Você está na página 1de 11

JFS R: Concise Reviews/Hypotheses in Food Science

Eukaryotic Antimicrobial Peptides: Promises and Premises in Food Safety


TALI RYDLO, JOSEPH MILTZ, AND AMRAM MOR
ABSTRACT: There is a lack of efficient and safe preservatives in the food industry. Massive use of some common food preservation methods has led, over the years, to development of a resistance to different treatments by various food pathogens. Enteric bacteria are especially tolerant to adverse environmental conditionssuch as low pH and high salt concentrations which limits efficiency of some preservation methods. Consumers demand for natural, preservative-free, and minimally processed foods and worldwide concern regarding disease outbreaks caused by food-related pathogens have created a need for development of new classes of antimicrobial (AM) agents. The twentieth century revealed a massive array of new peptide-based antimicrobials. Small ribosomally made compounds are found in practically all living species where they act as important component of host defense. Certain indubitable advantages of peptidespertaining to simplicity, activity spectra, and bacterial resistanceover known preservative agents advocate their potential for food preservation. Nisin, an AM compound originating from bacteria, is so far the only FDA-approved peptide. However, a growing number of reports describe the potential of animal-derived antimicrobial peptides as food preservatives. These studies have yielded various native compounds and/or derivatives that possess markedly improved antimicrobial properties under a broad range of incubation conditions. The present work reviews the most investigated peptides and accounts for their potential use as alternatives to the preservatives used today. The focus is on research aspects aiming at understanding the mechanism of action of these peptides at extreme environments of various food systems. Collectively, the data accumulated are convincingly indicative of potential applications of these peptides in food safety, namely, with respect to fighting multidrug-resistant pathogens. Keywords: antimicrobial packaging, antimicrobial resistance, drug design, food pathogens

he use, and sometimes misuse, of antimicrobials in both human and veterinary medicine during the years has resulted in the emergence of bacterial strains that no longer respond to antimicrobial therapy. So far, resistance has developed to almost all antimicrobial drugs (Gold and Moellering 1996). Moreover, bacterial strains resistant to all available antimicrobial agents have been identified among clinical isolates of different bacterial species (Tomasz 1994). A study aiming at evaluating the resistance of 5 major foodborne pathogens (Campylobacter, Salmonella, Yersinia enterocolitica, pathogenic Escherichia coli, and Listeria monocytogenes) isolated from 922 meat products showed that enteric bacteria are highly resistant against a wide range of antimicrobial agents and also possess multiresistance phenotypes of distinct advantages (Mayrhofer and others 2004). The alarming emergence of resistance among bacteria has led the World Health Organization (WHO) to announce antimicrobial drug resistance as a main public health concern (WHO 1995). Over the years, massive use of common food safety barriers led to the development of resistance by various food microorganisms (Samson and others 1995; Holyoak and others 1996; Lin and others 1996; Park and others 1996; Henriques and others 1997; Kathariou 2002). The difficulty in the food industry in preventing outbreaks of foodborne pathogens such as Salmonella, E. coli O157:H7, Staphylococcus, L. monocytogenes, and Clostridium perfringens resulted in

thousands deaths and million cases of foodborne illnesses each year (Mead and others 1999). In addition to well-known pathogens that may cause global pandemics, new unrecognized and uncontrolled pathogens are emerging as a result of changing ecology and technology or by the transfer of mobile virulence factors such as bacteriophage, thus dramatically changing the spectrum of foodborne illnesses (Tauxe 2002). Enteric bacteria are especially tolerant to adverse environmental conditions such as low pH and high salt concentrations (Small and others 1994; Cheville and others 1996; Brown and others 1997; Mayrhofer and others 2004). Virulent strains of E. coli are increasingly recognized as foodborne pathogens. Among the 6 virotypes, enterohemorrhagic E . coli (EHEC) are considered to be highly significant due to their low infectious dose and the severe consequences of infection (Buchanan and Doyle 1997). Escherichia coli O157:H7 may cause hemorrhagic colitis and hemolytic uremic syndrome (Riley and others 1983). Outbreaks of foodborne illness due to E. coli O157:H7 have been reported from various parts of the world (Doyle 1991; Griffin and Tauxe 1991; Besser and others 1993; Anonymous 1995). Acid foods such as apple juice were reported to be associated with these outbreaks (Steele and others 1982). Therefore, contamination and proliferation of food pathogens are a great concern for food safety and public health. Consumer awareness and demands as well as food legislation have made the task of providing high-quality products even more challenging to the food industry. Consumers demand higher quality, preservative-free, safe yet mildly processed foods with an exMS 20060146 Submitted 3/2/2006, Accepted 8/26/2006. Authors are with tended shelf life. Since acidity and sterilization are the most comDept. of Biotechnology & Food Engineering, TechnionIsrael Inst. of mon preservation techniques that control outgrowth of pathogenic Technology, Haifa, 32000, Israel. Direct inquiries to author Mor (E-mail: spore-forming bacteria, addressing this consumers need calls for amor@tx.technion.ac.il). innovative approaches to ensure product preservation.

Introduction

C 2006 Institute of Food Technologists doi: 10.1111/j.1750-3841.2006.00175.x

Vol. 71, Nr. 9, 2006JOURNAL OF FOOD SCIENCE

R125

Further reproduction without permission is prohibited

R: Concise Reviews in Food Science

Promises and premises in food safety . . . Antimicrobial Peptides


ntimicrobial peptides (AMPs) are widely distributed in nature and are used by nearlyif notall life forms as essential components of nonspecific host defense systems. The list of discovered AMPs has been constantly increasing, mostly in the last 2 decades (representative AMPs are shown in Table 1). The skin of frogs (Csordas and Mich 1970) and lymph of insects (Habermann 1972) were initially shown to contain peptides that kill bacteria in culture. Since then, more than 800 AMPs were described in many living organisms: microorganisms, insects, amphibians, plants, and mammals (Boman 1995; Nicolas and Mor 1995; Hancock and Lehrer 1998; Simmaco and others 1998; Lehrer and Ganz 1999; Tossi and others 2000; Cleveland and others 2001; Mor 2001). Moreover, typical mammalian peptides may be produced in fungi as well (Mygind and others 2005). In addition to their ability to kill microorganisms directly, AMPs seem to be able to recruit and promote various elements of host immunity (Boman 1995; Gudmundsson and Agerberth 1999; Hancock and Diamond 2000; Scott and Hancock 2000). AMPs are produced both constitutively and by induction, predominantly in the animals most exposed tissues (for example, skin, eyes, and lungs), which are most likely to come in contact with microorganisms. They are rapidly synthesized at low metabolic cost, easily stored in large amounts, and have a wide spectrum as well as synergistic activity, thus providing the producing organism with a broad spectrum of coverage against a wide range of pathogens (Bals 2000). Structure-activity relationship (SAR) studies of various AMPs demonstrated that the molecular size of native peptides can be significantly reduced while maintaining antimicrobial properties and sometimes improving them (Mor and Nicolas 1994b; Oh and others 1999; Chen and others 2000; Feder and others 2000; Tossi and others 2000; Kustanovich and others 2002; Gaidukov and others 2003; Fazio and others 2005; Radzishevski and others 2005; Robinson and others 2005; Shalev and others 2005; Rydlo and others 2006). Similar optimization attempts were reported also using nonnative model peptide sequences (Blondelle and Houghten 1992; Haynie and others 1995; Park and others 2004). There are a growing number of reports using biotechnology techniques either for SAR studies or for mass production of AMPs (Kilara and Panyam 2003; Chen and others 2005; Donini and others 2005; Zho and others 2005).

R: Concise Reviews in Food Science

Biosynthesis, Structural Features, and Mechanism of Action


nimal-derived AMPs are synthesized as primary translation products, prepropeptides, which are further processed to give the active forms (Boman 1995; Nicolas and Mor 1995). In addition to proteolytic processing, posttranslation modifications include, in some cases, glycosylation (Bulet and others 1993), carboxy-terminal amidation, amino acid isomerization (Simmaco and others 1998), halogenation (Shinnar and others 1996), and cyclization (Tang and others 1999). Some AMPs can be derived from proteolysis of functional polypeptides such as skin PYY (sPYY) (Vouldoukis and others 1996), buforin II (Kim and others 2000), and lactoferricin (Ulvatne and Vorland 2001) or from proteins a priori not suspected to be involved in antimicrobial function (Rotem and others 2006). AMPs are often produced as closely related multimembered families that may vary in only a few amino acid residues (Nicolas and Mor 1995). Within each family there is a remarkably high conservation of cDNA and amino acid sequence in the prepro regions, suggesting their functional significance, although comparison of peptides from all organisms, even those that are closely related, shows practically no conservation of amino acid sequence. Despite the enormous variety of sequences and structures, AMPs possess certain common features (Boman 1995; Nicolas and Mor 1995; Andreu and Rivas 1998; Hancock and Lehrer 1998). They are usually made of less than 50 amino acids, bear a net positive charge due to an excess of basic (often lysine and/or arginine) over acidic residues, and contain about 50% hydrophobic amino acids. They often fold into 3-dimensional amphipathic structures stabilized by cysteine disulphide bridges. Linear peptides lacking cysteines tend to fold, only upon contact with membranes, into a variety of amphipathic helixes, pleated-sheets, loops, or less defined extended structures in which positively charged hydrophilic domains are well delineated from hydrophobic domains. These features seem to be the main factors affecting their known and diverse biological activities. The mechanism of action of AMPs seems to involve multiple targets. The most cited target is the plasma membrane while more recent studies suggest intracellular targets at least for some peptides (Zasloff 2002; Brogden 2005). Whether representing a final or intermediate step in the mechanism of action, it is clear that the interaction of AMPs with the plasma membrane plays an important role in

Table 1 --- Amino acid sequence of representative gene-encoded antimicrobial peptides Designation Sequencea Reference Rogers 1928 Nielsen and others 1990 Hastings and others 1991 Tichaczek and others 1994 Tomita and others 1996 Aymerich and others 1996 Cintas and others 1997 Zasloff 1987 Ge and others 1999a Shi and others 1996 Thouzeau and others 2003 Cole and others 1997 Mor and Nicolas 1994a Rydlo and others 2006 Lee and others 1997 Wilcox and Eisenberg 1992 Johansson and others 1998 Lee and others 1997 Ganzle and others 1999

