of either of two genes, dlt or MprF, increases the sensitivity of the bacteria to defensins [68,69]. The gene dlt is required for covalent modification of cell wall teichoic acid by alanine, and MprF is necessary for covalent modification of membrane phosphatidylglycerol with L-lysine. These modifications probably act by decreasing the negative charge of the cell wall and bacterial membrane respectively and diminishing their attraction for the cationic defensins. Homologues of these resistance genes have been identified in many bacterial species indicating that these mechanisms may be widespread.

3.4 Other activities of defensins

Various defensins have been reported to have chemotactic activity for monocytes, T-lymphocytes and dendritic cells [70–73]. In the case of human β-defensins 1 and 2, which attract memory T-cells and immature dendritic cells, the chemoattractant activity may be due to defensin binding to the chemokine receptor CCR6 [72]. Although the physiologic significance of this interaction has not yet been demonstrated, the high concentrations of HBD-2 in inflamed skin make it likely that this defensin could compete effectively with the natural chemokine ligand (variously named CCL20, LARC, MIP-3α) despite the higher affinity of the latter for the CCR6 receptor. Recent structural analysis of CCL20 pointed out remarkable similarities to HBD-2 in the putative receptor-binding region of CCL20. The role of this region in the chemotactic activity of HBD-2 needs to be confirmed by mutating the amino acid residues suspected in its interaction with CCR6. Human neutrophil defensins HNP1–3 have been reported to be chemotactic for monocytes [70], naı̈ve T-cells and immature dendritic cells [73] but a specific receptor has not yet been identified.

Some defensins (called ‘corticostatins’) [74–76] oppose the action of adrenocorticotropic hormone (ACTH) by binding to ACTH receptor [77] without activating it. Although such activity would inhibit the production of the immunosuppressive hormone cortisol, and could thus be useful in responding to infections, the physiologic role of this in vitro interaction has not yet been demonstrated.

Yet another reported activity of some defensins is their ability to activate nifedipine-sensitive calcium channels in mammalian cells [78,79]. This effect required only nanomolar concentrations of defensins. The structural basis of this effect is not understood. Certain mouse Paneth cell defensins (cryptdins) activate chloride secretion most likely by forming channels in the apical membrane of epithelial cells [80,81]. This activity is limited to a subset of cryptdins, and its structural basis is not yet known.

Most recently, several peptides genetically and structurally related to defensins have been found in the male reproductive tract, and in particular in the epididymis [82,83]. While some peptides expressed in the male reproductive tract are typical defensins also found in other organs [84], most are larger peptides from genes that undergo complex alternative splicing. These peptides could have an important role in the host defense of germ cells as well as in the regulation of sperm maturation.

3.5 Defensin biosynthetic pathways

At least eight genes encoding α- and β-defensins are located in a cluster on chromosome 8p23 [39,85–88] and recent studies document additional defensin clusters with multiple transcribed defensin genes [89]. Mapping of the 8p23 cluster has been problematic, probably due to its polymorphic nature, with individuals and their chromosomes differing in the number of copies of individual defensin genes [90,91]. Alpha-defensins are generally encoded as a tripartite prepropeptide sequence, wherein a 90–100 amino acid precursor contains an N-terminal ∼19 amino signal sequence, ∼45 amino acid anionic propiece and a C-terminal ∼30 amino acid mature cationic defensin [92] (Fig. 3). In many cases, the charge of the propiece and the mature defensin approximately balance [93], and this arrangement may be important for folding and/or to prevent intracellular interactions with membranes [94,95]. For neutrophil α-defensins, synthesis takes place in the bone marrow, in neutrophil precursor cells, promyelocytes [96–98]. Mature neutrophils circulating in blood or found in inflamed tissues contain large amounts of defensins but are no longer synthesizing the peptides or their mRNAs. During defensin synthesis in myeloid cell lines, the signal sequence is rapidly removed but the subsequent proteolytic processing to mature defensins takes many hours, and the final proteolytic cleavage may take place in maturing granules [99]. The process is very efficient so that only small amounts of partially processed intermediates are detectable in mature neutrophils [100]. In the case of murine Paneth cell defensins (cryptdins), the metalloproteinase matrilysin (MMP-7) is required for processing since mice with homozygous disruption of the matrilysin gene do not process Paneth-cell defensin past the removal of the signal sequence. The structure of β-defensin precursors is simpler, consisting of a signal sequence, a short or no propiece and the mature defensin peptide at the C-terminus. The lack of anionic propiece in β-defensin precursors contrasts with the relatively large anionic propiece in α-defensin precursors, a difference that has not been satisfactorily explained.

Fig. 3

Processing of the α-defensins HNP1-3 as deduced from studies in HL-60 myeloid leukemia cell line and in mature PMN. Arrows indicate the forms detected by direct analysis or radiosequencing. The segments are denoted as PRE (signal sequence), PRO (propiece) and MAT (mature peptide).

