Sei sulla pagina 1di 18

Anti-Infective Agents in Medicinal Chemistry, 2009, 8, 327-344

327

Peptidomimetics and their Applications in Antifungal Drug Design


Shoeib Moradi, Saeed Soltani, Alireza M. Ansari and Soroush Sardari*
Drug Design and Bioinformatics Unit, Medical Biotechnology Department, Biotechnology Research Center, Pasteur Institute of Iran, #69 Pasteur Ave., Tehran, Iran 13164
Abstract: The antimicrobial peptides represent diverse structures for drug design. They have been looked at as potential sources of new antimicrobial drugs to combat the increasing threat posed by multiple drug resistant microorganisms. Unfortunately, peptides themselves provide inferior drug candidates because of their low oral bioavailability, potential immunogenicity, poor in vivo metabolic stability, high molecular weight and most importantly being exposed by enzymes like proteases. Recent efforts to resolve disadvantageous peptide characteristics, and thus generating practical pharmaceutical therapies, have focused on the creation of non-natural peptide mimetics. Peptidomimetic molecules may have reduced immunogenicity and improved bioavailability relative to peptide analogues. Also the artificial backbone makes most peptidomimetics resistant to degradative enzymes thus increasing the stability of peptidomimetic drugs in the body. In this article, after introducing antifungal peptides, benefits and limitations, and peptidomemetics usage are discussed and applications in drug discovery process and antifungal research will be presented.

INTRODUCTION The occurrence of fungal infections that have been seen in the growing populations of immunocompromised hosts, such as individuals infected with HIV, transplant recipients and patients with cancer, has increased dramatically in the last few decades [1]. Of the thousands of recognized fungal species, practically a couple, are pathogenic and produce mycotic infections in humans and animals. In tropical and subtropical developing countries, dermatophytes and Candida sp. can cause infections in humans and other animals, especially in immunocompromised patients. In addition, the number of reported cases of immunocompromised patients which often develop opportunistic and superficial mycoses, such as candidiasis, cryptococcosis and aspergillosis, has increased dramatically in recent years, especially in those with AIDS [2] The yeast fungus, Cryptococcus neoformans, has been identified as the fourth most common cause of lifethreatening infection in AIDS patients. Potentially fatal infections with Candida albicans and other species of Candida are also known [3]. According to Sardari and Dezfulian, the fungi have varied susceptibility to different antifungal agents [4]. In other words, the dearth of wide-spectrum antifungal agents is of great alarm to medical mycologists and practitioners. Long-term treatment with the commonly used antifungals such as amphotericin B has toxic effects on the patient; flucytosine and echinocandins have a restricted spectrum, while azoles may result in strain resistance [5]. Therefore, in the search for a substitute form of treatment for fungal infections, the last decade has seen a rise in novel approaches, such as therapeutic antibodies and peptide molecules. Antimicrobial peptides have newly become the focus of considerable interest as a candidate for a new type of antibiotic, primarily due to their potency against pathogenic mi*Address correspondence to this author at the Drug Design and Bioinformatics Unit, Department of Biotechnology, Pasteur Institute of Iran, #69 Pasteur Ave., Tehran, Iran 13164; Tel: (98-21) 6640-5535; Fax: (98-21) 6646-5132; E-mail: ssardari@hotmail.com;sardari@pasteur.ac.ir 1871-5214/09 $55.00+.00

crobes that are resistant to conventional antibiotics, as well as their broad-spectrum activity [6]. Since the first identification of cecropin [7] and defensin [8] in insect hemolymph and human neutrophils, respectively, several antimicrobial peptides have been isolated from a wide variety of organisms, including vertebrates, invertebrates, and plants [9]. Other than pathogen-lytic activities, these peptides have other properties like anti-tumor, mutagen activity, or act as signaling molecules [10]. In addition, they have a number of biotechnological applications, e.g. in transgenic plants [11], in aquaculture, and as aerosol spray for patients of cystic fibrosis [12]. Among AMPs, there is a group with considerable fungicidal effect referred to as antifungal peptides (AFPs). Antifungal and antimicrobial peptides are becoming the focus of interesting molecules among scientists, as they are vital components of the innate defense of all species, they kill very quickly, do not simply select resistant mutants and are synergistic with potentially toxic conventional therapeutic agents against microbes. Some of these agents have reached clinical trials, while others are undergoing detailed preclinical testing [13]. Therefore, the search continues for new antibiotics that are active in vivo, fast acting and broadspectrum, do not induce fungal and bacterial resistance and have limited side effects. In addition to the properties described above, they have low MICs and broad-spectrum activity in both low and high ionic strength conditions [14], neutralize lipopolysacharides [15], encourage injury healing [16] and have synergistic activity with conventional antibiotics [17]. Very few side effects have been reported. Synthetic congeners of natural antimicrobial peptides are good candidates. Synthetic congeners including a different series of peptides truncated successively from the carboxyl-terminal end of larger monomer and some analogs, which have lysine residues in place of two internal histidines or have a lysine added to the amino terminus of the original molecule. Not only synthetic congeners benefit from natural properties of antifungal peptides, but also with alteration in structural

2009 Bentham Science Publishers Ltd.

328 Anti-Infective Agents in Medicinal Chemistry, 2009, Vol. 8, No. 4

Moradi et al.

characteristic of original antifungal peptides, their bioactivity would increase. ANTIFUNGAL PEPTIDES Classification of Antifungal Peptides There is a great diversity of anti-fungal peptides, with large variations in molecular mass, N-terminal sequence and antifungal specificity. Antifungal peptides are classified in several ways. Two major classification of antifungal peptide included classification by their mode of action and by the source. They will be described briefly here. According to the classification by the mode of action, AFPs are divided to the two groups. The first group acts by lysis, which occurs by means of several mechanisms. Lytic peptides can be amphipathic, that is, molecules with two faces, with one being positively charged and the other being neutral and hydrophobic. Some amphipathic peptides bind merely to the membrane surface and can disrupt the membrane structure without traversing the membrane. Others traverse membranes and interact distinctively with certain molecules. Lastly, other amphipathic peptides aggregate in a discriminating manner, forming aqueous pores of variable sizes, allowing passage of ions or other solutes [18]. The second peptide group interferes with cell wall synthesis or the biosynthesis of critical cellular components such as glucan or chitin [19]. An excellent review of lipopeptide antifungal agents affecting cell wall synthesis has been published previously [20]. Also Antifungal peptide can be classified according to their origin of isolation. Table 1 shows some antifungal peptide from different origin. Mode of Action of Antifungal Peptides The method of fungal cell lysis by peptides usually include non-specific interaction with the membrane phospholipids rather than binding to exact receptors on the cell membrane so that microorganisms develop resistance to antimicrobial peptides at rates that are orders of magnitude less than those observed for conventional antibiotics [21]. On the down side, the toxicities of many of the peptides and their fast rate of clearance from the circulation mean that topical or in vitro applications may be more appropriate than systemic administration [22]. Antifungal peptides are found in numerous organisms, ranging from fungi through mammals [23]. The majority of antifungal peptides constitute a simple structural motif; they are most commonly short (<40 residues) linear, cationic, amphipathic -helices. They exert antibacterial activity by cell membrane permeabilization and lysis, even though the accurate lytic mechanism has not been conclusively determined. Selectivity for the fungal cell appears to be mediated by favorable electrostatic interaction between positively charged peptides and the negatively charged cell surface [24]. However, an excessively hydrophobic peptide can bind arbitrarily to any cell membrane [25]. Antifungal peptide selectivity is dependent upon a precise balance of peptide hydrophobicity and electrostatic charge. In general, studies propose that the overall physicochemical parameters of antifungal peptides, rather than any specific receptorligand interactions, are responsible for antifungal activity [26]. As a result, antifungal peptides are attractive targets for biomimicry and peptidomimetic lead de-

velopment, as reproduction of critical peptide biophysical characteristics in an unnatural, sequence-specific mimicking the source oligomer should presumably be enough to donate antifungal efficacy, while circumventing the restrictions related to peptide pharmaceuticals [27]. Models of Peptide Interaction with Lipid Membranes According to the above-mentioned mode of action, a peptide which has lethality effect on fungi should be capable of forming ion channels in membrane by aggregation (pseudoionophores) and insertion in order to span cell membrane with 2.5-4.0 nm thickness (depending upon lipid composition) and should have at least 12 amino acids [28, 29]. Being laterally amphipathic meaning that one face of the helix displaying hydrophobic residues, while the opposite face displays hydrophilic residues, the peptide can form hydrophilic ion channels or pores and at the same time remain in contact with the hydrophobic components, e.g. fatty acyl moieties [30]. Based on these properties, two principal modes of action for membrane-perturbing peptides have been proposed: pore formation across the lipid bilayer or a "carpet" mechanism, lysing the membrane in a detergent-like manner. In the latter model, peptides "carpet" the surface of a target membrane and when sufficiently accumulated, create numerous pores [30]. The transmembrane model involves the peptides forming pores through the outer membrane: the "barrelstave" [31] and toroidal pore [32] mechanisms. In these models, the peptides oligomerize to form pores through the membrane. The pores act as non-selective channels for ions, toxins and metabolites, thus preventing the organism from maintaining homeostasis. Peptides with 20 or more amino acids lend themselves to these mechanisms, as they are able to span the lipid bilayer when in -helical conformation. A key difference between these two mechanisms is the positioning of the headgroup region of the lipid molecules with respect to the peptide. In the barrel-stave mechanism, the headgroups remain located along the membrane surface, while the pore is formed by the interaction of the peptide within the hydrophobic core of the membrane. The transmembrane pore is lined by the hydrophilic surface of the peptide. According to BSPM (Barrel-Stave Pore Model), AFPs should have distinct structure such as -helix or sheet or both and have hydrophobic interaction with the target membrane. By contrast, toroidal pores are formed when the peptides insert in such a way to cause the inner and outer membrane leaflets to curve and the lumen is lined by the hydrophilic surface of the peptide interspersed by the phospholipid headgroups. TPM (Toroidal Pore Model) is based on formation of several short-lived clusters of an undefined nature of secondary structure [75]. Preference for barrelstave vs. toroidal pore may depend on several factors including the peptide length and membrane thinning effect induced by the peptide [76]. Disadvantages of Antifungal Peptides The development of peptides as drugs is problematic as a result of poor oral and tissue absorption, rapid proteolytic cleavage and poor shelf-life or stability. Most proteins and small peptides are easily proteolyzed, rapidly excreted and poorly bioavailable. Much effort has been expended to find ways to replace portions of peptides with nonpeptide struc-

Peptidomimetics and their Applications in Antifungal Drug Design

Anti-Infective Agents in Medicinal Chemistry, 2009, Vol. 8, No. 4

329

Table 1.

Antifungal Peptides from Different Origins


Source Length Mode of Action Target fungus MIC (g/ml) Refs.