AMPs produced by microbial cells Nisin ITSISLCTPGCKTGALMGCNMKTATCHCSIHVSK Pediocin PA1 KYYGNGVTCGKHSCSVDWGKATTCIINNGAMAWATGGHQGNHKC Leucocin A KYYGNGVHCTKSGCSVNWGEAFSAGVHRLANGGNGFW Sakacin P KYYGNGVHCGKhSGCTVDWGTAIGNIGNNAAANWATGGNAGWNK Bacteriocin 31 ATYYGNGLYCNKQKCWVDWNKASREIGKIIVNGWVQHGPWAPR Enterocin A TTHSGKYYGNGVYCTKNKCTVDWAKATTCIAGMSIGGFLGGAIPGQC Enterocin P ATRSYGNGVYCNNSKCWVNWGEAKENIAGIVISGWASGLAGMGH AMPs produced by animal cells Magainin GIGKFLHSAKKFGKAFVGEIMNS MSI-78 GIGKFLKKAKKFGKAFVKILKK CONH2 PR-39 RRRPRPPYLPRPRPPPFFPPRLPPRIPPGFPPRFPPRFP Spheniscin SFGLCRLRRGFCAHGRCRFPSIPIGRCSRFVQCCRRVW Pleurocidin GWGSFFKKAAHVGKHVGKAALHTYL Dermaseptin S4 ALWMTLLKKVLKAAAKALNAVLVGANA K 4 S4(1-14) ALWKTLLKKVLKAA CONH2 Cecropin P1 SWLSKTAKKLENSAKKRISEGIAIAIQGGPR Melittin GIGAVLKVLTTGLPALISWIKRKRQQ LL-37 LLGDFFRKSKEKIGKEFKRIVQRIKDFLRNLVPRTES Clavanin A VFQFLGKIIHHVGNFVHGFSHVF Curvacin A ARSYGNGVYCNNKKCNVNRGEATQSIIGGMISGWASGLAGM
a

Amino acid sequence in the 1 letter code. Positively charged residues are highlighted in bold and underlined.

R126

JOURNAL OF FOOD SCIENCEVol. 71, Nr. 9, 2006

URLs and E-mail addresses are active links at www.ift.org

their biological activity (Figure 1). Accordingly, cytotoxicity results from nonspecific interactions involving either membrane perturbations (Matsuzaki 1998; Wu and others 1999; Huang 2000) and/or cytoplasmic translocation followed by interaction with anionic elements such as nucleic acids (Friedrich and others 2000; Brogden 2005). In the case of cell-wall containing microorganisms (for example, Gram-negative bacteria), the external membranes may act as additional barriers. It has been proposed that AMPs first target the external membrane and undergo a self-promoted uptake (Hancock 1997), a process that involves displacement of divalent cations from their binding site on lipopolysaccharides (LPS) in the external membrane by competitor peptides. This leads to increased permeability and access to the cytoplasmic membrane. Hydrophobic peptides, which often aggregate in aqueous media, fail in this competition, resulting in poor bactericidal potency. Due to their nonspecific target and mode of action, the generation of resistance toward these antimicrobial agents was shown to be less likely to occur (Navon and others 2002; Yeaman and Yount 2003). Although most AMPs act by nonspecific mechanisms, they often display some selectivity between different microorganisms, for example, Gram-negative compared with Gram-positive bacteria (Boman and others 1991; Meister and others 1997), fungi preference over other eukaryotic cells (Tailor and others 1997), as well as normal compared with cancerous mammalian cells (Utsugi and others

1991; Perez-Paya and others 1994). The basis for this discrimination appears to be linked to a number of parameters, including lipid composition, membrane fluidity, extent of trans-membrane electric potential (Andreu and Rivas 1998; Matsuzaki 1999; Tossi and others 2000), and peptides self-assembly in solution (Pouny and others 1992; Ghosh and others 1997; Feder and others 2000; Kustanovitch and others 2002; Radzishevski and others 2005; Rydlo and others 2006).

Potential Therapeutic Applications


any AMPs exhibit potent activity against a wide range of microbes, including most known Gram-negative and Grampositive bacteria, yeasts, and fungi, but also against enveloped viruses such as HIV, herpes simplex virus, influenza A virus, and vesicular stomatitis virus (Zhang and Hancock 2000; Lorin and others 2005). AMPs also display activity against eukaryotic parasites, including trypanosomes, malaria parasites, and nematodes (Ghosh and others 1997; Krugliak and others 2000; Dagan and others 2002; Efron and others 2002) and a variety of types of tumor cells (Kamysz and others 2003; Papo and others 2004). Most AMPs display minimal inhibition concentration (MIC) at low micromolar range, including against antibiotic-resistant microorganisms. Many AMPs kill bacteria faster than conventional bactericidal antibiotics and are not affected by antibiotic-resistance mechanisms that often limit the use of other antibiotics. For example, no evidence of cross-resistance between magainin and antibiotics has been documented in clinical use (Ge and others 1999a). The peptides natural role(s) may thus involve synergy both with each other, as seen with dermaseptins (Mor and others 1994a), and with other agents of the host defense system. Magainin shows synergistic killing with another skin AMP termed PGLa (Matsuzaki and others 1998). Synergy has also been shown with lysozyme and various antibiotics against selected wildtype and mutant bacteria (Yan and Hancock 2001), with antifungal or antiprotozoan agents as well as with the anticancer drug doxorubicin, when tested against fungi, protozoa, and cancer cells, respectively (Scott and others 1999). AMPs are currently studied intensively because of their advantages over conventional antibiotics. They show prospects to be used in various antimicrobial applications: Microbe-derived antibiotic peptides (for example, Gramicidin S and Polymyxin B) have been used, over the years, mostly as topical creams and solutions (Hancock and Chapple 1999). Currently, several clinical trials are ongoing to assess various topical treatments by animal-derived AMPs (Hancock and Patrzykat 2002). These include phase III trials of IB367 for treating oral mucositis, most commonly associated with radiotherapy or chemotherapy for cancer; phase II clinical trials of IB-367 in aerosol formulation for Pseudomonas aeruginosa lung infections in cystic fibrosis. An indolicidin, MBI-226, is undergoing phase III clinical trials for sterilization of insertion sites for central venous catheters, and other indolicidin-like peptides are being investigated in phase II clinical trials for therapy of acute acne. Additional potential applications are also being considered, including (1) therapy of stomach ulcers due to Helicobacter pylori infections (Projan and Blackburn 1993); (2) contraceptive agents limiting the spread of sexually transmitted diseases from Neisseria, Chlamydia, human immunodeficiency virus, and Herpes simplex virus (Yasin and others 2000); (3) imaging probes for bacterial and fungal infections (Welling and others 2000; Melendez-Alafort and others 2004); (4) agents enhancing the potency of existing antibiotics, by facilitating access of antibiotics into the bacterial cell (Darveau and others 1991); (6) introduction of antimicrobial genes into plants, which by expressing the peptide become resistant to pathogens (Osusky and others 2000); (7) delivery agents for
Vol. 71, Nr. 9, 2006JOURNAL OF FOOD SCIENCE R127

Figure 1 --- Proposed antibacterial mechanism of action of linear AMPs: Unfolded cationic peptides associate with the negatively charged surface of the outer membrane (OM) and neutralize the charge over a patch of the membrane or competitively displace divalent cations from their binding sites on lipoplysacharides (LPS), creating cracks through which peptides can cross the outer membrane. In the periplasmic space (PS), the peptides adhere to negatively charged phospholipids in the cytoplasmic membrane (CM), which induces their amphipathic fold. Insertion of multiple monomers within the membrane lipid core ultimately leads to disruption of the membrane structure and function (Hancock 2001; Rydlo and others 2006).
URLs and E-mail addresses are active links at www.ift.org

R: Concise Reviews in Food Science

Promises and premises in food safety . . .

Promises and premises in food safety . . .


conjugated drugs and compounds (Feder and others 2001; Futaki trum of microorganisms. The potential of several native AMPs as and others 2001; Hariton- Gazal and others 2002); and (8) food safety. food preservatives is increasingly being reported. The main results from typical studies are summarized below according to the pepThe food-related potential applications are detailed below. tides origin.

R: Concise Reviews in Food Science

Food Preservation
onsumers concern about possible adverse health effects of some food additives has stimulated research of new effective alternatives. The bacteriocins produced by Gram-positive LAB bacteria have been found to be appropriate candidates to fulfill these requirements, mostly due to their natural origin in foods such as fermented dairy and meat products and due to their safety aspects (Cleveland and others 2001; Chen and Hoover 2003; Papagianni 2003). Nisin, a bacteriocin produced by Lactococcus lactis bacteria, was found to be of a particular interest. Discovered in 1928 (Rogers 1928), nisin was initially evaluated as a clinical antibiotic in the 1940s (Hirsch and Mattick 1949). Later, it was found to be suitable for food preservation due to its lethal activity against foodborne pathogens and spoilage microorganisms, inhibition of cell wall biosynthesis and spore outgrowth, and, most of all, its safe use and consumption with no apparent adverse effects (Hurst 1981; Montville and others 1995). So far, nisin is the only purified antimicrobial peptide approved by the U.S Food and Drug Administration for use in particular food products (FDA Federal Register 1988). It is used in a crude extract containing up to 5% of nisin of the solid. Of the different variants of nisin, the most used commercial form is NisaplinTM , which contains 2.5% of the active ingredient (Nisin A), 77.5% NaCl, and 12% nonfat dry milk (Chen and Hoover 2003). Currently it is permitted for use mostly in dairy products (especially cheese) and canned goods (vegetables, soups). In European countries it is also used in baby foods, baked goods, mayonnaise, and milk shakes. Nisin however, displays several shortcomings: low solubility at physiologic pH reduces its activity and limits its use in most cured meat products (Rayman and others 1983), and it is inactive against yeast, molds, and Gram-negative bacteria, unless other processing technologies are used in combination. These technologies include adding chelator agents like EDTA (Cutter and Siragusa 1995, Branen and Davidson 2004), preheating the product (Boziaris and others 1998), or pH reduction (Rayman and others 1983). The partial success of nisin as a food preservative has prompted examination of other bacteriocins. Pediocin PA-1 showed the most promising results (Nielsen and others 1990). But it is not yet an approved food additive in the United States. The use of bacteriocins in food preservation presents serious limitations because of their relatively narrow activity spectra and moderate antibacterial effects. Synthetic peptides have been tested in food products such as apple juice. SAR studies of synthetic model AMPs aiming at understanding their mechanism of action resulted in a 14-residue, long peptide, 8K6L, composed of 8 lysine and 6 leucine residues (Blondelle and Houghten 1992; Haynie and others 1995). The peptide showed bactericidal effect against E. coli O157, reducing its population by 6 log units after 1 h of incubation at the concentration of 50 g/mL in a buffer medium (Appendini and Hotchkiss 1999). A more recent study aimed at using this synthetic peptide as a preservative in food packaging materials demonstrated the ability of the peptide to reduce E.coli O157 population by 3.5 log units after 10-min incubation in citrate buffer (pH 3.5) at a concentration of 5 g/mL (Appendini and Hotchkiss 2000). Testing the bactericidal effect of the peptide in apple juice at 25 C (pH 3.7) revealed only 3.5 log unit reduction after a long incubation period of 8 h at the high peptide concentration of 100 g/mL. As pointed out above, unlike lactobacteria AMPs that inhibit growth of mainly Gram-positive organisms, AMPs that are produced by animal cells often display activity against a much larger specR128 JOURNAL OF FOOD SCIENCEVol. 71, Nr. 9, 2006