3.6 Amino acid sequence and composition of defensins

The amino acid sequences of mature defensins are highly variable, except for the conservation of the cystine framework (Fig. 3) in each defensin subfamily. Clusters of positively charged amino acids are characteristic of most α- and β-defensins, but their specific distribution within the defensin molecule is variable. In leukocytes and in Paneth cells of the small intestine, defensins are stored in granules, subcellular storage organelles rich in negatively charged glycosaminoglycans. With the exception of chicken gallinacins, these α- and β-defensins contain arginine as the predominant cationic amino acid. In contrast, β-defensins that are secreted from epithelial cells contain similar amounts of arginine and lysine. The preferential use of arginine in defensins stored in granules may reflect the constrains imposed by packing defensin molecules into the glycosaminoglycan matrix of granules [101,102].

3.7 Structure-function considerations

A unitary hypothesis of how defensins permeabilize membranes is complicated by the marked differences in net charge, amino acid sequence and quaternary structure (monomers vs. dimers) among the defensins. It is possible that these differences evolved so that various defensins can target different types of bacteria with differing structures of cell walls and membranes. Further complexity is introduced by the flexibility of the basic amino acid side chains that permit a variety of potential spatial interactions with phospholipids headgroups or water. Although the interactions of defensins with membranes have been modeled [25] there are only rudimentary experimental data on the structure of the defensin complexes within the membrane. Further work in this area is clearly needed since the considerable progress in understanding the interactions of amphipathic α-helical peptides with membranes does not readily translate to defensins, which are larger, more complicated and more variable structures.

3.8 Functions of defensins in vivo

When initially proposed [1], the name ‘defensins’ represented a risky conjecture since it was largely based on in vitro antimicrobial activity and the peptides' location in neutrophils, the prototypic host defense cells. Since then, experiments with transgenic mice have largely supported the idea that the dominant function of defensins is antimicrobial. Mice with homozygous disruption of the matrilysin gene failed to activate intestinal prodefensins to defensins, and were more susceptible to infection with Salmonella typhimurium, requiring an eight-fold lower oral dose for 50% mortality [103]. After oral administration of E. coli test bacteria, the counts of viable bacteria were similar in the proximal intestine of wild-type and matrilysin-knockout mice, but the wild-type mice had lower bacterial counts in the mid- and distal small intestine, where Paneth cells are present at higher density. In vitro, segments of intestine from wild-type mice contained and secreted more antimicrobial activity than those of matrilysin knockout mice [51]. Moreover, in wild-type mice the antimicrobial activity could be largely neutralized by anti-defensin antibody, indicating that defensins were responsible for much of the activity. Taken together, these experiments provided important circumstantial evidence for the protective role of defensins in the early stages of infection.

More recently, a gain-of-function model was reported, in transgenic mice expressing the human Paneth cell defensin gene HD-5 [104]. HD-5 compared to murine Paneth cell defensins has greater antibacterial potency against the murine pathogen Salmonella typhimurium. HD-5 mice were fully protected against death from Salmonella typhimurium infection at oral doses that killed all of the wild-type mice. Protection from infection was seen early, already at 6 hours, and correlated with lower S. typhimurium counts in the intestinal lumen, and prevention of the spread of infection to other organs. The effect of transgenic defensin was local, since intraperitoneal inoculation that bypassed the intestine caused equal mortality in the transgenic and wild-type strains. The intestinal lumen-specific effects of transgenic defensin early in the course of infection provide the strongest evidence to date that defensins act as locally secreted antibiotics.

Mice deficient in murine β-defensin-1 show only very mild defects [105,106] in host defense of the urinary and respiratory tracts, most likely due to the redundancy amongst mouse defensin genes. Unlike the many defensin genes present in the mouse genome, there is only one, or at most very few murine cathelicidins (the number depends on how the family is defined). The murine cathelicidin (cathelin-related antimicrobial peptide, CRAMP) is similar to its human ortholog, LL-37, and both are expressed in predominantly in neutrophils. Mice with homozygous disruption of the CRAMP gene showed diminished resistance to skin infection with group A Streptococcus [107]. Taken together, data from loss of function models support a host defense role for cathelicidins and defensins.

4 Conclusions and future prospects

Evidence is continuing to accumulate that vertebrate defensins function as antimicrobial effectors of innate immunity. In addition, some defensins may have also evolved additional roles in host defense, inflammation and even reproduction. Some of the many structurally and genetically diverse antimicrobial peptides in animals and plants should provide useful templates for the development of new antibiotics.