Peptide

Mammalian antifungal peptides Histatin 3 Halocidin Protegrins 1 Tritrptcin PAMP NP-3B SAMP29 Saliva of humans Halocynthia aurantium Human, porcine Human, porcine Human, porcine Rabbit Sheep 31 18 16 13 24 33 45 Lysis Lysis Lysis Lysis Unknown Lysis Lysis C. albicans C. albicans C. albicans A. flavus C. albicans A. fumigates C. albicans 75 10 3 250 23 100 4 [33] [34] [35] [36] [37] [38] [39]

Insect and amphibian antimicrobial peptides Melitin Cecropins B DSP Magainin 2 Drosomycin Tenecin Cicadin Esculentin-1 Brevin -2 Apis mellifera H. cecropia Leaf beetle Xenopus laevis D. melanogaster Inset defencin Juvenile cicadas Phyllomedusa sauvagii Phyllomedusa sauvagii 25 35 41 23 44 43 45 47 40 Lysis Lysis Unknown Lysis Lysis Lysis Lysis Lysis Lysis C. albicans A. fumigatus C. albicans C. albicans F. oxysporum C. albicans C. albicans C. albicans C. albicans 1.5 9.5 7 80 5 100 70 11 10 [40] [41] [42] [43] [44] [45,46] [47] [48] [49]

Bacterial and fungal antifungal peptides Anafp Fungicin M-4 HP Helioferin A Iturin A Leucinostatin A Nikkomycin X Polyoxin D Schizotrin A Trichopolyn A Pneumocandin A0 FR900403 CB-1 Aspergillus niger Bacillus licheniformis Helicobacter pylori Mycogone rosea Bacillus subtilis Phaeohelotium lilacinum Streptomyces tendae Streptomyces cacaoi Schizotrix sp. Trichoderma polysporum Zalerion arboricola 55 Cyclic peptide 11 Lipopeptide Lipopeptide Amino-lipopeptide Peptide-nucleoside Trinucleoside peptide Cyclic undecapeptide Amino-lipopeptide Lipopeptide Lysis Unknown Unknown Unknown Lysis Unknown Chitin synthesis Chitin synthesis Unknown Lysis Glucan synthesis C. albicans Mucor sp. C. albicans C. albicans S. cerevisiae C. neoformans C. immitis C. immitis C. albicans C. neoformans C. albicans 50 8.0 12.5 5.0 22. 0.5 0.125 0.125 0.02 0.78 0.1 [50] [51] [52] [53] [54] [55] [56] [57] [58] [59] [60]

Kernia sp. Bacillus licheniformis

Lipopeptide Lipopeptide

Chitin synthesis Chitin binding

F. oxysporum F. oxysporum

0.4 50

[61] [62]

Plant antifungal peptides Zeamatin Zea mays 27 Lysis C. albicans 0.5 [63]

330 Anti-Infective Agents in Medicinal Chemistry, 2009, Vol. 8, No. 4


(Table 1) contd.

Moradi et al.

Peptide ACE-AMP1 Pe-AFP1 Pinin Coccinin Cicerarin Gymnin Beta-basrubrin

Source Allium cepa Passiflora edulis Red Bean Runner beans Green chickpea Yunnan bean Ceylon spinach seeds

Length 84 25 45 10 20 10 36

Mode of Action Unknown Lysis Unknown Unknown Unknown Unknown Unknown

Target fungus F. oxysporum A. fumigatus F. oxysporum F. oxysporum F. oxysporum F. oxysporum F. oxysporum

MIC (g/ml) 0.3 32 3.5 7.5 8.2 59 12.5

Refs. [64] [46] [65] [66] [67] [68] [69]

Synthetic and semisynthetic antifungal peptides LY303366 D4E1 MP1 Dhvar Cilofungin Lipopeptide Linear peptide Linear peptide Linear peptide Lipopeptide 6 cyclic 17 11 14 Lipopeptide Glucan synthesis Lysis Unknown Unknown Glucan synthesis Candida krusei A. flavus C. albicans C. albicans C. albicans 0.5 26 5 14 0.62 [70] [71] [72] [73] [74]

tures, called peptidomimetics, in the hope of obtaining more bioavailable entities [30]. Peptidomimetics can be seen as probes used in the transition pathway of small molecule drug design. Cyclization of the peptide backbone and its modification with aromatic residues constitutes an effective approach to mimic drug structures in the design process and circumvents obstacles associated with delivery and formulation of peptides. In the recent years [25] examples of mimicking pturn structures has led to combining design strategies with molecular libraries, demonstrating that peptidomimetics can provide valuable clues about receptor similarities not revealed by their endogenous ligands. PEPTIDOMIMETICS Recent efforts to reorganize disadvantageous peptide characteristics, and thus generate viable pharmaceutical therapies, have focused on the creation of non-natural peptide mimics. A peptidomimetic is a small protein-like chain designed to mimic a peptide which is often used in the literature to indicate a multitude of structural types that differ in fundamental ways. The term of peptidomimetic is often applied to highly modified analogs of peptides without distinguishing how these differ from classical analogs of peptides. These Non-natural, sequence-specific peptidomimetic oligomers are being designed to mimic bioactive peptides, with potential therapeutic application. These peptidomimetics can be based on any oligomer that mimics peptide primary structure through use of amide bond isosteres and/or modification of the native peptide backbone, including chain extension or heteroatom incorporation. Peptidomimetic oligomers are often protease-resistant, and may have reduced immunogenicity and improved bioavailability relative to peptide analogues [77]. In addition to primary structural mimicry, a select subset of the sequence-specific peptidomimetic oligomers, the so-called foldamers [78], exhibits well-defined

secondary structural elements such as helices, turns and small, sheet-like structures [79]. When peptide bioactivity is contingent upon a precise 3D structure, the capacity of a biomimetic oligomer to fold can be very important. They typically arise from modification of an existing peptide in order to alter the molecule's properties. For example, they may arise from modifications to change the molecule's stability or biological activity [80, 81]. This can have a role in the development of drug-like compounds from existing peptides. These modifications involve changes to the peptide that will not occur naturally (such as altered backbones and the incorporation of non-natural amino acids). Peptidomimetics are part of the wide effort by researchers, research labs and institutions to create cures for cancer by means of restoring or activating apoptotic pathways in specific cells. An example of peptidomimetics were those designed and synthesized with the purpose of binding to target proteins in order to induce cancer cells into a form of programmed cell death called apoptosis [82]. Basically, these work by mimicking key interactions that activate apoptotic pathway in the cell. The unfavorable pharmacokinetic properties associated with peptides when used as orally administered drugs can, in principle, be avoided by development of peptidomimetics. The general strategy when preparing peptidomimetics is to replace segments related to undesired properties with non-peptidic structures, while attempting to maintain the ability to elicit the same or improved biological response as the native peptide [83, 84]. Classification of Peptidomimetics Based on Rikpa and Rich classification [85], three most important classes of peptidomimetics have been described. The first type includes structures that mimic the local topography about an amide bond, i.e., amide bond isosteres, pyrrolinones or short portions of secondary structure ( turns).

Peptidomimetics and their Applications in Antifungal Drug Design

Anti-Infective Agents in Medicinal Chemistry, 2009, Vol. 8, No. 4

331

The peptide backbone mimetics are also known as type-I mimetics. The first class is characterized by backbone changes, such as incorporation of amide bond isosteres and turn mimetics. The literature concerning peptidomimetics of class I including stabilized turn mimetics represented as bicycles [86, 87], aromatics [88, 89] and cyclic compounds [90]. These mimetics regularly match the peptide backbone atom-for-atom, while retaining functionality that makes significant contacts with binding sites. Latest studies paid attention to transition-state isosteres or collected substrate/product mimetics designed to mimic reaction pathway intermediates of the enzyme-catalyzed reactions. The reduced amide isostere developed by Szelke et al. [91] and the statine (hydroxylmethylene) isostere were early mimetics used to design inhibitors of a variety of aspartic proteases [92], and their success led to other tetrahedral intermediate mimics such as the hydroxylethylene and hydroxyethylamine isosteres. Cathepsin K, is new type-I peptidomimetic which is inhibitor of cysteine protease [86]. Two small-molecule inhibitors with an IC50 of 0.3 M and 0.013M have been designed lately, by Maibaum and co-workers [93, 94] which are orally active inhibitors of renin, the protease that catalyzes the first reaction in the renin-angiotensin system. The cysteine protease interleukin-lP-converting enzyme (ICE) inhibitor was designed by Dolle et al. [95] to generate potent aldehyde inhibitors of ICE. McKittrick et al. [96] designed a novel triple inhibitor of endothelin-converting enzyme (ECE), angiotensin-converting enzyme (ACE) and neutral endopeptidase (NEP) as a type-I peptidomimetic inhibitor. It has been shown that opportunistic fungi such as Candida albicans require N-myristoyltransferase (NMT) for viability. Brown et al. [97] created type-I mimetic compounds which are selective for C. albicans over human NMT. The second class of peptidomimetic is referred to as ligands exerting the same biological response as the native peptide ligand without any obvious structural resemblance (functional mimetics). The second type of mimetics is type-II or the functional mimetics, which are small nonpeptide molecules that binds to a peptide receptor. Initially these were presumed to be direct structural analogs of the natural peptide, but characterization of both the endogenous peptide and antagonist binding sites by site-directed mutagenesis [98] indicated that antagonists bind to subsites that are different from those used by the parent peptide for a large number of receptors. As a result, these functional mimetics do not essentially mimic the structure of the parent compound or hormone. The third class is represented by peptidomimetics with a nonpeptidic core structure, which position key functionalities for interactions with the receptor in a closely related way as the native peptide. Some examples from the vast literature [86] in the field are peptidomimetics of vasopressin [99], oxytocin [100], LHRH [101], somatostatin and angiotensin II [102, 103]. Type-III mimetics represent the idyllic peptidomimetics in that they possess novel templates which, though appearing unrelated to the original peptides, contain the necessary groups positioned on a novel nonpeptide scaffold to serve as topographical mimetics [104, 105]. At present, there is no systematic way to transform the structure of an enzyme-bound peptide substrate analog into a nonpeptide mimetic.

Type III peptidomimetics resemble the peptides, but employing a non-peptide scaffold to position key pharmacophores for receptor interaction, whereas type II mimetics, which are functional agonists or antagonists, do not structurally mimic the native peptide [106]. A wide variety of peptidomimetic building blocks have been developed in the past: e.g. peptoids [107], sulphonamides [108], ureapeptidomimetics [109], hydrazinopeptidomimetics [110], peptides [111], oligopyrrolinones [112], peptidosulfonamides [113], oligocarba-mates [114], oligoureas [110], azatides [111], and ethoxyformacetais [115] with unnatural oligoamide backbones which are self-organized at the molecular level to form stable helices useful to mimic protein secondary structure elements [116]. The predictability of folding of these oligomeric strands led to the development of molecules with functions including potent inhibitors of protein-protein interactions [117-119]. A set of novel type-III mimetics have been obtained and found to have activity towards Factor Xa with IC50=5 M [120]. Sall et al. and Chirgadze et al. [121, 122] designed Type-III peptidomimetic inhibitors of thrombin also. Another type-III mimetic was developed by a group that used pyranoses as templates to design thrombin inhibitors [98]. Some type-III mimetics of Inhibitors of Ras-farnesyltransferase (R-FT) have been developed by mimicking the carboxy-terminal CAAX motif which is the signal for farnesylation of Ras proteins [123125]. Such "unnatural biopolymers '' may have markedly different physicochemical proper-ties than natural peptides, including a superior pharmacological profile for development into therapeutic agents [126]. Finding unnatural oligoamide backbones can adopt well-defined and controlled helical secondary structures suggested that one could use them as scaffolds to distribute charged side chains in a predictable manner for de novo design of cationic amphiphilic molecules mimicking natural host defense -peptides. Antimicrobials based on either -peptides 314, 2.512, or 2.710,12 helical folds [127-131] or on peptoids polyproline type I-like [132] have been found to exhibit selective (non-hemolytic) and potent antibacterial activity against both Gram-positive and Gram-negative bacteria and fungi. Most antimicrobial peptidomimetics are typically discrete non-natural oligomers whose units are in many cases, connected via amide bonds. Much of the synthetic interest in peptidomimetics comes from the fact that these oligomers can present a wide variety of side chains which could be chemically identical to those found in natural peptides, but along an artificial backbone. The consequence of this hybrid structure is that peptidomimetics can mimic the conformation and functionality of biopolymers yet are not limited by the side chains of the main twenty naturally occurring -amino acid building blocks. Also the artificial backbone makes most peptidemimetics resistant to bio-degradation enzymes thus increasing the stability of peptidomimetic drugs in the body [133]. SPECIFICATIONS AND BIOACTIVITY OF PEPTIDOMIMETICS -Peptides -peptides consist of amino acids, which have their amino group bound to the carbon rather than the carbon