Mammalian AMPs
A 26 amino acid peptide derived from the sequence of PR39, a proline-arginine rich polypeptide isolated from porcine neutrophils, was found to be effective in killing E. coli, Salmonella Typhimurium, and Streptococcus suis, at 37 C (Shi and others 1996). The peptide was proposed to penetrate the outer membrane of E. coli, gain entry into the cytoplasm, and thus affect bacterial viability by interfering with DNA and/or protein synthesis (Boman and others 1993; Cabiaux and others 1994). Scanning electron microscopy studies of bacterial cells exposed to PR-26 indicated that the peptide does not lyse cells by pore-forming mechanisms (Shi and others 1996). Its effectiveness against pathogenic strains E. coli O157:H7 and L. monocytogenes was tested at different temperatures (Annamalai and others 2001). Although after 24-h incubation the peptide decreased bacterial populations by 4 and 5 log units at 24 and 37 C, respectively, it had poor activity at the lower temperatures representing normal refrigeration.

Avian AMPs
Two beta-defensins, termed spheniscins, were recently isolated from the stomach content of the king penguin Aptenodytes patagonicus (Thouzeau and others 2003). It has been proposed that in combination with other AMPs, spheniscins may be involved in a longterm preservation of food in the male birds stomach. This 38-residue AMP displayed mainly bacteriostatic effect against Gram-negative bacteria but showed bactericidal effect against most Gram-positive bacteria tested at a concentration range of 0.4 to 15 M and was active against yeast and filamentous fungi at concentrations ranging from 1.5 to >100 M. The peptide showed stable activity at a range of moderate acidity (pH 4 to 6), suggesting that it is not affected by conserving conditions in the stomach. No studies aiming to characterize these peptides in food systems as preservatives have been published to date.

Fish AMPs
Protamine, extracted from fish milt, demonstrated antimicrobial activity against a range of Gram-negative and Gram-positive bacteria, yeasts, and molds (Islam and others 1984; Uyttendaele and Debevere 1994; Johansen and others 1995). This 30 amino acid AMP (of which 66% are arginine) is believed to disrupt the cytoplasmatic membrane by inducing leakage of K+ , ATP , and intracellular enzymes of sensitive cells (Islam and others 1987; Johansen and others 1997; Stumpe and Bakker 1997). In recent studies this AMP was evaluated for its efficacy against L. monocytogenes and E. coli at pH levels ranging from 5.5 to 8 (Hansen and Gill 2000) and at temperatures ranging from 5 to 30 C (Hansen and others 2001). The peptide showed better activity against Gram-negative bacteria, showing similar MIC and MBC values, but was mainly bacteriostatic toward Gram-positive bacteria. In both cases the peptides activity increased at alkaline pH in a medium containing positively charged protein (gelatin A) where the competitive electrostatic interactions between protamine and culture media and bacterial surface were in favor for the latter. Addition of 0.9 mM EDTA was found to be either synergistic to protamine (reduced MIC) or bactericidal, depending on the bacterial strain tested. The effect of temperature on protamine MICs varied greatly among strains, species, and genera. Therefore, no definite conclusion regarding the effect of temperature on bactericidal potency could be made.
URLs and E-mail addresses are active links at www.ift.org

Another fish AMP , pleurocidin, isolated from the edible winter flounder, was active against Gram-positive and Gram-negative bacteria and was heat and salt tolerant (Cole and others 1997, 2000). Pleurocidin was active against foodborne microorganisms at levels well below the legal limit for nisin (10,000IU/g) without significant effect on human red blood cells (Burrowes and others 2004), indicating its potential as a food preservative and a natural alternative to conventional chemicals. It is noteworthy, however, that pleurocidin was inhibited by magnesium and calcium (Cole and others 2000), which may limit the use of this AMP in environments laden with these cations.

Amphibian AMPs
In addition to mammalian-like neuropeptides and hormones, the skin of amphibians also contains a rich arsenal of broad-spectrum, cytolytic AMPs (Nicolas and Mor 1995; Simmaco and others 1998). Prepro-dermaseptins are processed and then stored in the large granules of dermal glands (Nicolas and others 2003) and released onto the skin surface upon stimulation to provide an effective and fast-acting defense against noxious microorganisms (Bevins and Zasloff 1990; Simmaco and 1998). All known amphibian AMPs are linear and able to form amphipathic helical structures in hydrophobic environments (Papagianni 2003). As wide-spectrum microbicides, amphibian AMPs stimulate increasing interest because of their unique characteristics and potential therapeutic usefulness. Two of the most investigated representative frog AMPs, magainins and dermaseptins, are discussed below. Magainins discovery was triggered by the observation that wounded frogs manage to thrive in waters dense with bacteria (Zasloff 1987). Initially, the magainins were isolated from the skin of the African clawed frog Xenopus laevis, but have also been found in their stomach (Moore and others 1991). These 21 to 27 residue AMPs form an amphipathic -helix upon binding to membranes (Marion and others 1988) and displayed a broad spectrum of antimicrobial activity. At micromolar concentrations, they induced lysis of bacteria, fungi, protozoa, and tumor cells (Zasloff and others 1988; Soballe and others 1995). SAR studies yielded a 22-residue peptide termed MSI-78 showing improved potency and broader spectrum of activity, including against multiresistant bacterial species (Ge and others 1999a), and was found to be suitable for development as human therapeutic agent. A formulation based on this peptide has been taken through all phases of clinical trials for topical treatment of infected diabetic foot ulcers (Ge and others 1999b). The bactericidal effect of magainin and its synthetic analogs was investigated against 13 food-related pathogens (Abler and others 1995) displaying MICs ranging from <3 to 50 g/mL. When tested at various temperatures, Mag2 showed reduced potency at 4 and 25 C. Bovine serum albumin, which was selected as a representative food constituent that may interfere with the peptides inhibitory effect, indeed reduced its activity, suggesting that magainin peptides might be employed, during food processing as a final spray or dip for products such as meat and poultry or integrated into primary food packaging materials. The South American hylid frogs of the phyllomedusinae subfamily produce a rich array of AMPs (Vanhoye and others 2003). Several intriguing peptides have been isolated from the skin of these arboreal frogs, most of them derived from a group of precursors collectively known as preprodermorphin/dermaseptin (Amiche and others 1994; 1999; Auvynet and others 2006). Eight members of the dermaseptin S family (Mor and others 1991; Mor and Nicolas 1994a; Chen and others 2003) and 6 dermaseptin B members (Mor and others 1994b; Charpentier 1998) have been isolated and characterized. These linear AMPs often display rapid cytolytic ac-

tivity against a wide range of microorganisms, including viruses (Belaid and others 2002; Lorin and others 2005), Gram-negative and Gram-positive bacteria and filamentous fungi (Mor and others 1991; Mor and Nicolas 1994a; DeLucca and others 1998), protozoa (Hernandez and others 1992), and yeasts (Coote and others 1998). They were also found to possess intracellular antiparasitic activities (Ghosh and others 1997; Krugliac and others 2000; Dagan and others 2002; Efron and others 2002). As shown by circular dichoism, FTIR, and NMR analysis, these peptides are medium sensitive and tend to adopt helical conformations in membrane-mimicking environments (Mor and others 1991; Mor and others 1994b; Kustanovich and others 2002; Shalev and others 2002, 2005; Lequin and others 2003). Their cytolytic activity is typically mediated by interaction of the Nterminal sequence with plasma membrane phospholipids (Hernandez and others 1992; Mor and Nicolas 1994b; Gaidukov and others 2003; Radzishevski and others 2005; Shalev and others 2005). Various dermaseptin derivatives display selective activity (Mor and Nicolas 1994a; Mor and others 1994a; Feder and others 2000, 2001; Krugliac and others 2000; Navon and others 2002). Members from the dermaseptin S family exhibited dramatic synergy in combination, resulting in some cases in a 100-fold increase in activity compared to the activity of the peptides separately (Mor and others 1994a). Dermaseptins efficiently kill slowly growing and nongrowing bacteria, suggesting a potential use in the eradication of bacteria placed in a dormant state and/or subjected to low oxygen tension (Jouenne and others 1998). Dermaseptins are also effective killers of spoilage yeasts and pathogenic food-related bacteria, implying a possible application as food preservatives (Coote and others 1998; Yaron and others 2003; Rydlo and others 2006). Due to its unique molecular structure, dermaseptin S4 was used to unravel structure-function relationships that eventually led to potent derivatives (Mor and Nicolas 1994b; Feder and others 2000; Kustanovich and others 2002; Gaidukov and others 2003; Radzishevski and others 2005; Shalev and others 2005; Rydlo and others 2006). Dermaseptin S4 and it derivatives were subjected to CD and NMR studies correlating the peptides 3-D amphipathic organization with its cytotoxic profile (Mor and Nicolas 1994b; Kustanovich and others 2002; Shalev and others 2005). C-terminus truncated derivatives were investigated as model AMPs for investigating the effect of N-terminal acyl conjugation which resulted in potent and selective derivatives (Radzishevsky and others 2005; Shalev and others 2005). Binding parameter studies of the peptides to model membranes were used to investigate the mode of action and membrane permeating properties of dermaseptins. Assuming that peptide binding to the target membrane proceeds by a 2-stage mechanism (that is, an initial surface adsorption stage, due to electrostatic interaction, followed by insertion into the membrane bilayer, due to hydrophobic interaction), it was shown that the 2nd step is the crucial step leading to cell death (Gaidukov and others 2003; Shalev and others 2005; Rydlo and others 2006). Two recent studies investigated the molecular elements in dermaseptin S4 that are necessary for maintaining antimicrobial potency against E. coli O157:H7, where full length, truncated, and acylconjugated derivatives were examined for their potency, secondary structure, self-assembly state, and binding parameters (Yaron and others 2003; Rydlo and others 2006). The studies emphasize the effect of net charge and nonaggregative hydrophobicity on the peptides activity and point to 3 elements that affect peptide activity under such incubation conditions: peptide-membrane interaction, bacterial stress response, and peptide availability. The outcome of these and similar studies is discussed below in terms of how peptide performance might be affected by incubation conditions.