Bibliographie

[1] T. Ganz; M.E. Selsted; D. Szklarek; S.S. Harwig; K. Daher; D.F. Bainton; R.I. Lehrer Defensins. Natural peptide antibiotics of human neutrophils, J. Clin. Invest., Volume 76 (1985), pp. 1427-1435

[2] M. Zanetti; R. Gennaro; D. Romeo Cathelicidins: a novel protein family with a common proregion and a variable C-terminal antimicrobial domain, FEBS Lett., Volume 374 (1995), pp. 1-5

[3] H. Tsai; L.A. Bobek Human salivary histatins: promising anti-fungal therapeutic agents, Crit. Rev. Oral Biol. Med., Volume 9 (1998), pp. 480-497

[4] B. Schittek; R. Hipfel; B. Sauer; J. Bauer; H. Kalbacher; S. Stevanovic; M. Schirle; K. Schroeder; N. Blin; F. Meier; G. Rassner; C. Garbe Dermcidin: a novel human antibiotic peptide secreted by sweat glands, Nat. Immunol., Volume 2 (2001), pp. 1133-1137

[5] K.A. Brogden; M. Ackermann; K.M. Huttner Small, anionic, and charge-neutralizing propeptide fragments of zymogens are antimicrobial, Antimicrob. Agents Chemother., Volume 41 (1997), pp. 1615-1617

[6] J.A. Hoffmann; F.C. Kafatos; C.A. Janeway; R.A. Ezekowitz Phylogenetic perspectives in innate immunity, Science, Volume 284 (1999), pp. 1313-1318

[7] Y. Ge; D.L. MacDonald; K.J. Holroyd; C. Thornsberry; H. Wexler; M. Zasloff In vitro antibacterial properties of pexiganan, an analog of magainin, Antimicrob. Agents Chemother., Volume 43 (1999), pp. 782-788

[8] D.A. Steinberg; M.A. Hurst; C.A. Fujii; A.H. Kung; J.F. Ho; F.C. Cheng; D.J. Loury; J.C. Fiddes Protegrin-1: a broad-spectrum, rapidly microbicidal peptide with in vivo activity, Antimicrob. Agents Chemother., Volume 41 (1997), pp. 1738-1742

[9] B. Yasin; M. Pang; J.S. Turner; Y. Cho; N.N. Dinh; A.J. Waring; R.I. Lehrer; E.A. Wagar Evaluation of the inactivation of infectious Herpes simplex virus by host-defense peptides, Eur. J. Clin. Microbiol. Infect. Dis., Volume 19 (2000), pp. 187-194

[10] B. Haimovich; J.C. Tanaka Magainin-induced cytotoxicity in eukaryotic cells: kinetics, dose-response and channel characteristics, Biochim. Biophys. Acta, Volume 1240 (1995), pp. 149-158

[11] M.E. Selsted; D. Szklarek; R.I. Lehrer Purification and antibacterial activity of antimicrobial peptides of rabbit granulocytes, Infect. Immun., Volume 45 (1984), pp. 150-154

[12] M.E. Selsted; D. Szklarek; T. Ganz; R.I. Lehrer Activity of rabbit leukocyte peptides against Candida albicans, Infect. Immun., Volume 49 (1985), pp. 202-206

[13] R.I. Lehrer; T. Ganz; D. Szklarek; M.E. Selsted Modulation of the in vitro candidacidal activity of human neutrophil defensins by target cell metabolism and divalent cations, J. Clin. Invest., Volume 81 (1988), pp. 1829-1835

[14] A. Lichtenstein; T. Ganz; M.E. Selsted; R.I. Lehrer In vitro tumor cell cytolysis mediated by peptide defensins of human and rabbit granulocytes, Blood, Volume 68 (1986), pp. 1407-1410

[15] A.K. Lichtenstein; T. Ganz; T.M. Nguyen; M.E. Selsted; R.I. Lehrer Mechanism of target cytolysis by peptide defensins. Target cell metabolic activities, possibly involving endocytosis, are crucial for expression of cytotoxicity, J. Immunol., Volume 140 (1988), pp. 2686-2694

[16] A. Lichtenstein Mechanism of mammalian cell lysis mediated by peptide defensins. Evidence for an initial alteration of the plasma membrane, J. Clin. Invest., Volume 88 (1991), pp. 93-100

[17] R.I. Lehrer; K. Daher; T. Ganz; M.E. Selsted Direct inactivation of viruses by MCP-1 and MCP-2, natural peptide antibiotics from rabbit leukocytes, J. Virol., Volume 54 (1985), pp. 467-472

[18] K.A. Daher; M.E. Selsted; R.I. Lehrer Direct inactivation of viruses by human granulocyte defensins, J. Virol., Volume 60 (1986), pp. 1068-1074

[19] H.W. Huang Action of antimicrobial peptides: two-state model, Biochemistry, Volume 39 (2000), pp. 8347-8352