332 Anti-Infective Agents in Medicinal Chemistry, 2009, Vol. 8, No. 4

Moradi et al.

as in the 20 standard biological amino acids. The only commonly naturally occurring amino acid is -alanine. Although it is used as a component of larger bioactive molecules, -peptides in general do not appear in nature Fig. (1). Due to this reason, -peptide-based antibiotics are being explored as ways of evading antibiotic resistance. In amino acids, both the carboxylic acid group and the amino group are bound to the same carbon, termed the carbon (C) because it is one atom away from the carboxylate group. In amino acids, the amino group is bound to the carbon (C), which is found in most of the 20 standard amino acids. Only glycine lacks a carbon, which means that there is no glycine molecule. The chemical synthesis of amino acids can be challenging, especially given the diversity of functional groups bound to the carbon and the necessity of maintaining chirality. In alanine molecule, the carbon is achiral, however, larger amino acids have a chiral C atom. The -peptides are by far the best well-studied peptidomimetics. Because the backbones of -peptides are longer than those of peptides that consist of -amino acids, peptides form different secondary structures. The alkyl substituent at both the and positions in a amino acid favor a gauche conformation about the bond between the carbon and -carbon. This also affects the thermodynamic stability of the structure. -peptides are stable against proteolytic degradation in vitro and in vivo, an important advantage age over natural peptides in the preparation of peptide-based drugs [8]. -peptides have been used to mimic natural peptide-based antibiotics such as magainins, which are extremely powerful but difficult to use as drugs because they are degraded by proteolytic enzymes in the body [9].
H H N+ H CH2 H H C O O-

CH3 H H H C N+ H C O O-

Many types of helix structures consisting of -peptides have been reported [6]. These conformation types are distinguished by the number of atoms in the hydrogen-bound ring that is formed in solution; 8-helix, 10-helix, 12-helix, 14helix, and 10/12-helix. -peptides have more conformational freedom than -peptides, because of an additional methylene unit present in the polymer backbone. Consequently, whereas -peptide helices most commonly adopt the -helix conformation, -peptide sequences have been shown to adopt several distinct helical conformations, the choice of which depends largely upon the substitution pattern at backbone C and C atoms [134-136]. Of these, the -peptide 12helix and the 14-helix have been successfully employed as magainin mimics. The terms 12-helix and 14-helix correspond to the number of atoms participating in a ring created by intrachain hydrogen bonds. Recent studies in which peptides were designed to mimic the magainin have helped to illustrate which physical characteristics are critical for ideal antibacterial efficacy and biocompatibility in nonnatural oligomers. DeGrado and co-workers [137] first reported the de novo design of amphipathic, cationic, monosubstituted -peptide 14-helices as antibacterial compounds against the Gram-negative bacterium K91 Escherichia coli. Their study led to design of non-hemolytic analogues such as 1 in Fig. (3) [133]. Although potent antibiotics, these peptides displayed poor selectivity for bacteria, as indicated by extensive mammalian red blood cell lysis (haemolysis). Assuming that excessive side-chain hydrophobicity was responsible for the poor selectivity observed, Liu and colleagues [138] modified their original 14-helix designs to substitute a valine-like (-HVal) residue with a less hydrophobic alanine-like (-HAla) residue. In support of their original hypothesis, this modification abolished haemolytic activity, while retaining good antibacterial efficacy in both 12- and 15-residue oligomers. Generally, compounds that are not hemolytic at relative concentrations of their antimicrobial activity are typically deemed selective and have the potential for promising therapeutics that may kill microbial but not mammalian cells. In addition, hemolysis, by far the most common measure of mammalian toxicity, may not be the best predictor [139]. Simultaneously and independently, Seebach and Mathews [140] studied quite similar mono-substituted peptides, also designed to adopt the -peptide 14-helix for antimicrobial and hemolytic activity and demonstrated that "2/3" type peptides such as 2 in Fig. (3) can display selectivity [137, 140]. Interestingly, although haemolytic activity was comparable, this sequence difference alone resulted in a one order of magnitude reduction in antibacterial activity relative to DeGrados refined oligomers. In 2000, Gellman and coworkers has reported a 17-residue -peptide 12-helix, which they called -17, such as 3 in Fig. (3) that formed helices just like the class of host-defense peptides known as the magainins, which have 2030 residues [127]. This particular -17 was reported to be effective against two pathogens that are resistant to common antibiotics plus they are not hemolytic. Unlike DeGrados monosubstituted 14helices, -17 is disubstituted at C and C to form intraresidue five-membered rings and consequently possesses significantly less conformational freedom than either DeGrados or Seebachs oligomers. Therefore, conformational

L--alanine

-alanin

Fig. (1). Structure of alpha and beta alanine.

Two main types of -peptides exist: as it shown in Fig. (2), those with the organic residue (R) next to the amine are called 3-peptides and those with position next to the carbonyl group are called 2-peptides [5].
O ... O NH O
O NH
O O NH R R NH R

(1)

NH O
R

NH

NH

...

O NH

O NH
O

O ...
O

(2)

...

NH

(3)

...

NH

NH R

...

Fig. (2). 1) -peptide, 2) 3-peptide, and 3) 2-peptide general structures.

Peptidomimetics and their Applications in Antifungal Drug Design

Anti-Infective Agents in Medicinal Chemistry, 2009, Vol. 8, No. 4

333

H2N O H NH NH O NH O NH2
4 or 5

1
H2N O H2N NH O NH O NH O NH O NH O NH O OH

2
H3C NH NH NH2
+

NH

NH NH2
+

NH

NH

NH2

3
Fig. (3). 1: Selective antimicrobial -peptides. 2: 2/3 peptide. 3: "-17" -peptide.

rigidity does not appear to adversely affect activity or selectivity. In fact, Gellmans group has recently described a peptide 12-helix with both mono- and di-substituted residues, with activity comparable to -17 and magainin [141]. Regardless of the particular oligomer identity, it is clear that the repertoire of peptidomimetic antibiotics will soon expand beyond the exclusive realm of -peptides. Peptoids Peptoids, or N-substituted glycines, with -chiral side chains are of particular interest for their ability to adopt stable, helical secondary structure in organic and aqueous solution [142]. These molecules are specific subclass of peptidomimetics. They are closely related to their natural peptide counterparts, but differ chemically in that their side chains are appended to nitrogen atoms along the molecule's backbone, rather than to the -carbons, as they are in amino acids (Fig. 4). Such family of molecules is essentially invulnerable to protease degradation [143] and therefore is biostable and less prone to immune system recognition [144]. Despite the achirality of the N-substituted glycine backbone and its absence of hydrogen bond donors, oligopeptoids are able to adopt stable, chiral helices when substituted with -chiral, sterically large side chains (Fig. 5). [145, 146]. Several different families of biological oligomers have been synthesized [147, 148] and proposed as novel mimics of natural molecules such as magainin, a helical, amphipathic antimicrobial peptide (Fig. 6). [127, 145, 149] Their ability to form stable helices makes peptoids an excellent candidate for mimicry of bioactive molecules that rely on helical structure for proper function. In addition, peptoid oligomers have been shown to be protease resistant and are easily and inexpensively pro-

duced by solid-phase synthesis [144, 150], traits that are critical for the successful development of a peptidomimetic therapeutic agent. In order to probe for peptoid, selective peptidomimetics of magainin-II was designed by Barron and Patch, which signified the first bioactive folded peptoid (Fig. 7) [132].
H NH2 N R O n

Fig. (4). N-Substituted glycine oligomers or polypeptoid.

Another study was done by Statz et al. due to synthesize peptoid like 6, which specifically designed for robust, waterresistant anchorage to biomaterial surfaces and long-term resistance to fouling in the biological environment. The peptidomimetic polymer was synthesized by synthesis of a 20mer N-methoxyethyl glycine peptoid (Fig. 8) [151]. N, N-Linked Oligoureas N, N-linked oligoureas with proteinogenic side chains are peptide backbone mimetics that belong to the -peptide lineage. They are formally obtained by simple substitution of NH for the CH2 of the amino acid constituents of 4peptides. By using combination of NMR spectroscopy and circular dichroism it has been shown that oligoureas, as short as seven residues can adopt a stable right- handed 2.5-helical secondary structure stabilized by 12- and 14-membered Hbonded rings in various solvents such as pyridine, methanol,

334 Anti-Infective Agents in Medicinal Chemistry, 2009, Vol. 8, No. 4

Moradi et al.
-

and trifluoromethanol [152-154] reminiscent of the helix described for the corresponding 4-peptides.
H3C

NH NH O O

Anchoring peptide

Nrpe = (R)-(N)-(1-phenylethyl)glycine
H

20-20

Npm = N-(1-phenylmethyl)glycine
O

Fig. (8). Biomimetic antifouling N-substituted glycine polymer (peptoid) containing a C-terminal peptide anchor derived from residues found in mussel adhesive proteins for robust attachment of the polymer onto surfaces.

CH3

Nme = N-(methoxyethyl)glycine

Nchm = N-(Cyclohexylmethyl)glycine Fig. (5). Large peptoid oligomers.


H N 2

This strong helix folding propensity, together with the diversity of available side chain appendages and the expected resistance to protease degradation, make the oligourea backbone a promising candidate for biomedical applications. It has been shown recently that -amino acid residues bearing side chains branched at the first carbon have a strong 314 helix propensity [155,156]. The helix stability can be significantly enhanced by acylating the free amino terminus of oligoureas [157]. Although the amide bond is undoubtedly a consensual motif for the elaboration of peptidomimetics folding oligomers ("foldamers"), it was shown that the urea moiety, by its capacity to form auto-complementarities and bidirectional hydrogen bonds can be substituted for the amide linkage to generate oligomeric strands with strong propensity for helix formation [158,159]. DESIGN AND MODELING OF PEPTIDOMIMETICS

O H N N O H C 3 N

O N O CH 3 N

H 3C O N O H 3C C H 3 N H H N N H 2 N N O H N O
14

H O N H 2

Fig. (6). The chemical structure of a peptoid 22-mer surfactant protein C, a small protein that plays an important role in surfactant replacement therapy for the treatment of neonatal respiratory distress syndrome.

Recently significant progress has been made in use of Computer-Aided Molecular Design (CAMD) of novel molecules with desired properties, these methods typically rely on two stages: the first stage is forward modeling [160], through which Quantitative Structure Activity Relationship (QSAR) process is accomplished by application of non-linear modeling procedures such as Artificial Neural Networks (ANN). As example of this strategy, Soltani et al analyzed over 100 antifungal peptides by artificial neural networks and observed that most important physicochemical parameter affected on bioactivity of antifungal peptides are Log P and relative amphipaticity [161]. The second stage is model inversion/optimization which is the use of optimization algorithms in exploitation of the first stage results in discovery of molecules with improved activity [159]. Several parameters affect the activity of antifungal peptides such as sequence, size, charge, degree of structure formation, cationicity, hydrophobicity and amphipathicity Use of ANN could help us evaluate the importance of these structural parameters in bioactivity of AFPs and eventually the most probable model of AFPs' mechanism of action. The antimicrobial peptides have little sequence homology, despite common properties [13]. Thus it is difficult to develop methods for predicting the antimicrobial peptides based on similarity. Moreover, experimental methods for identification and design of antimicrobial peptides are costly, time consuming and resource intensive.