URLs and E-mail addresses are active links at www.ift.org

Vol. 71, Nr. 9, 2006JOURNAL OF FOOD SCIENCE

R129

R: Concise Reviews in Food Science

Promises and premises in food safety . . .

Promises and premises in food safety . . .


Effects of temperature
Only limited bactericidal activity of AMPs was reported at low temperatures (4 and 20 C) when examining the inactivation of foodborne pathogens with magainin analogs (Alber and others 1995). Low incubation temperatures may affect the peptides solubility, leading to formation of activity-limiting aggregates. Maisnier-Patin and others (1996) found that the antimicrobial effect of the bacteriocin EFS2 on L. innouua was highest at 35 C, with activity loss at 15 C due to low solubility. But this may not be the case for other peptides: Annamalali and others (2001) reported that activity of PR-26 against E. coli O157:H7 and L. monocytogenes was reduced at 4 and 10 C compared to 25 and 37 C. Similarly, Mendoza and others (1999) reported that as temperature is decreased (from 37 to 6 C) the bactericidal effect of enterocin AS-48, which is produced by Enterococcus faecalis and tested on L. monocytogenes, decreased substantially. Both of the above-mentioned peptides were highly soluble at low temperatures, indicating that low temperatures may alter susceptibility to antimicrobial agents rather by affecting fluidity and/or fatty acid profile of bacterial cell membrane. For other peptides, lower temperatures have an opposite effect due to structure stabilization. The C-terminal domain of class IIa bacteriocins, a region involved in receptor recognition, was used to study the effects of temperature on structure and antimicrobial activity (Kaur and others 2004). Peptides that did not possess a 2nd C-terminal disulfide bond in addition to the N-terminal disulfide bond were found to experience partial disruption of the helical section at elevated temperatures (37 C) and were 30- to 50-fold less potent. In a study on dermaseptins (Rydlo and others 2006), low temperatures of 25 C and 4 C were found to limit the bactericidal activity of dermaseptin S4 derivatives (Figure 2). The difference in potency could not be linked to obvious changes in peptides secondary structure. Moreover, the peptide exhibited the most ordered structure at 25 C although it was most active at 42 C. A slight decrease in the bacterial population at 4 C in the control group (that was not exposed to the peptides) suggests that bacterial resistance increased due to stress conditions.

R: Concise Reviews in Food Science

Effects of salt
Electrostatic interactions between peptide and negatively charged targets maybe masked at high salt concentrations. Salt might also affect peptide organization in solution in terms of either secondary structure, salting out or induce aggregation, thus limiting its activity. The -helical content of a synthetic bactericidal peptide [RLLR] 5 was reduced from 72% to 13% in the presence of 200 mM NaCl, leading to an 8- to 32-fold decrease in its antimicrobial activity against bacteria and fungi (Park and others 2004). This study also indicated that conjugation of -helix-capping motives at the peptides termini helped maintain its helical structure under salty conditions, resulting in improved activity. In different studies, however, salt supplement actually improved bactericidal activity: Adding up to 300 mM NaCl did not impair cecropin P1 activity against E. coli but progressively led to activity loss for magainin (Lee and others 1997). Activity of nisin against E.coli O157 was achieved at high salt concentrations of 5% irrespective of whether bacteria were precultured in the presence or absence of salt in the incubation medium (Ganzle and others 1999). Salt resistance to -helical cationic AMPs was also reported against P . aeruginosa PAO1 at salt concentrations of 300 mM (Freidrich and others 1999). Unfortunately, none of these studies addressed changes in secondary structure in the presence of high salt concentrations. The bee venom melittin was reported to convert from a monomeric random coil to a helical tetramer as ionic strength was increased from 0 to 500 mM NaCl (Wilcox and Eisenberg 1992). Furthermore, 2 -helix forming peptides produced by gene engineering methods were able to maintain the secondary structure at extremely high salt concentrations up to 1.5 M (Kojima and others 1996). High salt concentration was explained to mask electrostatic repulsion between similarly charged side chains of lysine (in acidic conditions) or glutamic acid (in alkali conditions) on the hydrophilic

Figure 2 --- Effects of incubation conditions on the antimicrobial properties of dermaseptin derivatives on E. coli 0157:H7. Plotted in column A are data obtained after 2 h of incubation with a 28 residue dermaseptin. Columns B and C present data obtained with a 14-residue truncated derivative and its acylated version, respectively, at same peptide concentration (8 M). CFU values represent the mean standard deviations (SD) obtained from two independent experiments performed in duplicates. Bars represent the SD from the mean. Absence of a bar indicates that the SD value is smaller than the symbol size. Zero CFU indicate negative cultures (Rydlo and others 2006). R130 JOURNAL OF FOOD SCIENCEVol. 71, Nr. 9, 2006
URLs and E-mail addresses are active links at www.ift.org

surface of the helix bundle, thus stabilizing the peptide structure. In addition, hydrophobic interaction was strengthened when increasing salt concentration, thus facilitating peptide association. Another study showed that the human cathelicidin, LL-37, converted from random coil in water to -helix in the presence of various salts, including 160 mM NaCl, and that the -helix content correlated the observed antibacterial activity (Johansson and others 1998). The salting out effect explained in this study causes peptide oligomerization observed with increasing helical content. Reducing repulsive forces between positively charged residues along the peptide chain was also explained as an outcome of salt addition, resulting in stabilization of the helix structure. Changes in antibacterial activity of dermaseptin derivatives did not correlate with structural changes as function of salt concentration since the peptides exhibited the highest helical structure at 6% NaCl where activity was significantly reduced (Rydlo and others 2006). Acylation remarkably improved efficacy, enabling the peptide to eliminate large bacterial populations (107 CFU) at high salt concentrations of 1.5% (approximately 0.3 M NaCl) as shown in Figure 2. At 3%, on the other hand, the peptide did not display improved activity. At such high salt concentrations, growth of the control group was also hampered.

tion (Richard and Foster 2004). The pH effect is bound to influence the peptides interaction with bacterial membrane components in a variety of manners. In support of this view, surface plasmon resonance (SPR) experiments performed with dermaseptin derivatives revealed high binding affinity at low pH but a low tendency to penetrate into the membrane inner core (Rydlo and others 2006). A solidstate NMR spectroscopy study also reported histidine-rich peptides to alter their orientation in a model membrane as a function of the surrounding pH, changing from parallel to transmembrane orientations, respectively, at acidic and elevated pH conditions (Bechinger 1996). The pH differences also affected the mode of action by which clavanin A permeabilized Lactobacillus sake membrane (van Kan and others 2002). Whereas the peptide efficiently released fluorophores from unilamellar vesicles at neutral pH according to a nonspecific permeabilization mechanism, it did not permeabilize model bilayers at low pH levels although displaying antimicrobial activity at those conditions. It was therefore suggested that this peptide uses distinct modes of action at acidic and neutral pH.

Bactericidal Activity in Food Model Systems


variety of studies attempted to define the efficacy of several peptides in food model systems such as apple juice, milk, and meat products: Activity of magainin, pediocin PA-1, and sakacin A was decreased in the presence of bovine serum albumin or foods of high protein content (Muriana 1993; Abler and others 1995). Activity of dermaseptin S4 derivatives in apple juice was limited by 2-fold for certain derivatives (Yaron and others 2003) and up to 8-fold for other derivatives (Rydlo and others 2006) compared with standard incubation conditions. After preincubation in apple juice, the peptide was assayed for its activity in standard LB medium showing reduced activity compared to nonpreincubated peptide. Thus, peptide inactivation also depends on juice component(s). Activity of nisin against L. monocytogenes in fluid milk decreased as the milk fat content increased (Jung and others 1992). Therefore, it has been speculated that components such as lipids, proteins, and sugars in foods interact with the peptide and hamper its activity.

Effects of pH
Variations in pH can affect the peptides net charge, which in turn affects its secondary structure, binding properties, and cytotoxic activity. Histidine-containing amphipathic helical peptides were reported to exhibit random coil structure at acidic pH levels and typical -helix structure at basic conditions (Vogt and Bechinger 1999). Another histidine-based peptide, clavanin A, showed enhanced activity at acidic conditions while at neutral pH, the uncharged peptide became inactive. Its substituted analog, clavanin AK, having a higher pKa value, exhibited potency at all conditions (Lee and others 1997). The human cathelicidin LL-37 displayed random coil structure at acidic conditions (pH 2 and 3.8) and -helix at neutral and basic pH levels (Johansson and others 1998). The acidic conditions reportedly destabilized ion pairs between acidic and basic side chains due to the protonated state of acidic residues, whereas at basic conditions, repulsive forces between basic side chains were limited due to deprotonation of positively charged residues, leading to helix stabilization. Thus, a peptides pKa value can determine loss or gain of activity as a function of the environmental pH conditions. Dermaseptin S4 derivatives also displayed reduced bactericidal potency against E. coli O157:H7 at acidic conditions of pH 3.6 (Figure 2). A few possible reasons might account for differences in peptide efficacy: Bacterial stress responses may reduce bacterial susceptibility to AMPs either by modifications of the membrane composition, SOS gene expression, or modification of the trans-membrane potential. Several studies revealed that rpoS, a gene involved in regulating the expression of a variety of stress response genes, could be induced in response to a number of inimical processes used in the food industry, including changes in water activity, osmotic shock, and highly acidic conditions (Foster and Spector 1995; Fang and others 1996; Turner and others 2000; Dodd and Aldsworth 2002). Preculture of E. coli in acidic conditions led to resistance to curvacin A (Ganzle and others 1999). A similar outcome was obtained with dermaseptins where bacteria that had been pre-incubated in pH 3.6 and assayed in neutral pH became less susceptible (Rydlo and others 2006). However, the fact that bacterial viability was hampered in absence of peptide may point to stress response adaptation of bacteria being the cause for reduced potency. As part of the stress response, it has been reported that E. coli transmembrane potential decreases significantly at acidic condiURLs and E-mail addresses are active links at www.ift.org

Food Packaging
ntimicrobial packaging has become one of the most interesting and challenging topics in the area of active packaging (AP). In AP , the product, the package, and the environment interact during food preparation and storage, resulting either in an improved product quality and safety and an extended shelf life or in the attainment of some product characteristics that cannot be obtained by any other means (Miltz and others 1995; Yam and others 2005). The most developed AP technology is oxygen scavenging. A reduction of oxygen in a package can inhibit oxidative reactions as well as the growth of microorganisms. However, a reduction in the oxygen concentration in a package to very low levels may encourage the growth of pathogenic anaerobic microorganisms. Antimicrobial packaging overcomes this problem. Several antimicrobial packaging systems have been proposed and reviewed in detail (Appendini and Hotchkiss 2002) where most of them use synthetic additives. The subject of antimicrobial packaging has been reviewed recently in 2 additional articles by Suppakul and others (2003a, 2003b). In recent years, however, the public perception is that synthetic agents may cause side effects and therefore consumers prefer natural over synthetic additives. Miltz and others (2004) have developed an antimicrobial film containing the natural components of basil: linalool and methyl chavicol. This film has shown a significant inhibition of E. coli and Listeria as well as other microorganisms and extended the shelf life of cheddar cheese. Miltz and others (2006) have studied very recently the properties of an antimicrobial coating based
Vol. 71, Nr. 9, 2006JOURNAL OF FOOD SCIENCE R131

R: Concise Reviews in Food Science

Promises and premises in food safety . . .