[20] Y. Shai Mechanism of the binding, insertion and destabilization of phospholipid bilayer membranes by alpha-helical antimicrobial and cell non-selective membrane-lytic peptides, Biochim. Biophys. Acta, Volume 1462 (1999), pp. 55-70

[21] K. Matsuzaki Magainins as paradigm for the mode of action of pore forming polypeptides, Biochim. Biophys. Acta, Volume 1376 (1998), pp. 391-400

[22] R.I. Lehrer; A. Barton; K.A. Daher; S.S. Harwig; T. Ganz; M.E. Selsted Interaction of human defensins with Escherichia coli. Mechanism of bactericidal activity, J. Clin. Invest., Volume 84 (1989), pp. 553-561

[23] B.L. Kagan; M.E. Selsted; T. Ganz; R.I. Lehrer Antimicrobial defensin peptides form voltage-dependent ion-permeable channels in planar lipid bilayer membranes, Proc. Natl Acad. Sci. USA, Volume 87 (1990), pp. 210-214

[24] G. Fujii; M.E. Selsted; D. Eisenberg Defensins promote fusion and lysis of negatively charged membranes, Protein Sci., Volume 2 (1993), pp. 1301-1312

[25] W.C. Wimley; M.E. Selsted; S.H. White Interactions between human defensins and lipid bilayers: evidence for formation of multimeric pores, Protein Sci., Volume 3 (1994), pp. 1362-1373

[26] K. Lohner; A. Latal; R.I. Lehrer; T. Ganz Differential scanning microcalorimetry indicates that human defensin, HNP-2, interacts specifically with biomembrane mimetic systems, Biochemistry, Volume 36 (1997), pp. 1525-1531

[27] M.E. Selsted; S.S. Harwig; T. Ganz; J.W. Schilling; R.I. Lehrer Primary structures of three human neutrophil defensins, J. Clin. Invest., Volume 76 (1985), pp. 1436-1439

[28] C.P. Hill; J. Yee; M.E. Selsted; D. Eisenberg Crystal structure of defensin HNP-3, an amphiphilic dimer: mechanisms of membrane permeabilization, Science, Volume 251 (1991), pp. 1481-1485

[29] A. Pardi; D.R. Hare; M.E. Selsted; R.D. Morrison; D.A. Bassolino; A.C.2d Bach Solution structures of the rabbit neutrophil defensin NP- 5, J. Mol. Biol., Volume 201 (1988), pp. 625-636

[30] A. Pardi; X.L. Zhang; M.E. Selsted; J.J. Skalicky; P.F. Yip NMR studies of defensin antimicrobial peptides. 2. Three- dimensional structures of rabbit NP-2 and human HNP-1, Biochemistry, Volume 31 (1992), pp. 11357-11364

[31] X.L. Zhang; M.E. Selsted; A. Pardi NMR studies of defensin antimicrobial peptides. 1. Resonance assignment and secondary structure determination of rabbit NP-2 and human HNP-1, Biochemistry, Volume 31 (1992), pp. 11348-11356

[32] J.J. Skalicky; M.E. Selsted; A. Pardi Structure and dynamics of the neutrophil defensins NP-2, NP-5, and HNP-1: NMR studies of amide hydrogen exchange kinetics, Proteins, Volume 20 (1994), pp. 52-67

[33] G.R. Zimmermann; P. Legault; M.E. Selsted; A. Pardi Solution structure of bovine neutrophil beta-defensin-12: the peptide fold of the beta-defensins is identical to that of the classical defensins, Biochemistry, Volume 34 (1995), pp. 13663-13671

[34] D.M. Hoover; K.R. Rajashankar; R. Blumenthal; A. Puri; J.J. Oppenheim; O. Chertov; J. Lubkowski The structure of human beta-defensin-2 shows evidence of higher-order oligomerization, J. Biol. Chem. (2000)

[35] M.V. Sawai; H.P. Jia; L. Liu; V. Aseyev; J.M. Wiencek; P.B. McCray; T. Ganz; W.R. Kearney; B.F. Tack The NMR structure of human beta-defensin-2 reveals a novel alpha- helical segment, Biochemistry, Volume 40 (2001), pp. 3810-3816

[36] M.E. Selsted; S.S. Harwig Determination of the disulfide array in the human defensin HNP-2. A covalently cyclized peptide, J. Biol. Chem., Volume 264 (1989), pp. 4003-4007

[37] Y.Q. Tang; M.E. Selsted Characterization of the disulfide motif in BNBD-12, an antimicrobial beta-defensin peptide from bovine neutrophils, J. Biol. Chem., Volume 268 (1993), pp. 6649-6653

[38] Y.Q. Tang; J. Yuan; G. Osapay; K. Osapay; D. Tran; C.J. Miller; A.J. Ouellette; M.E. Selsted A cyclic antimicrobial peptide produced in primate leukocytes by the ligation of two truncated alpha-defensins, Science, Volume 286 (1999), pp. 498-502 (see comments)