H H N O N

O N H NH 2 O

NH 3 +

Fig. (7). Helical peptoid mimics of magainin-II amide.

Peptidomimetics and their Applications in Antifungal Drug Design

Anti-Infective Agents in Medicinal Chemistry, 2009, Vol. 8, No. 4

335

Thus there is a need to develop computational tools for predicting antifungal peptides, which could be used to design potent peptides against fungal pathogens. Recently researchers attempted to model antimicrobial peptide by QSAR and other in silico methods. For example a HMM (Hidden Markov Model) based method has been developed for searching conserved motifs of -defensin family in genome databases [157]. Combinatorial chemistry, high-throughput screening, and analogous techniques have become powerful tools to promote drug discovery in peptidomimetics research. Current pharmaceutical research has taken advantages of newer computational methods, the so-called computer-aided drug design, and other physiochemical techniques such as Xray crystallography and NMR. The main goal in rational mimetic design is to translate the structural information in the native peptide into low molecular weight non-peptidic molecules. Over the past years, many 3D structures of biological targets have been solved and have been successfully used to design new, pharmacologically useful compounds. Different computer-aided design methods, e.g. 3D pharmacophore model, 3D quantitative SAR (QSAR), docking and de novo design have been extensively used [162]. As an example for QSAR and SAR studies, we follow up this section by a SAR study of some peptidomimetics that play role of microbial inhibition, through their structures. This example significantly shows the relation between peptide structure of these substrates and their ability for inhibition. Inhibit activity of cell wall biosynthesis enzymes are categorized under antimicrobial agent family. One of the most suitable targets for these inhibitors is Glucosamine-6phosphate synthase (G6Ps), which is known as the first step enzyme of cell wall biosynthesis process. N3-(4-methoxyfumaroyl)-L-2,3 diaminopropanoic acid (FMDP), N3-(fumaramoyl)-L-2, 3-diaminopropanoic acid (FCDP) or N4-(4methoxyfumaramoyl)-L-2,4diaminobutanoic acid (FMDB) are mentioned as an instance for peptidomimetic glutamine analogues, play the role of powerful inhibitor of G6Ps (Fig. 9).
CH3 H3C O O O N H O
(1)
O CH3

absorb the metabolites of proteins. A suggested idea in order to resolve the problem is incorporating native peptide, and producing peptidomimetics prodrugs. Peptide transporters accumulate them actively and will be reactivated by intracellular peptidase action. Table 2 shows the structureantibacterial activity relationship, which have been constructed for mentioned peptidomimetic prodrugs, using E. coli K-12 Morse 2034 (trp, leu) (CGSC 5071), which is a wild-type with respect to Opp, Dpp and Tpp traits. Opp, Dpp and Tpp are peptide transporters with wellcharacterized features that let them to be distinguished from each other, but they also share certain substrate recognition features, in fact, some peptides can be taken up by more than one transporter. On the other hand, Table 3 introduces structure-penetration activity of peptidomimetc prodrugs as ligands of Dpp and Opp transporters. Dpp and Opp are typical ATP consuming transporters; each comprises four membrane proteins with a periplasmic peptide binding protein DppA or OppA, respectively. These values directly influence the total effectiveness of the mentioned antibacterial peptidomimetics because of the problem of penetration as it was described before. QSAR bioactivity modeling of antimicrobial dodeca peptide of Bac2A indicated that good activity was not solely dependent on the composition of amino acids or the overall charge or hydrophobicity, but rather required particular linear sequence patterns [163]. In a study, the construction of a mathematical model for prediction, prior to synthesis, of peptide antibacterial activity against Pseudomonas aeruginosa had been described [164]. By use of novel descriptors quantifying the contact energy between neighboring amino acids in addition to a set of inductive and conventional quantitative structure-activity relationship descriptors, it is possible to model the peptides antibacterial activity [164]. Furthermore, by use of smoothed amino acid sequence descriptors, the structural characteristics important for antimicrobial activity are determined [165]. Although there is no unified way of designing peptidomimetics, certain computer applications might be of help in this regard. Recently SuperMimic software is introduced by Goede et al. [166] which is a tool for finding potential non-peptidic building blocks that can replace or mimic parts of a protein, or peptide and conversely for identifying locations within a protein where such building blocks can be inserted. It identifies compounds that mimic parts of a protein, or positions in proteins that are suitable for inserting mimetics. Photo-switchable compounds are becoming increasingly popular for a series of biological applications based on the reversible photo-control of structure and function of biomolecules. Villin is a 92.5 kDa tissue-specific actin-binding protein associated with the actin core bundle of the brush border. Villin contains multiple gelsolin-like domains capped by a small (8.5 kDa) "headpiece" at the C-terminus consisting of a fast and independently-folding three-helix bundle that is stabilized by hydrophobic interactions. The headpiece domain is a commonly studied protein in molecular dynamics due to its small size and fast folding kinetics and short primary sequence. Recently, Fullbeck et al. [167] used Supermimic software in order to be able to induce folding of the 35-residue subdomain (H35) at its C-

H N O OH

N H

O H N N H O O OH N H R

O H 3C

O O CH3

(2)

NH H N N H O R

(3)

OH

Fig. (9). 1) FMDP general structure. "R" shows additional amino acids. 2) FCDP general structure. 3) FMDB general structure.

Unfortunately, some of these substrates are not capable to penetrate or pass the cell wall and reach target enzymes, therefore, they may not act effectively against whole cells. Thus, observations have been focused on new solutions for this problem. Microorganisms have peptide transporters to

336 Anti-Infective Agents in Medicinal Chemistry, 2009, Vol. 8, No. 4

Moradi et al.

Table 2.

Antibacterial Activities of Glutamine Analogue Peptidomimetics Against E. coli K-12 Morse 2034 (trp , leu). Activities are expressed as the amounts in nmol producing inhibition zones of 25 mm, obtained by extrapolation from plots of inhibition zone diameters versus amounts of peptidomimetics [173]. Nva stands for norvalyl (2-amino pentanoic acid); Sar stands for sarcosyl (Nmethyl glycyl)
Peptidomimetics Leu-FMDP Lys-FMDP Met-FMDP Nva-FMDP Phe-FMDP Tyr-FMDP Nva-FMDB FMDP-Ala Gly-FMDP M2034 80 180 90 60 80 200 150 700 >3000 Peptidomimetics AcNva-FMDP FMDP-Leu FMDP-Met FMDP-FMDP FCDP-Ala FCDP-Met Nva-FMDP-Nva SarNva-FMDP LysNva-FMDP M2034 >5000 >1000 >3500 >5000 >5000 >5000 230 290 150

Table 3.

Percentage of Binding Peptidomimetics to DppA and OppA. The percentage inhibition of binding to DppA and OppA were determined using Gly [125I]Tyr, and [125I2]TyrGlyGly as ligands respectively. The molar ratio of peptidomimetic to ligand was 10:1 in all cases [173]
inhibition of binding to DppA% 0 80 88 99 94 73 inhibition of binding to OppA%

Peptidomimetics

Peptidomimetics

inhibition of binding to DppA% 75 0 93 14 36 5 0 49

inhibition of binding to OppA% 24 -

Gly-FMDP SarNva-FMDP Leu-FMDP Met-FMDP Nva-FMDP Phe-FMDP Tyr-FMDP LysNva-FMDP Nva-FMDP-Nva

83 89 91

Val-FMDP AcNva-FMDP Nva-FMDB FMDP-Ala FMDP-Leu FMDP-Met Lys-FMDP FMDP-FMDP FCDP-Ala

terminus of the villin headpiece by irradiation, to replace parts of its main-chain by a photoswitch without changing the overall structure of the sub-domain. OTHER APPLICATIONS Peptidomimetics are not only applied in the field of antifungal peptides design but also has applications in other research area in order to design new and potent drugs. Following, some other applications are described. Agonists of the gonadotropin-releasing hormone (GnRH) receptor were amongst the first successful examples of using a peptidomimetic approach to drug design. As determined by structureactivity and conformational studies, the inclusion of a D-amino acid residue in the 6-position of decapeptide

agonist analogs is sufficient to induce a turn conformation e.g. in triptorelin [Trelstar, Debiopharm; Fig. (10-1)], which significantly enhances the biological effect [168]. One highly successful example currently marketed for the treatment of endocrine cancers, Gosarelin [Zoladex, Astra Zeneca; Fig. (10-2)], also includes an aza-glycine mimetic at the Cterminus to improve the stability of the molecule [169]. Koerber et al. [170], defined the active conformation of antagonist deca-peptides by the incorporation of a series of conformationally restricting cyclisations, and the activity of compounds of this type was shown to be dependent on their ability to adopt a turn structure around residues four to eight. Nuclear magnetic resonance (NMR) studies indicate that the exact location of the backbone turn may not be a key feature for receptor binding affinity, provided that the side chains of

Peptidomimetics and their Applications in Antifungal Drug Design

Anti-Infective Agents in Medicinal Chemistry, 2009, Vol. 8, No. 4

337

the critical amino acids can adopt the proper spatial orientation, which is encouraging for the design of new analogs [171]. Using a combination of turn mimetic design and directed screening, a bicyclic scaffold was discovered which, when appropriately elaborated with both hydrophobic and basic side chains, produced potent type III antagonists [172]. Refined solution conformations and SAR of these and other linear and cyclic Complement factor (C5a) antagonists modulators have paved the way for the transition between peptides and type III mimetics analogous to cases described for Somatostatin (SST). Random screening and subsequent hit optimization at Merck produced the C5a antagonist Fig. (10-3). The similarity between the guanidine, spiroindane and cyclohexyl pharmacophore of Fig. (10-3) and the Arg Trp and cyclohexylalanine side-chains of antagonist peptides such as Fig. (10-4) is indicative of some degree of type III mimicry. The rational design of compound sets that mimic the well-defined turn of the cyclic peptide derivatives appears to be a promising avenue for the development of potent and selective non-peptide C5a modulators for inflammatory diseases. Eli Lilly has patented a family of D(p-Cl)Phe DTic-like structures, of which compound Fig. (10-5) is reported to have a potency of 8.4 nM [173]. This specific small molecule peptidomimetic melanocortin-4 receptor (MC4R) agonists exhibit little effect on feeding, while having a marked effect on penile erection in rats. In this case, the design of peptidomimetic compounds has made use of the observation that the majority of known non-peptide agonists of peptide hormone receptors share close structural similarity with known antagonists of those receptors [174]. Recent advances in the development of potent and selective peptide and non-peptide ligands are anticipated to help further unravel the roles of class I and II G-protein-coupled receptors (GPCRs) in the pathogenesis of human diseases and to accelerate the clinical use of small molecule mimetics. Peptidomimetic drug discovery of GnRH antagonists, vasopressin agonists, neurotensin agonists, C5a, SST modulators and melanocortin-4 agonists illustrates the success of amalgamating the fields of conformational-based drug-design, site-directed mutagenesis, screening, smart de novo library design and classical medicinal chemistry. The recognition that discreet well-defined secondary structures are involved in the interactions of peptides with receptors has stimulated the development of molecules that mimic or stabilize such pharmacophoric descriptors. Researchers are currently developing many different non-natural oligomers for use in the mimicry of diverse bioactive peptides. Non-natural azapeptide and peptoid ligands for MHC-II are being developed to help modulate immune system response, and possibly provide therapy for auto-immune disorders. Novel somatostatin analogues incorporating peptoid and -peptide residues have been reported. Helical peptoid mimics of lung surfactant protein C are the first non-natural synthetic replacements reported. Even peptoid-based collagen mimics have been generated. Bianchini et al. [175], designed and synthesized three novel peptidomimetic phosphinate inhibitors and evaluated as inhibitors of matrix metalloproteinases MMP-2 and MMP-8 using molecular dynamics (MD) simulation and molecular modeling. Their IC50 values are in the micromolar range, and one of them showed to be the most effective in-