Promises and premises in food safety . . .


on a dermaseptin AMP . This coating has shown significant microbial growth inhibition.
Blondelle SE, Houghten RA. 1992. Design of model amphipathic peptides having potent antimicrobial activities. Biochemistry 31(50):1268894. Boman HG. 1995. Peptide antibiotics and their role in innate immunity. Annu Rev Immunol 13:6192. Boman HG, Faye I, Gudmundsson GH, Lee JY, Lidholm DA. 1991. Cell-free immunity in Cecropia. A model system for antibacterial proteins. Eur J Biochem 201:2331. Boman HG, Agerberth B, Boman A. 1993. Mechanisms of action on Escherichia coli of cecropin P1 and PR-39, two antibacterial peptides from pig intestine. Infect Immun 61:297884. Boziaris IS, Humpheso L, Adams MR. 1998. Effect of nisin on heat injury and inactivation of Salmonella enteritidis PT4. Int J Food Microbiol 43:713. Branen JK, Davidson PM. 2004. Enhancement of nisin, lysozyme, and monolaurin antimicrobial activities by ethylenediaminetetraacetic acid and lactoferrin. Int J Food Microbiol 90:6374. Brogden KA. March 2005. Antimicrobial peptides: pore formers or metabolic inhibitors in bacteria? Nature Revi Microbiol 3:238-50. Available from: www.nature. com/reviews/micro. Brown JL, Ross T, McMeekin TA, Nichols PD. 1997. Acid habituation of Escherichia coli and potential role of cyclopropane fatty acids in low pH tolerance. Int J Food Microbiol 37:16373. Buchanan RL, Doyle MP. 1997. Foodborne disease significance of Escherichia coli O157:H7 and other enterohemorrhagic E. coli. Food Technol 51(10):6976. Bulet P, Dimarcq JL, Hetru C, Lagueux M, Charlet M, Hegy G, Dorsselaer AV, Hoffmann JA. 1993. A novel inducible antibacterial peptide of Drosophila carries an O-glycosylated substitution. J Biol Chem 268:148937. Burrowes OJ, Hadjicharalambous C, Diamond G, Lee TC. 2004. Evaluation of antimicrobial spectrum and cytotoxic activity of pleurocidin for food applications. J Food Sci 69(3):6671. Cabiaux V, Agerberth B, Johansson J, Homble F, Goormaghtigh E, Ruysschaert M. 1994. Secondary structure and membrane interaction of PR-39, a Pro+Arg-rich antibacterial peptide. Eur J Biochem 224:101927. Charpentier S, Amiche M, Mester J, Vouille V, Le Caer JP, Nicolas P , and Delfour D. 1998. Structure, synthesis and molecular cloning of dermaseptin B, family of skin peptide antibiotics. J Biol Chem 273:146907. Chen H, Hoover DG. 2003. Bacteriocins and their food applications. Comp Rev Food Sci Safety 2:82100. Chen H, Xu Z, Xu N, Cen P. 2005. Efficient production of a soluble fusion protein containing human beta-defensin-2 in E. coli cell-free system. J Biotechnol 115(3):307 15. Chen J, Falla TJ, Liu H, Hurst MA, Fujii CA, Mosca DA, Embree JR, Loury DJ, Radel PA, Cheng Chang C, Gu L, Fiddes JC. 2000. Development of protegrins for the treatment and prevention of oral mucositis: structure-activity relationships of synthetic protegrin analogues. Biopolymers 55(1):8898. Chen T, Tang L, Shaw C. 2003. Identification of three novel Phyllomedusa sauvagei dermaseptins (sVI-sVIII) by cloning from a skin secretion-derived cDNA library. Regul Pept 116(13):13946. Cheville AM, Arnold KW, Buchrieser C, Cheng CM, Kaspar CW. 1996. RpoS regulation of acid, heat, and salt tolerance in Escherichia coli O157:H7. Appl Environ Microbiol 62:18224. Cintas LM, Casaus P, Havarstein LS, Hernandez PE, Nes IF. 1997. Biochemical and genetic characterization of enterocin P , a novel secdependent bacteriocin from Enterococcus faecium P13 with a broad antimicrobial spectrum. Appl Environ Microbiol 63:432130. Cleveland J, Montville TJ, Nes IF, Chikindas ML. 2001. Bacteriocins: safe, natural antimicrobials for food preservation. Int J Food Microbiol 71:120. Cole A, Weis P, Diamond G. 1997. Isolation and characterization of pleurocidin, an antimicrobial peptide in the skin secretions of winter flounder. J Biol Chem 272(18):120813. Cole A, Darouiche R, Legarda D, Connell N, Diamond G. 2000. Characterization of a fish antimicrobial peptide: gene expression, subcellular localization and spectrum of activity. Antimicrob Agents Chemother 44:203945. Coote PJ, Holyoak CD, Bracey D, Ferdinando DP, Pearce JA. 1998. Inhibitory action of a truncated derivative of the amphibian skin peptide dermaseptin s3 on Saccharomyces cerevisiae. Antimicrob Agents Chemother 42:216070. Csordas A, Michl H. 1970. Isolation and structural resolution of a haemolytically active polypeptide from the immune secretion of a European toad. Monatsh Chem 101:1829. Cutter CN, Siragusa G. 1995. Population reductions of Gram negative pathogens following treatments with nisin and chelators under various conditions. J Food Prot 58:97783. Dagan A, Efron L, Gaidukov L, Mor A, Ginsburg H. 2002. In vitro antiplasmodium effects of dermaseptin S4 derivatives. Antimicrob Agents Chemother 46:105966. Darveau RP, Cunningham MD, Seachord CL, Cassiano-Clough L, Cosand WL, Blake J, Watkins CS. 1991. Beta-lactam antibiotics potentiate magainin 2 antimicrobial activity in vitro and in vivo. Antimicrob Agents Chemother 35:11539. DeLucca AJ, Bland JM, Jack TJ, Grimm C, Walsh YJ. 1998. Fungicidal and binding properties of the natural peptides cecropin B and dermaseptin. Med Mycol 36:291 8. Dodd CE, Aldsworth TG. 2002. The importance of RpoS in the survival of bacteria through food processing. Int J Food Microbiol 74:18994. Donini M, Lico C, Baschieri S, Conti S, Magliani W, Polonelli L, Benvenuto E. 2005. Production of an engineered killer peptide in Nicotiana benthamiana by using a potato virus X expression system. Appl Environ Microbiol 71(10):63607. Doyle MP. 1991. Escherichia coli O157:H7 and its significance in foods. Int J Food Microbiol 12:289301. Efron L, Dagan A, Gaidukov L, Ginsburg H, Mor A. 2002. Direct interaction of dermaseptin S4 aminoheptanoyl derivative with intraerythrocytic malaria parasite leading to increased specific antiparasitic activity in culture. J Biol Chem 277:24067 72. Fang FC, Chen CY, Guiney DG, Xu Y. 1996. Identification of sigma S-regulated genes in Salmonella typhimurium: complementary regulatory interactions between sigma S and cyclic AMP receptor protein. J Bacteriol 178:511220.
URLs and E-mail addresses are active links at www.ift.org

R: Concise Reviews in Food Science

Concluding Remarks

hile the data collected in this review highlight part of the massive research that is currently being conducted, many signs designate the potential of natural AMPs in food preservation. Future challenges lie in our ability to adapt these extraordinary compounds to perform tasks specifically related to food safety. Better understanding of the mode(s) by which AMPs rapidly eliminate microorganisms should provide solid grounds for engineering new and upgraded derivatives with optimized potency and stability under the range of incubation conditions typical to food. This endeavor is indeed stimulating an increasing number of multidisciplinary studies that tackle these issues by a large variety of strategies: So far, success in significantly reducing the molecular size while increasing resistance to proteases was achieved through SAR studies aiming at optimizing known native peptides or model peptide sequences through conventional and/or combinatorial methods. In the future, the use of biotechnology techniques should further enable mass production of AMPs to minimize their costs. The recent isolation of plectasin, a defensin-like AMP expressed in fungi, clearly represents a breakthrough toward achieving several of these goals. There is therefore little doubt that as progress is made, new derivatives will be designed which will enhance the suitability of these simple yet phenomenal molecules for various antimicrobial applications, including food safety, for their demonstrable ability to escape most of the known resistance mechanisms.

Acknowledgments
The authors gratefully acknowledge the financial support by THE ISRAEL SCIENCE FOUNDATION (grant nr 387/03) and by funds from the Technions Vice President for Research.