[39] L. Liu; C. Zhao; H.H.Q. Heng; T. Ganz The human β-defensin-1 and α-defensins are encoded by adjacent genes: two peptide families with differing disulfide topology share a common ancestry, Genomics, Volume 43 (1997), pp. 316-320

[40] H.I. Zeya; J.K. Spitznagel Antibacterial and enzymic basic proteins from leukocyte lysosomes: separation and identification, Science, Volume 142 (1963), pp. 1085-1087

[41] R.I. Lehrer; M.E. Selsted; D. Szklarek; J. Fleischmann Antibacterial activity of microbicidal cationic proteins 1 and 2, natural peptide antibiotics of rabbit lung macrophages, Infect. Immun., Volume 42 (1983), pp. 10-14

[42] M.E. Selsted; D.M. Brown; R.J. DeLange; R.I. Lehrer Primary structures of MCP-1 and MCP-2, natural peptide antibiotics of rabbit lung macrophages, J. Biol. Chem., Volume 258 (1983), pp. 14485-14489

[43] A.J. Ouellette; R.M. Greco; M. James; D. Frederick; J. Naftilan; J.T. Fallon Developmental regulation of cryptdin, a corticostatin/defensin precursor mRNA in mouse small intestinal crypt epithelium, J. Cell Biol., Volume 108 (1989), pp. 1687-1695

[44] G. Diamond; M. Zasloff; H. Eck; M. Brasseur; W.L. Maloy; C.L. Bevins Tracheal antimicrobial peptide, a cysteine-rich peptide from mammalian tracheal mucosa: peptide isolation and cloning of a cDNA, Proc. Natl Acad. Sci. USA, Volume 88 (1991), pp. 3952-3956

[45] S.S. Harwig; K.M. Swiderek; V.N. Kokryakov; L. Tan; T.D. Lee; E.A. Panyutich; G.M. Aleshina; O.V. Shamova; R.I. Lehrer Gallinacins: cysteine-rich antimicrobial peptides of chicken leukocytes, FEBS Lett., Volume 342 (1994), pp. 281-285

[46] C. Zhao; T. Nguyen; L. Liu; R.E. Sacco; K.A. Brogden; R.I. Lehrer Gallinacin-3, an inducible epithelial beta-defensin in the chicken, Infect. Immun., Volume 69 (2001), pp. 2684-2691

[47] E.W. Evans; G.G. Beach; J. Wunderlich; B.G. Harmon Isolation of antimicrobial peptides from avian heterophils, J. Leukoc. Biol., Volume 56 (1994), pp. 661-665

[48] C.W. Brockus; M.W. Jackwood; B.G. Harmon Characterization of beta-defensin prepropeptide mRNA from chicken and turkey bone marrow, Anim. Genet., Volume 29 (1998), pp. 283-289

[49] H. Marquardt, G.J. Todaro, D.R. Twardzik, Snake venom growth arresting peptide, Oncogen, Seattle, WA, US Patent No. 4774318-A 3, 9–27–1988

[50] T. Ganz Extracellular release of antimicrobial defensins by human polymorphonuclear leukocytes, Infect. Immun., Volume 55 (1987), pp. 568-571

[51] T. Ayabe; D.P. Satchell; C.L. Wilson; W.C. Parks; M.E. Selsted; A.J. Ouellette Secretion of microbicidal α-defensins by intestinal Paneth cells in response to bacteria, Nat. Immunol., Volume 1 (2000), pp. 113-118

[52] E.V. Valore; C.H. Park; A.J. Quayle; K.R. Wiles; P.B. McCray; T. Ganz Human beta-defensin-1: an antimicrobial peptide of urogenital tissues, J. Clin. Invest., Volume 101 (1998), pp. 1633-1642

[53] J. Harder; J. Bartels; E. Christophers; J.-M. Schroeder A peptide antibiotic from human skin, Nature, Volume 387 (1997), pp. 861-862

[54] J. Shi; G. Zhang; H. Wu; C. Ross; F. Blecha; T. Ganz Porcine epithelial beta-defensin 1 is expressed in the dorsal tongue at antimicrobial concentrations, Infect. Immun., Volume 67 (1999), pp. 3121-3127

[55] P.B. Eisenhauer; R.I. Lehrer Mouse neutrophils lack defensins, Infect. Immun., Volume 60 (1992), pp. 3446-3447

[56] P.B. Eisenhauer; S.S. Harwig; D. Szklarek; T. Ganz; M.E. Selsted; R.I. Lehrer Purification and antimicrobial properties of three defensins from rat neutrophils, Infect. Immun., Volume 57 (1989), pp. 2021-2027