hibitor of MMP-2. The differences in binding affinities for MMP-2 and MMP-8 of the three phosphinates have been rationalized by means of modeling studies and MD simulations, also. In order to fully prove therapeutic benefits of Peptide T (ASTTTNYT), which is a fragment of the gp120 envelop protein of the human immunodeficiency virus (HIV), shown to inhibit binding of gp120 to the CD4, Araya et al. [176], tried to design peptidomimetics of the peptide. Using computational methods, the natural product amygdalin was identified as a prospective peptidomimetic of the peptide and later proved to exhibit a similar chemotactic profile to the peptide that led to the synthesis of a set of amygdalin analogues lacking the cyanide group with improved chemotactic profiles. Targeting farnesyltransferase (FT) enzyme has become a promising strategy in cancer therapy in last decade. Transferring a farnesyl from farnesylpyrophosphate to the thiol of a cysteine side chain of protein residues is catalyzed by FT enzyme. In order to inhibit the activity of farnesyltransferase (FT) enzyme, a genetic neural network (GNN) approach, using radial distribution function descriptors was used by Gonzalez et al. [177] to achieve this goal by a set of 78 thiol and non-thiol peptidomimetic inhibitors. The caspase (cysteinyl-aspartate protease) family represents a class of intracellular proteases, playing a critical role in apoptotic cell death pathways and activation of pro-inflammatory cytokines. Their enzymatic properties are governed by a nearly absolute specificity for substrates containing aspartic acid at the P1 site, and by the use of a cysteine side-chain for peptide-bond hydrolysis. Caspase-3 is an attractive target for therapeutic intervention due to have functions in apoptosis, mediating apoptotic cascade from the intrinsic and extrinsic activation pathways. Zhang et al. [178] designed a novel peptidomimetic inhibitor of caspase-3, which had the properties of a reversible inhibitor, while the P1 site at the C-terminal remains and only Lamino acid has been replaced by D-amino acid. Thrombosisrelated disorders such as deep vein thrombosis, pulmonary embolism, and thromboembolic stroke remain a major cause of morbidity worldwide. The limitations associated with current therapies2 have driven the search for small-molecule direct inhibitors of specific enzymes involved in the coagulation cascade. In this regard, inhibitors of both thrombin and factor Xa have attracted considerable recent attention. Nantermet et al. [179], demonstrated during their experiments that the critical hydrogen bonding motif of the established 3aminopyrazinone thrombin inhibitors can be effectively mimicked by a 2-aminopyridine N-oxide. As this peptidomimetic core was more resistant toward oxidative metabolism, it also defeated the metabolic liability associated with the pyrazinones. Smith et al. [180], developed a new class of non-peptide peptidomimetic, designed to replace the 16membered macrolide ring with a 7-membered azepine ring for attachment of the cryptophycin (isolated from terrestrial blue-green algae, show potent activity against a variety of tumor cell lines) side chains with the required spatial orientation to mimic the conformation of the relevant region of the natural product. In Type II diabetes mellitus, the most prevalent form of the disease, tissues develop resistance to the actions of insulin even though, in most instances, the insulin receptors in those tissues are structurally normal and are in near normal

338 Anti-Infective Agents in Medicinal Chemistry, 2009, Vol. 8, No. 4

Moradi et al.

H N NH H N O NH H O NH OH N NH OH O NH O NH O NH O NH O NH O

NH NH N

NH2

H NH O NH O NH2

1
H N O NH H O NH OH N NH OH O NH O NH O NH O NH O O NH O O NH O N H NH NH2 NH NH NH2

H N

NH2 HN O O N O
NH

NH

N NH

Cl HN O O
R N N H O n H N O H O H N NH O N H NH O

O S O

4: R=H, n=2

HN NH2

NH

5
Fig. (10). Five applicable peptidomimetic structures.

abundance. One strategy to combat this insulin resistance therapeutically is to maintain insulin receptors (IR) in the active tyrosine-phosphorylated form by inhibiting enzymes that catalyze IR dephosphorylation. Based on substantial evidence that protein tyrosine phosphatase 1B (PTP1B) catalyzes IR dephosphorylation and is involved physiologically and pathologically in terminating insulin signaling, this enzyme has emerged as an attractive therapeutic target. Larsen et al. [181], designed low molecular weight peptidomimetic

compounds based on O-malonyl tyrosine and Ocarboxymethyl salicylic acid are potent inhibitors of PTP1B. One member of the family of Signal Transducer and Activator of Transcription (STAT) proteins, Stat3, participates in malignant transformation. The critical role of Stat3 in the growth and survival of human tumor cells provides a valid basis for targeting Stat3 for development of novel inhibitors. Turkson et al. [182], reported novel tripeptide mimics that have been developed for improved selectivity and efficacy

Peptidomimetics and their Applications in Antifungal Drug Design

Anti-Infective Agents in Medicinal Chemistry, 2009, Vol. 8, No. 4

339

with regard to inhibition of Stat3 activity. The presence of these peptidomimetic compounds in nuclear extracts results in a dose-dependent decrease in the level of Stat3 DNAbinding activity in vitro, with efficacies that are five to ten folds higher than previously obtained for tripeptides. Hematopoietic growth factors such as G-CSF, M-CSF, GM-CSF and EPO have received considerable attention in recent years. These factors play key roles in regulating cellular host defense mechanisms involved in combating bacterial, fungal and viral infections. They stimulate host defense mechanisms by both increasing immunocompetent cell numbers and enhancing effectors cell functions. The activity of a novel series of peptidomimetic hematoregulatory compounds, designed by Heerding et al. [183] based on a pharmacophore model inferred from the structure activity relationships (SAR) of a peptide SK&F 107647 (1), is reported. These compounds induce a hematopoietic synergistic factor (HSF) which in turn modulates host defense. The compounds may represent novel therapeutic agents in the area of hematoregulatio. The binding and internalization of pathogens by host cells are critical steps in the development of many infectious diseases. Many viruses and bacteria pathogens are capable of exploiting host cell surface integrins during their replication cycles. The integrins are a family of large alpha/beta heterodimeric membrane proteins that function as cell adhesion and signal transducing molecules affecting proliferation, survival, differentiation, and migration the ligands for many integrins contain an arginineglycineaspartic acid (RGD) amino acid sequence that is essential for proteinprotein interaction. Synthetic peptidomimetics of RGD have been shown to be antagonists of the activities of specific integrins both in vitro and in vivo. Infection of cultured human cells has been shown to be inhibited by RGD-containing peptides. In a study, Hippenmeyer et al. [184], tried to understand if these small molecules are antagonists of adenovirus infection. Their results suggested that integrins interact with adenoviral RGD in a manner similar to that of other protein ligands such as vitronectin. Furthermore, the results confirm the role of RGD in the replication cycle, and suggest peptidomimetic compounds may be useful antimicrobial agents in the treatment of a variety of diseases. The phosphoprotein p53 plays a key role in maintaining the genomic integrity of cells. In response to DNA damage and other types of stress stimuli, p53 causes cell-cycle arrest1 or activates apoptosis. In normal cells, p53 is held in check until needed by the murine double-minute clone 2 (MDM2). Detrimental mutations of p53 are common mechanisms for the loss of p53 wildtype activity in cancer cells.5 But another important mechanism is over expression of MDM2, which leads to constitutive inhibition of p53; this is commonly seen in cancerous cells containing wild-type (WT) p53. The p53MDM2 complex is a target for anticancer drug design due to its importance in cancer development. It has been shown that a p53 homologue is sufficient to induce p53-dependent cell death in cells over-expressing MDM2. Recently, Zhong et al. [185], used molecular dynamics (MD) simulations in order to design a potent -peptide mimic of p53 mimic based on a tetramer of -proline, a promising peptidomimetic oligomer,

to examine the binding interface and the effect of mutating key residues in the human p53MDM2 complex. The blood coagulation cascade is divided into extrinsic and intrinsic coagulation pathways. Factor VIIa (FVIIa) in complex with tissue factor (TF) initiates the extrinsic coagulation pathway. Recent studies on blood coagulation have suggested that selective inhibition of extrinsic coagulation provides effective anticoagulation and low risk of bleeding compared with other antithrombotic mechanisms. Thus, specific FVIIa/TF complex inhibition, which blocks only extrinsic coagulation, is seen as a promising target for developing new anticoagulant drugs. Kadono and co-workers [186], showed that compound 1 (Fig. 1) with the large P3 moiety D-biphenylalanine showed relatively high selectivity against thrombin. They were able to show in other experiment that structure-based designs of the P3 moiety in the peptidomimetic factor VIIa inhibitor successfully led to novel inhibitors with selectivity for FVIIa/TF and extrinsic coagulation the same as or even higher than those of previously reported peptidomimetic factor VIIa inhibitors. CONCLUSION Considering the growing rate of fungal infections, and the limitations of current existing therapies, it is widely speculated that the direction of research in the field of developing new agents be revisited and the policies be reevaluated. The problem of resistance caused by the pathogenic and opportunistic fungal organisms is becoming common in medical practice. The treatment of patients with suboptimal dosage, delivery issues are part of the wider picture to be considered in this respect. The administered drugs, however, may not reach the target or the target itself may have changed. In instances, where the molecular design needs modification, it is possible to initiate the molecular lead discovery process to present new structural features resistant to degradation by the organism or less prone to resistance due to new targets. The antifungal peptides are part of a category of compounds belonging to the larger group of agents called antimicrobial peptides. One of the common features of this group is paucity of developed resistance which is believed to be partly due to their mechanism of action involving membranes. In this group, there are physicochemical features that make them suitable candidates for further development and evaluation. These features were analyzed by our team separately and reported before [187]. The compounds known as peptidomimetics are capable of carrying the features of AFP compounds they were originated from, without actually presenting the disadvantages usually exist in such compounds. In this article, in addition to looking at the latest developments in this field, we tried to surface the potentials of such compounds in drug discovery for infectious diseases. It is evident that the general strategy that would lead to better and more effective peptidomimetic antifungals would depend on the target, types of compound, organism involved, the cell wall and other barriers which may exist on the way of drug penetration [188]. The optimization can be carried out by in silico tools. After all, the availability of building blocks or feasibility of synthetic tools should not be neglected [189]. Although the structural complexity and the above-mentioned