References
Abler LA, Klapes NA, Sheldon BW, Klaenhammer TR. 1995. Inactivation of food-borne pathogens with magainin peptides. J Food Prot 58:3818. Amiche M, Ducancel F, Mor A, Boulain JC, Menez A, Nicolas P. 1994. Precursors of vertebrate peptide antibiotics dermaseptin B and adenoregulin have extensive sequence identities with precursors of opioid peptides dermorphin, dermenkephalin, and deltorphins. J Biol Chem 269:1784752. Amiche M, Seon AA, Pierre T, Nicolasm P. 1999. The dermaseptin precursors: a protein family with a common preproregion and a variable C-terminal antimicrobial domain. FEBS Lett 456:3526. Andreu D, Rivas L. 1998. Animal antimicrobial peptides: an overview. Biopolymers 47:41533. Annamalai T, Venkitanarayanan KS, Hoagland TA, Khan MI. 2001. Inactivation of Escherichia coli O157: H7 and Listeria monocytogenes by PR-26, a synthetic antibacterial peptide. J Food Prot 64:192934. Anonymous. 1995. Escherichia coli O157:H7 outbreak linked to commercially distributed dry-cured salamiWashington and California. Morb Mortal Wkly Rep 44:15760. Appendini P, Hotchkiss JH. 1999. Antimicrobial activity of a 14-residue peptide against Escherichia coli O157:H7. J Appl Microbiol 87(5):7506. Appendini P, Hotchkiss JH. 2000. Antimicrobial activity of a 14-residue synthetic peptide against foodborne microorganisms. J Food Prot 63:88993. Appendini P, Hotchkiss JH. 2002. Review of antimicrobial food packaging. Inno Food Sci Emerging Technol 3:11326. Auvynet C, Seddiki N, Dunia I, Nicolas P, Amiche M, Lacombe C. 2006. Posttranslational amino acid racemization in the frog skin peptide deltorphin I in the secretion granules of cutaneous serous glands. Eur J Cell Biol 85(1):2534. Aymerich T, Holo H, Havarstein LS, Hugas M, Garriga M, Nes IF. 1996. Biochemical and genetic characterization of enterocin A from Enterococcus faecium, a new antilisterial bacteriocin in the pediocin family of bacteriocins. Appl Environ Microbiol 62:167682. Bals R. 2000. Epithelial antimicrobial peptides in host defense against infection. Respir Res 3:14150. Bechinger B. 1996. Towards membrane protein design: pH-sensitive topology of histidine-containing polypeptides. J Mol Biol 263:76875. Belaid A, Aounim M, Khelifa R, Trabelsi A, Jemmali M, Hani K. 2002. In vitro antiviral activity of dermaseptins against herpes simplex virus type 1. J Med Virol 66:22934. Besser RE, Lett SM, Weber JT, Doyle MP, Barrett TJ, Wells JG, Grin PM. 1993. An outbreak of diarrhea and hemolytic uremic syndrome from Escherichia coli O157:H7 in freshpressed apple cider. JAMA 269:221720. Bevins CL, Zasloff M. 1990. Peptides from frog skin. Annu Rev Biochem 59:395414.

R132

JOURNAL OF FOOD SCIENCEVol. 71, Nr. 9, 2006

Fazio MA, Oliveira VX Jr, Bulet P, Miranda MT, Daffre S, Miranda A. 2005. Structureactivity relationship studies of gomesin: importance of the disulfide bridges for conformation, bioactivities and serum stability. Biopolymer 84(2): 20518. FDA (U.S. Food and Drug Administration)/Federal Register. 1988. 21 CFR Part 184. Fed Reg 53:1124751. Feder R, Dagan A, Mor A. 2000. Structure-activity relationship study of antimicrobial dermaseptin S4 showing the consequences of peptide oligomerization on selective cytotoxicity. J Biol Chem 275:42308. Feder R, Nehushtani R, Mor A. 2001. Affinity driven molecular transfer from erythrocyte membrane to target cells. Peptides 22:168390. Foster JW, Spector MP. 1995. How Salmonella survive against the odds. Annu Rev Microbiol 49:14574. Friedrich C, Scott MG, Karunaratne N, Yan H, Hancock REW. 1999. Salt-resistant alphahelical cationic antimicrobial peptides. Antimicrob Agents Chemother 43:15428. Friedrich CL, Moyles D, Beveridge TJ, Hancock REW. 2000. Antibacterial action of structurally diverse cationic peptides on Gram-positive bacteria. Antimicrob Agents Chemother 44:208692. Futaki S, Suzuki T, Ohashi T, Yagami T, Tanaka S, Ueda K, Sugiura Y. 2001. Arginine-rich peptides. An abundant source of membrane permeable peptides having potential as carriers for intracellular protein delivery. J Biol Chem 276:583640. Gaidukov L, Fish A, Mor A. 2003. Analysis of membrane-binding properties of dermaseptin analogues: relationships between binding and cytotoxicity. Biochem 41:1286674. Ganzle MG, Hertel C, Hammes WP. 1999. Resistance of Escherichia coli and Salmonella against nisin and curvacin A. Int J Food Microbiol 48:3750. Ge Y, MacDonald DL, Halroyd KJ, Thornsberry C, Wexler H, Zasloff M. 1999a. In vitro antibacterial properties of pexiganan. Antimicrob Agents Chemother 43:7828. Ge Y, MacDonald DL, Henry MM, Hait HI, Nelson KA, Lipsky BA, Zasloff M. Holroyd KJ. 1999b. In vitro susceptibility to pexiganan of bacteria isolated from infected diabetic foot ulcers. Diagn Micrbiol Infect Dis 35:4553. Ghosh JK, Shaool D, Guillaud P, Ciceron L, Mazier D, Kustanovich I, Shay Y, Mor A. 1997. Selective cytotoxicity of dermaseptin S3 towards intraerythrocytic Plasmodium falciparum and the underlying molecular basis. J Biol Chem 267:65029. Gold HS, Moellering RC. 1996. Antimicrobial-drug resistance. N Engl J Med 335:1445 53. Griffin PM, Tauxe RV. 1991. The epidemiology of infections caused by Escherichia coli O157:H7, other enterohemorrhagic E. coli, and the associated hemolytic uremic syndrome. Epidemiol Rev 13:6098. Gudmundsson GH, Agerberth B. 1999. Neutrophil antibacterial peptides, multifunctional effector molecules in the mammalian immune system. J Immunol Methods 232:4554. Habermann E. 1972. Bee and wasp venoms. Science 177:31422. Hancock REW. 1997. Peptide antibiotics. Lancet 349:41822. Hancock REW. 2001. Cationic peptides: effectors in innate immunity and novel antimicrobials. Lancet Infectious Disease. 1:15664. Hancock REW, Lehrer RI. 1998. Cationic peptides: a new source of antibiotics. Trends Biotechnol 16:828. Hancock REW, Chapple DS. 1999. Peptide antibiotics. Antimicrob Agents Chemother 43:1317. Hancock REW, Diamond G. 2000. The role of cationic antimicrobial peptides in innate host defences. Trends Microbiol 8:40210. Hancock REW, Patrzykat A. 2002. Clinical development of cationic antimicrobial peptides: from natural to novel antibiotics. Curr Drug Targets Infect Disord 2:7983. Hansen LT, Gill TA. 2000. Solubility and antimicrobial efficacy of protamine on Listeria monocytogenes and Escherichia coli as influenced by pH. J Appl Microbiol 88:1049 55. Hansen LT, Austin JW, Gill TA. 2001. Antibacterial effect of protamine in combination with EDTA and refrigeration. Int J Food Microbiol 66:14961. Hariton-Gazal EH, Feder R, Mor A, Graessmann A, Werner RB, Jans D, Gilon C, Loyter A. 2002. Targeting of nonkaryophilic cell-permeable peptides into the nuclei of intact cells by covalently attached nuclear localization signals. Biochemistry 41(29):9208 14. Hastings JW, Sailer M, Johnson K, Roy KL, Vederas JC, Stiles ME. 1991. Characterization of leucocin A-UAL 187 and cloning of the bacteriocin gene from Leuconostoc gelidum. J Bacteriol 173:7491500. Haynie SL, Crum GA, Doele BA. 1995. Antimicrobial activities of amphiphilic peptides covalently bonded to a water-insoluble resin. Antimicrob Agents Chemother 39: 3017. Henriques M, Quintas C, Loureiro-Dias MC. 1997. Extrusion of benzoic acid in Saccharomyces cerevisiae by energy-dependent mechanism. Microbiology 143:187783. Hernandez C, Mor A, Dagger F, Nicolas P, Hernandez A, Benedetti EL, Dunia I. 1992. Functional and structural damage in Leishmania mexicana exposed to the cationic peptide dermaseptin. Eur J Cell Biol 59:41424. Hirsch A, Mattick ATR. 1949. Some recent applications of nisin. Lancet (2):1903. Holyoak CD, Stratford M, McMullin Z, Cole MB, Crim FK, Brown AJP, Coote P. 1996. Activity of the plasma membrane H(+)-ATPase and optimal glycolytic flux are required for rapid adaptation and growth of Saccharomyces cerevisiae in the presence of the weak-acid preservative sorbic acid. Appl Environ Microbiol 62:315864. Huang HW. 2000. Action of antimicrobial peptides: 2-state model. Biochem 39:8347 52. Hurst A. 1981. Nisin. Adv Appl Microbiol 27:85123. Islam NMD, Itakura T, Motohiro T. 1984. Antibacterial spectra and minimum inhibition concentration of clupeine and salmine. Bull Jpn Soc Sci Fish 50:17058. Islam NMD, Oda H, Motohiro T. 1987. Changes in the cell morphology and the release of soluble constituents from washed cells of Bacillus subtilis by the action of protamine. Nippon Suisan Gakkaishi 53:297303. Johansen C, Gill T, Gram L. 1995. Antibacterial effect of protamine assayed by impedimetry. J App Bacteriolog 78:297303. Johansen C, Verheul A, Gram L, Gill T, Abee T. 1997. Protamine-induced permeabilization of cell envelopes of Gram-positive and Gram-negative bacteria. Apple Enviro Microbiol 63:11559. Johansson J, Gudmundsson GH, Rottenberg ME, Berndt KD, Agerberth B. 1998.
URLs and E-mail addresses are active links at www.ift.org