[57] T. Ganz; J.R. Rayner; E.V. Valore; A. Tumolo; K. Talmadge; F. Fuller The structure of the rabbit macrophage defensin genes and their organ-specific expression, J. Immunol., Volume 143 (1989), pp. 1358-1365

[58] L.K. Ryan; J. Rhodes; M. Bhat; G. Diamond Expression of beta-defensin genes in bovine alveolar macrophages, Infect. Immun., Volume 66 (1998), pp. 878-881

[59] L.A. Duits; M. Rademaker; B. Ravensbergen; M.A. van Sterkenburg; E. van Strijen; P.S. Hiemstra; P.H. Nibbering Inhibition of hBD-3, but not hBD-1 and hBD-2, mRNA expression by corticosteroids, Biochem. Biophys. Res. Commun., Volume 280 (2001), pp. 522-525

[60] B. Agerberth; J. Charo; J. Werr; B. Olsson; F. Idali; L. Lindbom; R. Kiessling; H. Jornvall; H. Wigzell; G.H. Gudmundsson The human antimicrobial and chemotactic peptides LL-37 and alpha-defensins are expressed by specific lymphocyte and monocyte populations, Blood, Volume 96 (2000), pp. 3086-3093

[61] E.A. Groisman; E. Chiao; C.J. Lipps; F. Heffron Salmonella typhimurium phoP virulence gene is a transcriptional regulator, Proc. Natl Acad. Sci. USA, Volume 86 (1989), pp. 7077-7081

[62] S.I. Miller; W.S. Pulkkinen; M.E. Selsted; J.J. Mekalanos Characterization of defensin resistance phenotypes associated with mutations in the phoP virulence regulon of Salmonella typhimurium, Infect. Immun., Volume 58 (1990), pp. 3706-3710

[63] S.I. Miller PhoP/PhoQ: macrophage-specific modulators of Salmonella virulence?, Mol. Microbiol., Volume 5 (1991), pp. 2073-2078

[64] E.A. Groisman; F. Heffron; F. Solomon Molecular genetic analysis of the Escherichia coli phoP locus, J. Bacteriol., Volume 174 (1992), pp. 486-491

[65] L. Guo; K.B. Lim; C.M. Poduje; M. Daniel; J.S. Gunn; M. Hackett; S.I. Miller Lipid A acylation and bacterial resistance against vertebrate antimicrobial peptides, Cell, Volume 95 (1998), pp. 189-198

[66] T. Guina; E.C. Yi; H. Wang; M. Hackett; S.I. Miller A PhoP-regulated outer membrane protease of Salmonella enterica serovar typhimurium promotes resistance to alpha-helical antimicrobial peptides, J. Bacteriol., Volume 182 (2000), pp. 4077-4086

[67] W.M. Shafer; X. Qu; A.J. Waring; R.I. Lehrer Modulation of Neisseria gonorrhoeae susceptibility to vertebrate antibacterial peptides due to a member of the resistance/nodulation/division efflux pump family, Proc. Natl Acad. Sci. USA, Volume 95 (1998), pp. 1829-1833

[68] A. Peschel; M. Otto; R.W. Jack; H. Kalbacher; G. Jung; F. Gotz Inactivation of the dlt operon in Staphylococcus aureus confers sensitivity to defensins, protegrins, and other antimicrobial peptides, J. Biol. Chem., Volume 274 (1999), pp. 8405-8410

[69] A. Peschel; R.W. Jack; M. Otto; L.V. Collins; P. Staubitz; G. Nicholson; H. Kalbacher; W.F. Nieuwenhuizen; G. Jung; A. Tarkowski; K.P. van Kessel; J.A. van Strijp Staphylococcus aureus resistance to human defensins and evasion of neutrophil killing via the novel virulence factor MprF is based on modification of membrane lipids with l-lysine, J. Exp. Med., Volume 193 (2001), pp. 1067-1076

[70] M.C. Territo; T. Ganz; M.E. Selsted; R. Lehrer Monocyte-chemotactic activity of defensins from human neutrophils, J. Clin. Invest., Volume 84 (1989), pp. 2017-2020

[71] O. Chertov; D.F. Michiel; L. Xu; J.M. Wang; K. Tani; W.J. Murphy; D.L. Longo; D.D. Taub; J.J. Oppenheim Identification of defensin-1, defensin-2, and CAP37/azurocidin as T-cell chemoattractant proteins released from interleukin-8- stimulated neutrophils, J. Biol. Chem., Volume 271 (1996), pp. 2935-2940

[72] D. Yang; O. Chertov; S.N. Bykovskaia; Q. Chen; M.J. Buffo; J. Shogan; M. Anderson; J.M. Schroder; J.M. Wang; O.M. Howard; J.J. Oppenheim Beta-defensins: linking innate and adaptive immunity through dendritic and T cell CCR6, Science, Volume 286 (1999), pp. 525-528 (see comments)