340 Anti-Infective Agents in Medicinal Chemistry, 2009, Vol. 8, No. 4

Moradi et al. [23] Yang, D.; Chertov, O.; Bykovskaia, S. N.; Chen, Q.; Buffo, M. J.; Shogan, J.; Anderson, M.; Schroder, J. M.; Wang, J. M.; Howard, O. M. Z.; Oppenheim, J. J. Beta-defensins: linking innate and adaptive immunity through dendritic and T cell CCR6. Science., 1999, 286, 525-528. Wade, D.; Boman, A.; Whlin, B.; Drain, C. M.; Andreu, D.; Boman, H. G.; Merrifield, R. B. All-D amino acid-containing channel-forming antibiotic peptides. Proc. Natl. Acad. Sci. U.S.A., 1990, 87, 4761-4765. Matsuzaki, K.; Sugishita, K.; Harada, M.; Fujii, N.; Miyajima, K. Interactions of an antimicrobial peptide, magainin 2, with outer and inner membranes of Gram-negative bacteria. Biochim. Biophys. Acta, 1997, 1327, 119-130. Gennis, R. B. Molecular Structure and Function. Biomembrane, 1989, 1, 51-158. Steiner, H.; Andreu, D.; Merrifield, R. B. Binding and action of cecropin and cecropin analogues: antibacterial peptides from insects. Biochim. Biophys. Acta, 1988, 939, 260-266. Giangasparo, A.; Sandri, L.; Tossi, A. Amphipathic -helical antimicrobial peptides, a systematic study of the effects of structural and physical properties on biological activity. Eur. J. Biochem., 2001, 268, 5589-5600. Eunoh, J.; Hong S.U.; Lee, K.H. The comparison of characteristic between membrane-active antifungal peptide and its pseudo peptides. Bioorg. Med. Chem., 1999, 7, 2509-2515. Shai, Y. Mode of action of membrane active antimicrobial peptides. Biopolymers, 2002, 66, 236-248. Yang, L.; Harroun, T.A.; Weiss, T.M. L.; Ding, H.W. Barrel-stave model or toroidal model? A case study on melittin pores. Biophys. J., 2001, 81, 1475-1485. Ehrenstein, G.; Lecar, H. Q. Electrically gated ionic channels in lipid bilayers. Rev. Biophys., 1977, 10, 1-34. Banzet, N.; Lotris, M.P.; Bulet, P.; Francosis, E.; Deropierre, C.; Dubald, M. Expression of insect cystein-rich antifungal peptides in transgenic tobacco enhances resistance to a fungal disease. Plant Sci., 2002, 162, 995-1006. Jang, W.S.; Kim H.K.; Lee K.Y.; Kim S.A.; Han Y.S.; Lee, I.H. Antifungal activity of synthetic peptide derived from halocidin, antimicrobial peptide from the tunicate, Halocynthia aurantium. FEBS Lett., 2006, 580, 1490-1496. Kokryakov, V.N.; Harwig, S.S.; Panyutich, E.A.; Shevchenko, A.A.; Aleshina, G.M.; Shamova, O.; Korneva, H.A.; Lehrer, R.I. Protegrins: leukocyte antimicrobial peptides that combine features of corticostatic defensins and tachyplesins. FEBS Lett., 1993, 327(2), 23,1-6. Wang, H.X.; Ng, T.B. Concurrent isolation of a Kunitz-type trypsin inhibitor with antifungal activity and a novel lectin from Pseudostellaria heterophylla roots. Biochem. Biophys. Res. Commun., 2006, 342, 349-353. Kumar, M.; Chaturvedi, A.K.; Kavishavar, A.; Shukla, P.K.; Kesarwani, A.P.; Kundu, B. Identification of a novel antifungal nonapeptide generated by combinatorial approach. Int. J. Antimicrob. Agent., 2005, 25, 313-320. Selsted, M.E.; Brown, D.M.; DeLange, R.J.; Harwig, S.S.; Lehrer, R.I. Primary structures of six antimicrobial peptides of rabbit peritoneal neutrophils. J. Biol. Chem., 1985, 260(8), 4579-84. Skerlavaj, B.; Benincasa, M.; Risso, A.; Zanetti, M.; Renato, G. SMAP-29: a potent antibacterial and antifungal peptide from sheep leukocytes. FEBS Lett., 1999, 463, 58-62. Klotz, S.A.; Gaur, N.K.; Rauceo, J.; Lake, D.F.; Park, Y.; Hahm, S. Inhibition of Adherence and Killing of Candida albicans with a 23Mer Peptide (Fn/23) with Dual Antifungal Properties. Antimicrob. Agent Chemother., 2004, 48 (11), 4337-4341. Andra, J.; Berninghausen, O.; LEIPPE, M. Cecropins, antibacterial peptides from insects and mammals, are potently fungicidal against Candida albicans. Med. Microbiol. Immunol., 2001, 189, 169-173. Gao, G.H.; Liu, W.; Dai, J.X.; Wang, J.F.; Hu, Z.; Zhang, Y.; Wang, D.C. Molecular scaffold of a new pokeweed antifungal peptide deduced by 1H nuclear magnetic resonance. Int. J. Bio. Macromol., 2001, 29, 251-258.. Cruciani, R.A.; Barker, J.L.; Durell, S.R.; Raghunathan, G.; Guy, H.R.; Zasloff, M.; Stanley, E.F. Magainin 2, a natural antibiotic from frog skin, forms ion channels in lipid bilayer membranes. Eur. J. Pharmacol., 1992, 226(4), 287-96 Michut, L.; Fehlbaum, P.; Moniatte, M.; Dorsselaer, A.V.; Reichhart, J.M.; Bulet, P. Determination of the disulfide array of the first

aspects would make the design of such compound difficult, the development of computational tools and the internet servers provide powerful resources for further flourishing of this field. REFERENCES
[1] Rahalison, L.; Hamburger, M.; Monod, M.; Frenk, E.; Hostettman, K. A bioautographic agar overlay method for the detection of antifungal compounds from higher plants. Phytochem. Anal., 1991, 2, 199-203. Caceres, A.; Lo pez, B.R.; Giro n, M.A.; Logemann, H. Plants used in Guatemala for the treatment of dermatomucosal infections. 1: Screening of 38 plant extracts for anticandidal activity. J. Ethnopharmacol., 1991, 31, 263-276. Kovacs, J.A.; Kovacs, A.A.; Polis, M.; Wright, W.C.; Gill, V.J.; Tuazo n, C.U.; Gelmann, E.P.; Lane, H.C.; Longfield, R.; Overturf, G.; Macher, A.M.; Fauci, A.S.; Parillo, J.E.; Bennett, J.E.; Masur, H. Cryptococcosis in the acquired immunodeficiency syndrome. Ann. Int. Med., 1985, 103, 533-538. Sardari, S.; Dezfulian, M. Evaluation of SAR for Amphotericin B Derivatives by Artificial Neural Network. Trop. J. Pharm. Res ., 2005, 4, 517-521. Alcouloumre, M.S.; Ghannoum, M.A.; Ibrahim, A.S.; Selsted, M.E.; Edwards, J.E. Fungicidal properties of defensin NP-1 and activity against Cryptococcus neoformans in vitro. Antimicrob. Agents Chemother., 1993, 37, 2628-2632. Bulet, P.; Stocklin, R.; Menin, L. Anti-microbial peptides: from invertebrates to vertebrates. Immunol. Rev., 2004, 198, 169-184. Steiner, H.; Hultmark, D.; Engstrom, A.; Bennich, H.; Boman, H.G. Sequence and specificity of two antibacterial proteins involved in insect immunity. Nature, 1981, 292, 246-248. Ganz, T.; Selsted, M.E.; Szklarek, D.; Harwig, S.S.; Daher, K.; Bainton, D.F.; Lehrer, R.I. Defensins. Natural peptide antibiotics of human neuitrophils. J. Clin. Invest., 1985, 76, 1427-1435. Devine, D.A.; Hancock, R.E.W. Cationic peptides: distribution and mechanisms of resistance. Curr. Pharm. Des., 2002, 8, 703-714. Kamysz, W.; Okruj, M.; Lukasiak, J. Novel properties of antimicrobial peptides. Acta. Biochim., 2003, 50, 461-469. Baker, B., Zambryski, P.; Staskawicz, B.; Dinesh-Kumar, S.P. Signaling in plant-microbe interactions. Science, 1997, 276, 72633. Osusky, M.; Zhou, G; Osuska, L.; Hancock, R.E.; Kay, W.W.; Misra, S. Transgenic plants expressing cationic peptide chimeras exhibit broad-spectrum resistance to phytopathogens. Nat. Biotechnol., 2000, 18, 1162-1166. Loffet, A. Peptides as drugs: is there a market?. J. Pept. Sci., 2002, 8, 1-7. Travis, S.M.; Anderson, N.N.; Forsyth, W.R. Bactericidal activity of mammalian cathelicidin-derived peptides. Infect. Immun., 2000, 68, 2748-55. Hirata, M.; Shimomura, Y.; Yoshida, M. Characterization of a rabbit cationic protein (CAP18) with lipopolysaccharide-inhibitory activity. Infect. Immun., 1994, 62, 1421-6. Zanetti, M.; Gennaro, R.; Romeo, D. Cathelicidins: A novel protein family with a common proregion and a variable C-terminal antimicrobial domain. FEBS Lett., 1995, 374, 1-5. Sawa, T.; Kurahashi. K.; Ohara, M. Evaluation of antimicrobial and lipopolysaccharide-neutralizing effects of a synthetic CAP18 fragment against Pseudomonas aeruginosa in a mouse model. Antimicrob. Agents Chemother., 1998, 42, 3269-75. Hancock, R.E.W.; Chapple, D.S. Peptide antibiotics. Antimicrob. Agents Chemother., 1999, 43, 1317-1323. Yeaman, R.M.; Yount Y.N. Mechanisms of antimicrobial peptide action and resistance. Pharmacol. Rev., 2003, 55, 27-55. Shai, Y.; Oren, Z. From "carpet" mechanism to de novo designed diastereomeric cell-selective antimicrobial peptides. Peptides, 2001, 22, 1629 -1641. Yeaman, M.R.; Yount, N.Y. Mechanisms of antimicrobial peptide action and resistance. Pharmacol. Rev., 2003, 55, 27-55. Conlon, J.M.; Kolodziejek, J.; Nowotny, N. Antimicrobial peptides from ranid frogs: taxonomic and phylogenetic markers and a potential source of new therapeutic agents. Biochim. Biophys. Acta., 2004, 1696, 1-14.