Conformation-dependent antibacterial activity of the naturally occurring human peptide LL-37. J Biol Chem 273:371824. Jouenne T, Mor A, Bonato H, Junter GA. 1998. Antibacterial activity of synthetic dermaseptins against growing and non-growing Escherichia coli cultures. J Antimicrob Chemother 42:8790. Jung D, Bodyfelt FW, Daeschel MA. 1992. Influence of fat and emulsifiers on the efficacy of nisin in inhibiting Listeria monocytogenes in fluid milk. J Dairy Sci 75:38793. Kamysz W, Okroj M, Lukasiak J. 2003. Novel properties of antimicrobial peptides. Acta Biochimica Polonica 50(2):4619. Kathariou S. 2002. Listeria monocytogenes virulence and pathogenicity, a food safety perspective. J Food Prot 65:181129. Kaur K, Andrew LC, Wishart DS, Vederas JC. 2004. Dynamic relationships among type IIa bacteriocins: temperature effects on antimicrobial activity and on structure of the C-terminal amphipathic alpha helix as a receptor-binding region. Biochem 43:9009 20. Kilara A, Panyam D. 2003. Peptides from milk proteins and their properties. Crit Rev Food Sci Nutr 43(6):60733. Kim HS, Yoon H, Minn I, Park CB, Lee WT, Zasloff M, Kim SH. 2000. Pepsin-mediated processing of the cytoplasmic histone H2A to strong antimicrobial peptide buforin I. J Immunol 165:326874. Kojima S, Kuriki Y, Sato Y, Arisaka F, Kumagai I, Takahashi S, Miura K. 1996. Synthesis of alpha-helix-forming peptides by gene engineering methods and their characterization by circular dichroism spectra measurements. Biochim Biophys Acta 1294:12937. Krugliak M, Feder R, Zolotarev VY, Gaidukov L, Dagan A, Ginsburg H, Mor A. 2000. Antimalarial activities of dermaseptin S4 derivatives. Antimicrob Agents Chemother 44:244251. Kustanovich I, Shalev DE, Mikhlin M, Gaidukov L, Mor A. 2002. Structural requirements for potent versus selective cytotoxicity for antimicrobial dermaseptin S4 derivatives. J Biol Chem 277(19):1694151. Lee IH, Cho Y, Lehrer R. 1997. Effects of pH and salinity on the antimicrobial properties of clavanins. Infec Immun 65:2898903. Lehrer RI, Ganz T. 1999. Antimicrobial peptides in mammalian and insect host defence. Curr Opin Immunol 11:237. Lequin O, Bruston F, Convert O, Chassaing G, Nicolas P. 2003. Helical structure of dermaseptin B2 in a membrane-mimetic environment. Biochemistry 42:1031123. Lin JS, Smith MP, Chapin KC, Baik HS, Bennett GN, Foster JW. 1996. Mechanisms of acid resistance in enterohemorrhagic Escherichia coli. Appl Environ Microbiol 62:3094 100. Lorin C, Saidi H, Belaid A, Zairi A, Baleux F, Hocini H, Belec L, Hani K, Tangy F. 2005. The antimicrobial peptide dermaseptin S4 inhibits HIV-1 infectivity in vitro. Virology 334:26475. Maisnier-Patin S, Forni E, Richard J. 1996. Purification, partial characterisation and mode of action of enterococcin EFS2, an antilisterial bacteriocin produced by a strain of Enterococcus faecalis isolated from a cheese. Int J Food Microbiol 30:255 70. Marion D, Zasloff M, Bax M. 1988. A two-dimensional NMR study of the antimicrobial peptide magainin 2. FEBS Lett 227(1):216 Matsuzaki K. 1998. Magainins as paradigm for the mode of action of pore forming polypeptides. Biochim Biophys Acta 1376:391400. Matsuzaki K. 1999. Why and how are peptide-lipid interactions utilized for selfdefense? Magainins and tachyplesins as archetypes. Biochim Biophys Acta 1426:1 10. Matsuzaki K, Mitani Y, Akada KY, Murase O, Yoneyama S, Zasloff M, Miyajima K. 1998. Mechanism of synergism between antimicrobial peptides magainin 2 and PGLa. Biochemistry 37:1514453. Mayrhofer M, Paulsen P, Smulders FJM, Hilbert F. 2004. Antimicrobial resistance profile of five major food-borne pathogens isolated from beef, pork and poultry. Int J Food Microbiol 97(1):239. Mead PS, Slutsker L, Dietz V, McCaig LF, Bresee JF, Shapiro C, Griffin PM, Tauxe RV. 1999. Food-related illness and death in the United States. Emerging Infect Dis 5(5):60725. Meister M, Lemaitre B, Hoffmann JA. 1997. Antimicrobial peptide defense in Drosophila. Bioessays 19:101926. Melendez-Alafort L, Rodriguez-Cortes J, Ferro-Flores G, Arteaga De Murphy C, HerreraRodriguez R, Mitsoura E, Martinez-Duncker C. 2004. Biokinetics of (99m)Tc-UBI 29-41 in humans. Nucl Med Biol 31(3):3739. Mendoza F, Maqueda M, Galvez A, Martinez-Bueno M, Valdivia E. 1999. Antilisterial activity of peptide AS-48 and study of changes induced in the cell envelope properties of an AS-48-adapted strain of Listeria monocytogenes. Appl Environ Microbiol 65:61825. Miltz J, Passy N, Mannheim CH. 1995. Trends and applications of active packaging systems. In: Ackerman P, Jagerstad M, Ohlsson M, editors. Food and packaging materialschemical interaction. The Royal Soc. of Chemistry Publ. No 162, p 201 10. Miltz J, Bigger SW, Sonneveld C, Suppakul P. 2004. Antimicrobial Packaging Material. Australian Patent Application No. 2004903510. Miltz J, Rydlo T, Mor A, Polyakov V. 2006. Potency evaluation of a dermaseptin s4 derivative for antimicrobial food packaging applications. Packag Technol Sci. Online May 24. http://www3.interscience.wiley.com/cgi-bin/abstract/112636230/ ABSTRACT?CRETRY=1&SRETRY=0 Miltz J, Rydlo T, Mor A, Polyakov V. 24 May 2006. Potency evaluation of a dermaseptin s4 derivative for antimicrobial food packaging applications. Packag Technol SiPublished Online. Montville TJ, Winkowski K, Ludescher RD. 1995. Models and mechanisms for bacteriocin action and application. Int Dairy J 5:797814. Moore KS, Bevins C, Brasseur MM, Tomassini N, Turner K, Eck H, Zasloff M. 1991. Antimicrobial peptides in the stomach of Xenopus laevis. J Biol Chem 266:198517. Mor A, Nguyen VH, Delfour A, Migliore SD, Nicolas P. 1991. Isolation, amino acid sequence, and synthesis of dermaseptin, a novel antimicrobial peptide of amphibian skin. Biochemistry 3:882430. Mor A, Hani K, Nicolas P. 1994a. The vertebrate peptide antibiotics dermaseptins have

Vol. 71, Nr. 9, 2006JOURNAL OF FOOD SCIENCE

R133

R: Concise Reviews in Food Science

Promises and premises in food safety . . .

Promises and premises in food safety . . .