[73] D. Yang; Q. Chen; O. Chertov; J.J. Oppenheim Human neutrophil defensins selectively chemoattract naive T and immature dendritic cells, J. Leukoc. Biol., Volume 68 (2000), pp. 9-14

[74] Q.Z. Zhu; A.V. Singh; A. Bateman; F. Esch; S. Solomon The corticostatic (anti-ACTH) and cytotoxic activity of peptides isolated from fetal, adult and tumor-bearing lung, J. Steroid. Biochem., Volume 27 (1987), pp. 1017-1022

[75] Q. Zhu; A. Bateman; A. Singh; S. Solomon Isolation and biological activity of corticostatic peptides (anti-ACTH), Endocr. Res., Volume 15 (1989), pp. 129-149

[76] S. Solomon; J. Hu; Q. Zhu; D. Belcourt; H.P. Bennett; A. Bateman; T. Antakly Corticostatic peptides, J. Steroid Biochem. Mol. Biol., Volume 40 (1991), pp. 391-398

[77] T. Tominaga; J. Fukata; Y. Naito; Y. Nakai; S. Funakoshi; N. Fujii; H. Imura Effects of corticostatin-I on rat adrenal cells in vitro, J. Endocrinol., Volume 125 (1990), pp. 287-292

[78] R.J. MacLeod; J.R. Hamilton; A. Bateman; D. Belcourt; J. Hu; H.P. Bennett; S. Solomon Corticostatic peptides cause nifedipine-sensitive volume reduction in jejunal villus enterocytes, Proc. Natl Acad. Sci. USA, Volume 88 (1991), pp. 552-556

[79] A. Bateman; R.J. MacLeod; P. Lembessis; J. Hu; F. Esch; S. Solomon The isolation and characterization of a novel corticostatin/defensin-like peptide from the kidney, J. Biol. Chem., Volume 271 (1996), pp. 10654-10659

[80] W.I. Lencer; G. Cheung; G.R. Strohmeier; M.G. Currie; A.J. Ouellette; M.E. Selsted; J.L. Madara Induction of epithelial chloride secretion by channel-forming cryptdins 2 and 3, Proc. Natl Acad. Sci. USA, Volume 94 (1997), pp. 8585-8589

[81] D. Merlin; G. Yue; W.I. Lencer; M.E. Selsted; J.L. Madara Cryptdin-3 induces novel apical conductance(s) in Cl-secretory, including cystic fibrosis, epithelia, Am. J. Physiol. Cell Physiol., Volume 280 (2001), p. C296-C302

[82] O. Frohlich; C. Po; T. Murphy; L.G. Young Multiple promoter and splicing mRNA variants of the epididymis-specific gene EP2, J. Androl., Volume 21 (2000), pp. 421-430

[83] P. Li; H.C. Chan; B. He; S.C. So; Y.W. Chung; Q. Shang; Y.D. Zhang; Y.L. Zhang An antimicrobial peptide gene found in the male reproductive system of rats, Science, Volume 291 (2001), pp. 1783-1785

[84] E. Com; F. Bourgeon; B. Evrard; T. Ganz; D. Colleu; B. Jegou; C. Pineau Expression of antimicrobial defensins in the male reproductive tract of rats, mice, and humans, Biol. Reprod., Volume 68 (2003), pp. 95-104

[85] R.S. Sparkes; M. Kronenberg; C. Heinzmann; K.A. Daher; I. Klisak; T. Ganz; T. Mohandas Assignment of defensin gene(s) to human chromosome 8p23, Genomics, Volume 5 (1989), pp. 240-244

[86] J. Harder; R. Siebert; Y. Zhang; P. Matthiesen; E. Christophers; B. Schlegelberger; J.M. Schroder Mapping of the gene encoding human beta-defensin-2 (DEFB2) to chromosome region 8p22-p23.1, Genomics, Volume 46 (1997), pp. 472-475

[87] L. Liu; L. Wang; H.P. Jia; C. Zhao; H.H.Q. Heng; B.C. Schutte; P.B.J. McCray; T. Ganz Structure and mapping of the human β-defensin HBD-2 gene and its expression at sites of inflammation, Gene, Volume 222 (1998), pp. 237-244

[88] R. Linzmeier; C.H. Ho; B.V. Hoang; T. Ganz A 450-kb contig of defensin genes on human chromosome 8p23, Gene, Volume 233 (1999), pp. 205-211

[89] B.C. Schutte; J.P. Mitros; J.A. Bartlett; J.D. Walters; H.P. Jia; M.J. Welsh; T.L. Casavant; P.B. McCray Discovery of five conserved beta-defensin gene clusters using a computational search strategy, Proc. Natl Acad. Sci. USA, Volume 99 (2002), pp. 2129-2133