[24]

[25]

[2]

[26] [27] [28]

[3]

[4] [5]

[29]

[30] [31] [32] [33]

[6] [7]

[8] [9] [10] [11]

[34]

[35]

[12]

[36]

[13] [14] [15]

[37]

[38] [39]

[16] [17]

[40]

[18] [19] [20]

[41] [42]

[21] [22]

[43]

[44]

Peptidomimetics and their Applications in Antifungal Drug Design inducible antifungal peptide from insects: drosomycin from Drosophila melanogaster. FEBS Lett., 1996, 395, 6 -10. Kim, D. H.; Lee, Y.T.; Lee, Y. J.; Chung, J. H.; Lee, B. L.; Choi, B.S.; Lee, Y. Bacterial expression of tenecin 3, an insect antifungal protein isolated from Tenebrio molitor, and its efficient purification. Mol. Cell., 1998, 8(6), 786-9 Huang, R.H.; Xiang, Y.; Liu, X.Z.; Zhang, Y.; Hu, Z.; Wang, D. C. Two novel antifungal peptides distinct with a five-disulfide motif from the bark of Eucommiaulmoides Oliv. FEBS Lett., 2002, 521, 87-90. Wang, H.; Ng, T.B. Isolation of cicadin, a novel and potent antifungal peptide from Juvenile cicadas. Peptides, 2002, 23, 7 -11. Conlon, L.M.; Kolodziejek, J.; Nowontny, N. Antimicrobial peptides from rained frogs: Taxonomic and phylogenic markers and a potential source of new therapeutic agents. Biochim. Biophys. Acta, 2004, 1696, 1-14. Ng, T.B. Antifungal proteins and peptides of leguminous and nonleguminous origin. Peptides, 2000, 25, 1215-1222. Lee, D.G.; Shin, S.Y.; Maeng, C.Y.; Jin, Z.Z.; Kim, K.L.; Hahm, K.S. Isolation and characterization of a novel antifungal peptide from Aspergillus niger. Biochem. Biophys. Res. Commun., 1999, 263, 646 -651. Lebbadi, M.; Glvez, A.; Maqueda, M.; Martnez-Bueno, M.; Valdivia, E. Fungicin M4: a narrow spectrum peptide antibiotic from Bacillus licheniformis M-4. J. Appl. Bacteriol., 1994, 77(1), 49-53. Jung, H. J.; Park, Y.; Hahm, K. S.; Lee, D.G. Biological activity of Tat (47-58) peptide on human pathogenic fungi. Biochem. Biophys. Res. Commun., 2006, 345, 222 -228. Grfe, U.; Ihn, W.; Ritzau, M.; Schade, W.; Stengel, C.; Schlegel, B.; Fleck, W. F.; Knkel, W.; Hrtl, A.; Gutsche, W. Helioferins, novel antifungal lipopeptides from Mycogone rosea: screening, isolation, structure, and biological properties. J. Antibiot., 1995, 48(2), 128-133. Klich, A.; Arthur, K. S.; Lax, A.R.; Bland, J. M. Iturin A: a potential new fungicide for stored grains. Mycopathology, 1994, 127 (2), 453-460. Fresta, M.; Ricci, M.; Rossi, C.; Furneri, P.M.; Puglisi, G. Antimicrobial Nonapeptide Leucinostatin A-Dependent Effects on the Physical Properties of Phospholipid Model Membranes. J. Coll. Inter. Sci., 2000, 226 (2), 222-230. Krainer, E.; Naider, F.; Becker, J. M. Conformational studies of nikkomycin X in aqueous solution. Biopolymers, 1990, 29 (8), 1297 -1306. Endo, A.; Kakiki, K.; Misato, T. Mechanism of Action of the Antifugal Agent Polyoxin D. J Bacteriol., 1970, 104(1), 189-196. De Lucca, A. J.; Walsh, T. J. Antifungal peptides: novel therapeutic compounds against emerging pathogens. Antimicrob. Agents Chemother., 1999, 43(1), 1-11. Fuji, E.; Fujita, Y.; Takaishi, T.; Fujita, I.; Arita, M.; Hiratsuka, N. New antibiotics, trichopolyns A and B: isolation and biological activity. Cel. Mol. Life Sci., 1978, 34(2), 253-361. Adefarati, A.A.; Hensens, O.D.; Jones, E.T.; Tkacz, J.S. Pneumocandins from Zalerion arboricola. V. Glutamic acid- and leucine-derived amino acids in pneumocandin A0 (L-671,329) and distinct origins of the substituted proline residues in pneumocandins A0 and B0. J. Antibiot., 1992, 45(12), 1953-7. Iwamoto, T.; Fujie, A.; Tsurumi, Y.; Nitta, K.; Hashimoto, S.; Okuhara, M. WF11899A, B and C, novel antifungal lipopeptides. I. Taxonomy, fermentation, isolation and physico-chemical properties. J. Antibiot., 1990, 43(9), 1183-5. Shigeru, O.; Mitsuo, H.; Sonoe, Y. Purification and properties of a new chitin-binding antifungal CB-1 from Bacillus licheniformis. Biosci. Biotechnol. Biochem., 1996, 60 (3), 481-483. Wilson, S.; Mahiou, B.; Reiger, R.; Tentler, S.; Schimmoler, R.; Orndorff, S.; Selitrennikoff, C. P. Pilot-scale purification of zeamatin, an antifungal protein from maize. Biotechnol. Prog., 2000, 16(1), 38-43. Tassin, S.; Broekaert, W. F.; Marion, D.; Acland, D. P.; Ptak, M.; Vovelle, F.; Sodano, P. Solution structure of Ace-AMP1, a potent antimicrobial protein extracted from onion seeds. Structural analogies with plant nonspecific lipid transfer proteins. Biochemistry, 1998, 37(11), 3623-37. Wang, H.X.; Ng, T.B. An antifungal peptide from red lentil seeds. Peptides, 2007, 28 (3), 547-552.

Anti-Infective Agents in Medicinal Chemistry, 2009, Vol. 8, No. 4 [66]

341

[45]

[67] [68] [69]

[46]

[47] [48]

[70]

[49] [50]

[71]

[72]

[51]

[73]

[52] [53]

[74]

[75]

[54] [55]

[76]

[77] [78] [79]

[56]

[57] [58] [59]

[80]

[60]

[81] [82]

[61]

[83]

[62]

[84] [85] [86]

[63]

[64]

[87]

[65]

[88]

Ngai, P.H.K..; Ng, TB. Coccinin, an antifungal peptide with antiproliferative and HIV-1 reverse transcriptase inhibitory activities from large scarlet runner beans. Peptides, 2004, 25, 2063-2068. Chu, K.T.; Liu, K.H.; Ng, T.B. Cicerann, a novel antifungal peptide from the green chickpea. Pep., 2003, 24, 659-663. Wong, J.H.; Ng, T.B. Gymnin, a potent defensin-like antifungal peptide from the Yunnan bean (Gymnocladus chinensis Baill). Peptides, 2003, 24, 963-968. Tanaka, H.; Satio, K.; Satio, Y.; Yamashita, T.; Agoh, M.; Okkunishi, J. Insect diapuse-specific peptide from the leaf beetle has consensus with a putative iridovirus peptide. Peptides, 2003, 24, 1327-1333. Green, L.J.; Marder, P.; Mann, L. L.; Chio, L.; Current, W. L. LY303366 exhibits rapid and potent fungicidal activity in flow cytometric assays of yeast viability. Antimicrob. Agents Chemother., 1999, 43 (4), 830-835. De Lucca, A.J.; Bland, J.M.; Grimm, C.; Jacks, T.J.; Cary, J.W.; Jaynes, J.M.; Cleveland,T.E.; Walsh.; T.J. Fungicidal and binding properties of the natural peptides cecropin B and dermaseptin. J. Microbiol., 1998, 44, 514.520. Oh, J.E.; Hong, S.U.; Lee, K.H. The comparison of characteristics between membrane-active antifungal peptide and its pseudopeptides. Bioorg. Med. Chem., 1999, 7, 2509-2515. Faber, C.; Stallmann, H. P.; Lyaruu, D. M.; Blieck, J. M. A.; Bervoets, J. M.; Wuisman, P. I. J. M. Release of antimicrobial peptide Dhvar-5 from polymethylmethacrylate beads. J. Antimicrob. Chemother., 2003, 51, 1359-1364 Drouhet, E.; Dupont, B.; Improvisi, L.; Lesourd, M.; Prevost, M. C. Activity of cilofungin (LY 121019), a new lipopeptide antibiotic, on the cell wall and cytoplasmic membrane of Candida albicans. Structural modifications in scanning and transmission electron microscopy. J. Med. Vet. Mycol., 1990, 28(6), 425-36. Lee, M.T.; Chen, F.Y.; Huang, H.W. Energetics of pore formation induced by membrane active peptides. Biochemistry, 2004, 433, 590-3599. Kim, C.; Spano, J.; Park, E.K.; Wi, S. Evidence of pores and thinned lipid bilayers induced in oriented lipid membranes interacting with the antimicrobial peptides, magainin-2 and aurein-3. BBABiomembr., 2009, 1788 (7), 1482-1496. Patch, J.A.; Barron, A.E. Mimicry of bioactive peptides via nonnatural, sequence-specific peptidomimetic oligomers. Cur. Opin. Chem. Biol., 2002, 6, 872-877 Gellman, S.H. Foldamers: A Manifesto. Acc. Chem. Res., 1998, 31, 173-180 Guichard, G. In Pseudopeptides in Drug Development; P.E. Nielsen, Ed. Weinheim, Germany: Wiley-VCH Verlag, 2004; pp. 33-120. Raguse, T.L.; Porter, E.A.; Weisblum, B.; Gellman, S.H. Structureactivity studies of 14-helical antimicrobial beta-peptides: probing the relationship between conformational stability and antimicrobial potency. J. Am. Chem. Soc., 2002, 124, 12774-12785. Yoo, B.; Kirshenbaum, K. Peptoid architectures: elaboration, actuation, and application. Curr. Opin. Chem. Biol., 2008, 12 (6), 714721. Li, L., Thomas, R.M.; Suzuki, H.; De Brabander, J.K.; Wang, X.; Harran, P.G. A small molecule Smac mimic potentiates TRAIL- and TNFalpha-mediated cell death. Science, 2004, 3, 305, 1471-74. Hruby, V. J.; Balse, P. M. Conformational and topographical considerations in designing agonist peptidomimetics from peptide leads. Curr. Med. Chem., 2000, 7, 945-970. Hruby, V. J.; Li, G.; Haskell-Luevano, C.; Shenderovich, M. Design of peptides, proteins, and peptidomimetics in chi space. Biopolymers, 1997, 43, 219-266. Ripka, A.S.; Rich, D.H. Peptidomimetic design. Cur. Opin. Chem. Biol., 1998, 2, 441-452. Eguchi, M.; Shen, R. Y. W.; Shea, J. P.; Lee, M.S.; Kahn, M. Design, synthesis, and evaluation of opioid analogues with nonpeptidic beta-turn scaffold: enkephalin and endomorphin mimetics. J. Med. Chem., 2002, 45, 1395-1398. Wang, W.; Yang, J.; Ying, J.; Xiong, C.; Zhang, J.; Cai, C.; Hruby, V. J. Stereoselective synthesis of dipeptide beta-turn mimetics: 7benzyl and 8-phenyl substituted azabicyclo[4.3.0]nonane amino acid esters. J. Org. Chem., 2002, 67, 6353-6360. Jean, F.; Buisine, E.; Melnyk, O.; Drobecq, H.; Odaert, B.; Hugues, M.; Lippens, G.; Tartar, A. Synthesis and structural and functional