overlapping structural features but target specific microorganisms. J Biol Chem 269:3163541. Mor A. 2000. Peptide-based antibiotics: a potential answer to raging antimicrobial resistance. Drug Dev Res 50:4407. Mor A. 2001. Antimicrobial peptides. The Kirk-Othmer encyclopedia of chemical technology by Wiley InterScience. John Wiley & Sons Inc. Available from: http://www.mrw.interscience.wiley.com:8095/articles/peptwise.a01/frame.html. Mor A, Nicolas P. 1994a. Isolation and structure of novel defensive peptides from frog skin. Eur J Biochem 219:14554. Mor A, Nicolas P. 1994b. The NH2-terminal helical domain 1-18 of dermaseptin is responsible for antimicrobial activity. J Biol Chem 269:19349. Mor A, Amiche M, Nicolas P. 1994b. Structure, synthesis, and activity of dermaseptin B, a novel vertebrate defensive peptide from frog skin: relationship with adenoregulin. Biochemistry 33:664250. Muriana P. 1993. Antimicrobial peptides and their relation to food quality. In: Spanier AM, Okai H, Tamura M, editors. Food flavor and safety. Molecular analysis and design. Washington, D.C.: American Chemical Society. 30321. Mygind PH, Fischer RL, Schnorr KM, Hansen MT, Sonksen CP, Ludvigsen S, Raventos D, Buskov S, Christensen B, De Maria L, Taboureau O, Yaver D, Elvig-Jorgensen SG, Sorensen MV, Christensen BE, Kjaerulff S, Frimodt-Moller N, Lehrer RI, Zasloff M, Kristensen HH. 2005. Plectasin is a peptide antibiotic with therapeutic potential from a saprophytic fungus. Nature 437:97580. Navon-Venezia S, Feder R, Gaidukov L, Carmeli Y, Mor A. 2002. Antibacterial properties of dermaseptin S4 derivatives with in vivo activity. Antinicrob Agents Chemother 46:68994. Nicolas P, Mor A. 1995. Peptides as a weapon against microorganisms in the chemical defense system of vertebrates. Annu Rev Microbiol 49:277304. Nicolas P, Vanhoye D, Amiche M. 2003. Molecular strategies in biological evolution of antimicrobial peptides. Peptides 24:166980. Nielsen JW, Dickson JS, Crouse JD. 1990. Use of a bacteriocin produced by Pediococcus acidilactici to inhibit Listeria monocytogenes associated with fresh meat. Appl Environ Microbiol 56:21425. Oh JE, Hong SY, Lee KH. 1999. Structure-activity relationship study: short antimicrobial peptides. J Pept Res 53(1):416. Osusky M, Zhou G, Osuska L, Hancock RE, Kay WW, Misra S. 2000. Transgenic plants expressing cationic peptide chimeras exhibit broad-spectrum resistance to phytopathogens. Nat Biotechnol 18:11626. Papagianni M. 2003. Ribosomally synthesized peptides with antimicrobial properties: biosynthesis, structure, function, and applications. Biotechnol Adv 21:46599. Papo N, Braunstein A, Eshhar Z, Shai Y. 2004. Suppression of human prostate tumor growth in mice by a cytolytic D-, L-amino acid peptide: membrane lysis, increased necrosis, and inhibition of prostate-specific antigen secretion. Cancer Res 64:5779 86. Park IY, Cho JH, Kim KS, Kim YB, Kim MS, Kim SC. 2004. Helix stability confers salt resistance upon helical antimicrobial peptides. J Biol Chem 279:13896901. Park YK, Bearson B, Bang SH, Bang IS, Foster J. 1996. Internal pH crisis, lysine decarboxylase and the acid tolerance response of Salmonella typhimurium. Mol Microbiol 20:60511. Perez-Paya E, Houghten RA, Blondelle SE. 1994. Determination of the secondary structure of selected melittin analogues with different haemolytic activities. Biochem J 299:58791. Pouny Y, Rapaport D, Mor A, Nicolas P, Shai Y. 1992. Interaction of antimicrobial dermaseptin and its fluorescently labeled analogues with phospholipid membranes. Biochem 31:1241623. Projan SJ, Blackburn P. 1993. The bacteriocin nisin activated by chelating agents is bactericidal for Helicobacter pylori in vitro. Gastroenterology 104:173. Radzishevsky IS, Rotem S, Zaknoon F, Gaidukov L, Dagan D, Mor A. 2005. Effects of acyl versus aminoacyl conjugation on the properties of antimicrobial peptides. Antimicrob Agents Chemother 49(6):241220. Rayman K, Malik N, Hurst N. 1983. Failure of nisin to inhibit outgrowth of Clostridium botulinum in a model cured meat system. Appl Environ Microbiol 46:14502. Richard H, Foster JW. 2004. Escherichia coli glutamate- and arginine-dependent acid resistance systems increase internal pH and reverse transmembrane potential. J Bacteriol 186:603241. Riley LW, Remis RS, Helgerson SD, McGee HB, Wells JG, Davis BR, Herbert RJ, Olcott ES, Johnson LM, Hargrett NT, Blake PA, Cohen ML. 1983. Hemorrhagic colitis associated with a rare Escherichia coli serotype. N Engl J Med 308:6815. Robinson JA, Shankaramma SC, Jetter P, Kienzl U, Schwendener RA, Vrijbloed JW, Obrecht D. 2005. Properties and structure-activity studies of cyclic beta-hairpin peptidomimetics based on the cationic antimicrobial peptide protegrin. I Bioorg Med Chem 13(6):205564. Rogers LA. 1928. The inhibiting effect of Streptococcus lactis on Lactobacillus bulgaricus. J Bacteriol 16:3215. Rotem S, Radzishevsky I, Inouye RT, Samore M, Mor A. 2006. Identification of antimicrobial peptide regions derived from genomic sequences of phage lysins. Peptides 27(1):1826. Rydlo T, Rotem S, Mor A. 2006. Antibacterial properties of dermaseptin s4 derivatives under extreme incubation conditions. Antimicrob Agents Chemother 50(2):490 7. Samson RA, Hoekstra E, Frisvad JC, Filtenborg O. 1995. Introduction to food-borne fungi. The Netherlands: CBS, Baarn, p 299. Scott MG, Hancock REW. 2000. Cationic antimicrobial peptides and their multifunctional role in the immune system. Crit Rev Immunol 20:40731. Scott MG, Yan H, Hancock REW. 1999. Biological properties of structurally related alpha-helical cationic antimicrobial peptides. Infect Immun 67:20059. Shalev DE, Mor A, Kustanovich I. 2002. Structural consequences of carboxyamidation of dermaseptin S3. Biochemistry 41:73127. Shalev DE, Rotem S, Fish A, Mor A. 2005. Consequences of N-acylation on structure and binding properties of dermaseptin derivative K4S4(1-13). J Biol Chem 281(14):9432 8. Shi J, Ross CR, Chengappa MM, Style MJ, McVey DS, Blecha F. 1996. Antibacterial activity of a synthetic peptide (PR-26) derived from PR-39, a prolinearginine-rich neutrophil antimicrobial peptide. Antimicrob Agents Chemother 40:11521. Shinnar AE, Uzzell T, Rao MN, Spooner E, Lane WS, Zasloff MA. 1996. In: Kaumaya PTP, Hodges RS, editors. Peptides chemistry structure and biology, proceedings of the 12th American peptide symposium. London: Mayflower Scientific. 18991. Simmaco M, Mignogna G, Barra D. 1998. Antimicrobial peptides from amphibian skin: what do they tell us? Biopolymers 47:43550. Small P, Blankenhorn D, Welty D, Zinser E, Slonczewski, McGarrity JT, Armstrong JB. 1994. Acid and base resistance in Escherichia coli and Shigella flexneri: role of rpoS and growth pH. J Bacteriol 176:172937. Soballe PW, Maloy WL, Myrga ML, Jacob LS. Herlyn M. 1995. Experimental local therapy of human melanoma with lytic magainin peptides. Int J Cancer 60:2804. Steele BT, Murphy N, Rance CP. 1982. An outbreak of hemolytic uremic syndrome associated with ingestion of fresh apple juice. J Pediatr 101:9635. Stumpe S, Bakker EP. 1997. Requirement of a large K+-uptake capacity and of extracytoplasmic protease activity for protamine resistance of Escherichia coli. Arch Microbiol 167:12636. Suppakul P, Miltz J, Sonneveld K, Bigger SW. 2003a. Active packaging technologies with an emphasis on antimicrobial packaging and its applications. J Food Sci 68(2):408 20. Suppakul P, Miltz J, Sonneveld K, Bigger SW. 2003b. Antimicrobial properties of basil and its possible application in food packaging. J Agric Food Chem 5:3197207. Suppakul P, Miltz J, Sonneveld K, Bigger SW. 2006. Characterization of antimicrobial films containing basil extracts. Packag Technol Sci 19(5):25968. Tailor RH, Acland DP, Attenborough S, Cammue BP, Evans IJ, Osborn RW, Ray JA, Rees SB, Broekaert WF. 1997. A novel family of small cysteine-rich antimicrobial peptides from seed of Impatiens balsamina is derived from a single precursor protein. J Biol Chem 272:244807. Tang YQ, Yuan J, Osapay G, Osapay K, Tran D, Miller CJ, Ouellette AJ, Selsted ME. 1999. A cyclic antimicrobial peptide produced in primate leukocytes by the ligation of two truncated -defensins. Science 286:498502. Tauxe RV. 2002. Emerging foodborne pathogens. Int J Food Microbiol 78(1-2):31 41. Thouzeau C, Le Maho Y, Froget G, Sabatier L, Le Bohec C, Hoffmann JA, Bulet P. 2003. Spheniscins, avian beta-defensins in preserved stomach contents of the king penguin, Aptenodytes patagonicus. J Biol Chem 278:510538. Tichaczek PS, Vogel RF, Hammes WP. 1994. Cloning and sequencing of sakP encoding sakacin P , the bacteriocin produced by Lactobacillus sake LTH 673. Microbiology (New York) 140:3617. Tomasz A. 1994. Multiple-antibiotic-resistant pathogenic bacteria. A report on the Rockefeller Univ. Workshop. N Engl J Med 330:124751. Tomita H, Fujimoto S, Tanimoto K, Ike Y. 1996. Cloning and genetic organization of the bacteriocin 31 determinant encoded on the Enterococcus faecalis pheromoneresponsive conjugative plasmid pYI17. J Bacteriol 178:358593. Tossi A, Sandri L, Giangaspero A. 2000. Amphipathic, -helical antimicrobial peptides. Biopolymers 55:430. Turner K, Porter J, Pickup R, Edwards C. 2000. Changes in viability and macromolecular content of long-term batch cultures of Salmonella typhimurium measured by flow cytometry. J Appl Microbiol 89:909. Ulvatne H, Vorland LH. 2001. Bactericidal kinetics of 3 lactoferricins against Staphylococcus aureus and Escherichia coli. Scand J Infect Dis 33:50711. Utsugi T, Schroit AJ, Connor J, Bucana CD, Fidler IJ. 1991. Elevated expression of phosphatidylserine in the outer membrane leaflet of human tumour cells and recognition by activated human blood monocytes. Cancer Res 51:30626. Uyttendaele M, Debevere J. 1994. Protamine evaluation of the antimicrobial activity of protamine. Food Microb 11:41727. Vanhoye D, Bruston F, Nicolas P, Amiche M. 2003. Antimicrobial peptides from hylid and ranin frogs originated from a 150-million-year-old ancestral precursor with a conserved signal peptide but a hypermutable antimicrobial domain. Eur J Biochem 270:206881. Van Kan EJ, Demel RA, Breukink E, van der Bent BA, Kruijff B. 2002. Clavanin permeabilizes target membranes via two distinctly different pH-dependent mechanisms. Biochemistry 41:752939. Vogt TCB, Bechinger B. 1999. The interactions of histidine-containing amphipathic helical peptide antibiotics with lipid bilayers. The effects of charges and pH. J Biol Chem 274(41):2911521. Vouldoukis I, Shai I, Nicolas P, Mor A. 1996. Antimicrobial properties of skin-PYY. FEBS Lett 380:23740. Welling MM, Paulusma-Annema A, Balter HS, Pauwels EK, Nibbering PH. 2000. Technetium-99m labelled antimicrobial peptides discriminate between bacterial infections and sterile inflammations. Eur J Nucl Med 27(3):292301. [WHO] World Health Organization. 1995. The use of essential drugs. Sixth Report of the WHO expert committee, WHO Tech. Rep. Ser. No. 850. Rome:WHO. Wilcox W, Eisenberg D. 1992. Thermodynamics of melittin tetramerization determined by circular dichroism and implications for protein folding. Protein Sci 1:641 53. Wu M, Maier E, Benz R, Hancock REW. 1999. Mechanism of interaction of different classes of cationic antimicrobial peptides with planar bilayers and with the cytoplasmic membrane of Escherichia coli. Biochemistry 38:723542. Yam KL, Takhistov PT, Miltz J. 2005. Intelligent packaging: concepts and applications. J Food Sci 70(1):R1R10. Yan H, Hancock REW. 2001. Synergistic interactions between mammalian antimicrobial defense peptides. Antimicrob Agents Chemother 45:155860. Yaron S, Rydlo T, Shachar D, Mor A. 2003. Activity of dermaseptin K4-S4 against foodborne pathogens. Peptides 24(11):181521. Yasin B, Pang M, Turner JS, Cho Y, Dinh NN, Waring AJ, Lehrer RI, Wagar EA. 2000. Evaluation of the inactivation of infectious Herpes simplex virus by host-defense peptides. Eur J Clin Microbiol Infect Dis 19:18794. Yeaman MR, Yount NY. 2003. Mechanisms of antimicrobial peptide action and resistance. Pharmacol Rev 55:2755. Zasloff M. 1987. Magainins, a class of antimicrobial peptides from Xenopus skin:

R: Concise Reviews in Food Science

R134

JOURNAL OF FOOD SCIENCEVol. 71, Nr. 9, 2006

URLs and E-mail addresses are active links at www.ift.org

isolation, characterization of two active forms, and partial cDNA sequence of a precursor. Proc Natl Acad Sci USA 84:544953. Zasloff M. 2002. Antimicrobial peptides of multicellular organisms. Nature 415:38995. Zasloff M, Martin B, Chen HC. 1988. Antimicrobial activity of synthetic magainin peptides and several analogues. Proc Natl Acad Sci USA 85:9103. Zhang L, Hancock REW. 2000. Peptide antibiotics. In: Hughes D, Andersson DI, editors.

Antibiotic resistance and antibiotic development. Reading,Pa: Harwood Academic Publishers. 20932. Zhou YX, Cao W, Luo QP, Ma YS, Wang JZ, Wei DZ. 2005. Production and purification of a novel antibiotic peptide, adenoregulin, from a recombinant Escherichia coli. Biotechnol Lett 27(10):72530.

URLs and E-mail addresses are active links at www.ift.org

Vol. 71, Nr. 9, 2006JOURNAL OF FOOD SCIENCE

R135

R: Concise Reviews in Food Science

Promises and premises in food safety . . .

Você também pode gostar