[90] W.M. Mars; P. Patmasiriwat; T. Maity; V. Huff; M.M. Weil; G.F. Saunders Inheritance of unequal numbers of the genes encoding the human neutrophil defensins HP-1 and HP-3, J. Biol. Chem., Volume 270 (1995), pp. 30371-30376

[91] E.J. Hollox; J.A. Armour; J.C. Barber Extensive normal copy number variation of a beta-defensin antimicrobial-gene cluster, Am. J. Hum. Genet., Volume 73 (2003), pp. 591-600

[92] K.A. Daher; R.I. Lehrer; T. Ganz; M. Kronenberg Isolation and characterization of human defensin cDNA clones, Proc. Natl Acad. Sci. USA, Volume 85 (1988), pp. 7327-7331

[93] D. Michaelson; J. Rayner; M. Couto; T. Ganz Cationic defensins arise from charge-neutralized propeptides: a mechanism for avoiding leukocyte autocytotoxicity?, J. Leukoc. Biol., Volume 51 (1992), pp. 634-639

[94] E.V. Valore; E. Martin; S.S. Harwig; T. Ganz Intramolecular inhibition of human defensin HNP-1 by its propiece, J. Clin. Invest., Volume 97 (1996), pp. 1624-1629

[95] L. Liu; T. Ganz The pro region of human neutrophil defensin contains a motif that is essential for normal subcellular sorting, Blood, Volume 85 (1995), pp. 1095-1103

[96] N.Y. Yount; M.S.C. Wang; J. Yuan; N. Banaiee; A. Ouellette; M.E. Selsted Rat neutrophil defensins. Precursor structures and expression during neutrophilic myelopoiesis, J. Immunol., Volume 155 (1995), pp. 4476-4484

[97] K. Arnljots; O. Sorensen; K. Lollike; N. Borregaard Timing, targeting and sorting of azurophil granule proteins in human myeloid cells, Leukemia, Volume 12 (1998), pp. 1789-1795

[98] J.B. Cowland; N. Borregaard The individual regulation of granule protein mRNA levels during neutrophil maturation explains the heterogeneity of neutrophil granules, J. Leukoc. Biol., Volume 66 (1999), pp. 989-995

[99] E.V. Valore; T. Ganz Posttranslational processing of defensins in immature human myeloid cells, Blood, Volume 79 (1992), pp. 1538-1544

[100] S.S. Harwig; A.S. Park; R.I. Lehrer Characterization of defensin precursors in mature human neutrophils, Blood, Volume 79 (1992), pp. 1532-1537

[101] J.R. Fromm; R.E. Hileman; E.E. Caldwell; J.M. Weiler; R.J. Linhardt Differences in the interaction of heparin with arginine and lysine and the importance of these basic amino acids in the binding of heparin to acidic fibroblast growth factor, Arch. Biochem. Biophys., Volume 323 (1995), pp. 279-287

[102] R.E. Hileman; J.R. Fromm; J.M. Weiler; R.J. Linhardt Glycosaminoglycan-protein interactions: definition of consensus sites in glycosaminoglycan binding proteins, Bioessays, Volume 20 (1998), pp. 156-167

[103] C.L. Wilson; A.J. Ouellette; D.P. Satchell; T. Ayabe; Y.S. Lopez-Boado; J.L. Stratman; S.J. Hultgren; L.M. Matrisian; W.C. Parks Regulation of intestinal alpha-defensin activation by the metalloproteinase matrilysin in innate host defense, Science, Volume 286 (1999), pp. 113-117

[104] N.H. Salzman; D. Ghosh; K.M. Huttner; Y. Paterson; C.L. Bevins Protection against enteric salmonellosis in transgenic mice expressing a human intestinal defensin, Nature, Volume 422 (2003), pp. 522-526

[105] G. Morrison; F. Kilanowski; D. Davidson; J. Dorin Characterization of the mouse beta defensin 1, Defb1, mutant mouse model, Infect. Immun., Volume 70 (2002), pp. 3053-3060

[106] C. Moser; D.J. Weiner; E. Lysenko; R. Bals; J.N. Weiser; J.M. Wilson beta-Defensin 1 contributes to pulmonary innate immunity in mice, Infect. Immun., Volume 70 (2002), pp. 3068-3072

[107] V. Nizet; T. Ohtake; X. Lauth; J. Trowbridge; J. Rudisill; R.A. Dorschner; V. Pestonjamasp; J. Piraino; K. Huttner; R.L. Gallo Innate antimicrobial peptide protects the skin from invasive bacterial infection, Nature, Volume 414 (2001), pp. 454-457


Commentaires - Politique


Ces articles pourraient vous intéresser

Plant–pathogen interactions: will the understanding of common mechanisms lead to the unification of concepts?

Philippe Reignault; Michel Sancholle

C. R. Biol (2005)