342 Anti-Infective Agents in Medicinal Chemistry, 2009, Vol. 8, No. 4 evaluation of a protein modified with a beta-turn mimic. J. Am. Chem. Soc., 1998, 120, 6076-6083. Kaul, R.; Deechongkit, S.; Kelly, J. W. Synthesis of a negatively charged dibenzofuran-based beta-turn mimetic and its incorporation into the WW miniprotein-enhanced solubility without a loss of thermodynamic stability. J. Am. Chem. Soc., 2002, 124, 1190011907. Han, Y.; Giragossian, C.; Mierke, D. F.; Chorev, M. Ni-to-Ni+3ethylene-bridged partially modified retro-inverso tetrapeptide betaturn mimetic: design, synthesis, and structural characterization. J. Org. Chem., 2002, 67, 5085-5097. Szelke, M.; Leckie, B.; Hallett, A; Jones, D.M.; Sueiras, J.; Atrash, B. Potent new inhibitors of human renin. Nature, 1982, 299, 555557. Wiley, R.A.; Rich, D.H. Peptidomimetics derived from natural products. Med. Res. Rev. 1993, 13, 327-384. Maibaum, V.; Cohen, N.C.; Rijeger, H.; Gschke, .R; Maibaum, J.; Cumin, F.; Fuhrer, W.; Wood, J.M. Bioactive hydroxyethylene dipeptide isosteres with hydrophobic (P3-P1)-moieties. A novel strategy towards small non-peptide renin inhibitors. Bioorg. Med. Chem. Lett., 1996, 6, 1589-l 594. Goschke, R.; Cohen, N.C.; Wood, J.M.; Maibaum, J. Design and synthesis of novel 2,7-dialkyl substituted 5(S)-amino-4(S)hydroxy-8-phenyl-octanecarboxamides as in vitro potent peptidomimetic inhibitors of human rennin. Bioorg. Med. Chem. Lett., 1997, 7, 2735-2740. Dolle, R.E.; Prasad, C.V.C.; Prouty, C.P.; Salvino, J.M.; Awad, M.M.A.; Schmidt, S.J.; Hoyer, D.; Ross, T.M.; Graybill, T.L.; Speier, G.J. Pyridazinodiazepines as a high-affinity, P2-P3 peptidomimetic class of interleukin-1 beta-converting enzyme inhibitor. J. Med. Chem., 1997, 40, 1941-l 946. McKittrick, B.A.; Stamford, A.W.; Weng, X.; Ma, K.; Chackalamannil, S.; Czarniecki, M.; Cleven, R.M.; Fawzi, A.B. Design and synthesis of phosphinic acids that triply inhibit endothelin converting enzyme, angiotensin converting enzyme and neutral endopeptidase 24.11. Bioorg. Med. Chem. Lett., 1996, 6, 1629-l 634. Brown, D.L.; Devadas, B.; Lu, H.F.; Nagarajan, S.; Zupec, M.E.; Freeman, S.K.; McWherter, C.A.; Getman, D.P.; Sikorski, J.A. Replacements for lysine in L-seryl-L-lysyl dipeptide amide inhibitors of candida albicans myristoyl-CoA:protein N-myristoyltransferase. Bioorg. Med. Chem. Lett., 1997, 7, 379-382. Saute, M.; Rudolf, K.; Wittneben, H.; Herzog, H.; Martinez, R.; Munoz, M.; Eberlein, W.; Engel, W.; Walker, P.; Beck-Sickinger, A.G. Neuropeptide Y and the nonpeptide antagonist BIBP 3226 share an overlapping binding site at the human Y1 receptor. Pharmacology, 1996, 50, 285-292. Brickmann, K.; Yuan, Z.; Sethson, I.; Somfai, P.; Kihlberg, J. Synthesis of conformationally restricted mimetics of gamma-turns and incorporation into desmopressin, an analogue of the peptide hormone vasopressin. Chem. Eur. J., 1999, 5, 2241-2253. Yuan, Z.; Blomberg, D.; Sethson, I.; Brickmann, K.; Ekholm, K.; Johansson, B.; Nilsson, A.; Kihlberg, J. Synthesis and pharmacological evaluation of an analogue of the peptide hormone oxytocin that contains a mimetic of an inverse gamma-turn. J. Med. Chem., 2002, 45, 2512-2519. Yuan, Z.; Kihlberg, J. Synthesis of a -turn mimetic suitable for incorporation in the peptide hormone LHRH. Tetrahedron, 2005, 61, 4901-4909. Lindman, S.; Lindeberg, G.; Nyberg, F.; Karln, A.; Hallberg, A. Comparison of three -turn mimetic scaffolds incorporated into angiotensin II. Bioorg. Med. Chem., 2000, 8, 2375-2383. Fairlie, D.P.; Abbenante, G.; March, D.R. Macrocyclic peptidomimetics- forcing peptides into bioactive conformations. Curr. Med. Chem., 1995, 2, 654-686. Hanessian, S.; McNaughton-Smith, G.; Lombart, H.G.; Lubell, W.D. Design and Synthesis of Conformationally Constrained Amino Acids as Versatile Scaffolds and Peptide Mimetics. Tetrahedron, 1997, 53, 12789-l 2854. Liverton, N.J.; Armstrong, D.J; Claremon, D.A; Remy, D.C.; Baldwin, J.J.; Lynch, R.J.; Zhang, G.; Gould, R.J. Nonpeptide glycoprotein IIb/IIIa inhibitors: substituted quinazolinediones and quinazolinones as potent fibrinogen receptor antagonists. Bioorg. Med. Chem. Lett., 1998, 8, 483-486. Jones, M.R; Boatman, P.D; Semple, G.; Shin, Y.; Tamura, Y.S. Clinically validated peptides as templates for de novo pepti-

Moradi et al. domimetic drug design at G-protein-coupled receptors. J. Curr. Opin. Pharmacol., 2003, 3, 530-543. Kruijtzer, J.A.W.; Lefeber, D.J.; Liskamp, R.M.J. Approaches to the Synthesis of Ureapeptoid. Peptidomimetics. Tetrahedron Lett., 1997, 38 (30), 5335-5338, Terrett, N.K. Combinatorial Chemistry Online. J. Comb. Chem., 2008, 10 (8), 31-34. Burgess, K.; Ibarzo, J.; Linthicum, D. S.; Russell, D. H.; Shin, H.; Shitangkoon, A.; Totani, R.; Zhang, A. J. Solid Phase Syntheses of Oligoureas. J. Am. Chem. Soc, 1997, 119, 1556- 1564 Han, H.; Janda, K. D. Azatides: Solution and Liquid Phase Syntheses of a New Peptidomimetic. J. Am. Chem. Soc, 1996, 118, 25392544. Arvidsson, P.I.; Frackenpohl, J.; Ryder, N.S.; Liechty, B.; Petersen, F.; Zimmermann, H.; Camenisch, G.P.; Woessner, R.; Seebach, D. On the Antimicrobial and Hemolytic Activities of Amphiphilic beta-peptides. ChemBiochem., 2001, 2, 771-773. Smith, A. B.; Benowitz, A. B.; Favor, D. A.; Sprengeler, P. A.; Hirschmann, R. A second-generation synthesis of scalemic 3,5,5trisubstituted pyrrolin-4-ones: Incorporation of functionalized amino acid side-chains. Tetrahedron Lett., 1997,38, 3809-3812. Bont, D. B. A.; Moree, W. J.; Liskamp, R. M. Molecular diversity of peptidomimetics: Approaches to the solid-phase synthesis of peptidosulfonamides . J. Bioorg. Med. Chem., 1996, 4, 667-672. Wang, X.; Huq, I.; Rana, T. M. HIV-1 TAR RNA Recognition by an Unnatural Biopolymer. J. Am. Chem. Soc., 1997, 119, 64446445. Hall, D. G.; Schultz, P. G. Synthesis of Diverse Ethoxyformacetal Oligomers. Toward Libraries of Metal-Coordinating Unnatural Biopolymers. Tetrahedron Lett., 1997, 38, 7825-7529. Violette, A.; Fournel, S.; Lamour, L.; Chaloin, O.; Frisch, B.; Briand, J.; Monteil, H.; Guichard, G. Mimicking helical antibacterial peptides with nonpeptidic folding oligomers. J. Chem. Biol., 2006, 13, 531-538. Stephens, O.M.; Kim, S.; Welch, B.D.; Hodsdon, M.E.; Kay, M.S.; Schepartz, A. Inhibiting HIV fusion with a -peptide foldamer. J. Am. Chem. Soc., 2005, 127, 13126-13127. Kritzer, J.A.; Stephens, O.M.; Guarracino, D.A.; Reznik, S.K.; Schepartz, A. beta-peptides as inhibitors of protein-protein interactions. Bioorg. Med. Chem., 2005, 13, 11-16. Kritzer, J.A.; Lear, J.D.; Hodsdon, M.E.; Schepartz, A. Helical peptide inhibitors of the p53-hDM2 interaction. Chem. Soc., 2004, 126, 9468-9469. Klein, S.I.; Czekaj, M.; Gardner, C.J.; Guertin, K.R.; Cheney, D.L.; Spada, A.P.; Bolton, S.A.; Brown, K.; Colussi, D.; Heran, C.L. Identification and initial structure-activity relationships of a novel class of nonpeptide inhibitors of blood coagulation factor Xa. J. Med. Chem., 1998, 41, 437-450. Sall, D.J.; Bastian, J.A.; Briggs, S.L.; Buben, J.A.; Chirgadze, N.Y.; Clawson, D.K.; Denney, M.L.; Gierra, D.D.; Gifford-Moore, D.S.; Harper, R.W. Book of Abstracts. 214th American Chemical Society National Meeting: 1997 Sept 7-11: Las Vegas. Washington, DC: American Chemical Society; 1997:A58. Chirgadze, N.Y.; Sal, D.J.; Klimkowski, V.J.; Clawson, D,K.; Briggs, S.L.; Hermann, R.; Smith, G.F.; Gifford-Moore, D.S.; Werv, J.P. The crystal structure of human alpha-thrombin complexed with LY178550, a nonpeptidyl, active site-directed inhibitor. Protein Sci., 1997, 6, 1412-l 417. Hunt ,J.T.; Lee, V.G.; Leftheris, K.; Seizinaer, B.; Carboni, J.; Mabus, J.; Ricca, A.; Yan, N.; Manne V. Potent cell active not-thiol tetrapeptide inhibitors of Farnesyl transferase. J. Med. Chem., 1996, 39, 353-358. Nioroae, F.G.; Doll, R.J.; Vibulbhan, B.; Alvarez, C.S.; Bishoo, W.R.; Pktrin, J.; Kirschmeier, P.; Carruthers, N.I.; Wong, J. Discovery of Novel Nonpeptide Tricyclic Inhibitors of Ras Farnesyl Protein Transferase. Bioorg. Med. Chem., 1997, 5, 101-l 13. Bishop, W.R.; Bond, R.; Petrin, J.; Wang, L.; Patton, R.; Doll, R.; Njoroge, G.; Catino, J.; Schwartz, J.; Windsor, W. Novel tricyclic inhibitors of farnesyl protein transferase. Biochemical characterization and inhibition of Ras modification in transfected Cos cells. J. Biol. Chem., 1995, 270, 3061 l-30618. Lin, P.; Ganesan, A. Solid-phase synthesis of peptidomimetic oligomers with a phosphodiester backbone. Bioorg. Med. Chem. Lett., 1998, 8, 511-514.

[89]

[107]

[108] [109] [110]

[90]

[91] [92] [93]

[111]

[112]

[94]

[113] [114]

[95]

[115] [116]

[96]

[117]

[97]

[118] [119]

[98]

[120]

[99]

[121]

[100]

[122]

[101] [102]

[123]

[103] [104]

[124]

[125]

[105]

[126]

[106]

Peptidomimetics and their Applications in Antifungal Drug Design [127] Porter, E.A.; Wang, X.; Lee, H.S.; Weisblum, B.; Gellman, S.H. Non-haemolytic beta-amino-acid oligomers. Nature, 2000, 404 (6778), 565. Porter, E.A.; Weisblum, B.; Gellman, S.H. Mimicry of hostdefense peptides by unnatural oligomers: antimicrobial betapeptides. J. Am. Chem. Soc., 2002, 124, 7324-7330. Raguse, T.L.; Porter, E.A.; Weisblum, B.; Gellman, S.H. Structureactivity studies of 14-helical antimicrobial beta-peptides: probing the relationship between conformational stability and antimicrobial potency. J. Am. Chem. Soc., 2002, 124, 12774-12785. Epand, R.M.; Epand, R.F. Lipid domains in bacterial membranes and the action of antimicrobial agents. BBA-Biomembranes, 2009, 1788 (1), 289-294. Arvidsson, P.I.; Ryder, N.S.; Weiss, H.M.; Gross, G.; Kretz, O., Woessner, R.; Seebach, D. Antibiotic and hemolytic activity of a beta2/beta3 peptide capable of folding into a 12/10-helical secondary structure. ChemBioChem., 2003, 4, 1345-1347. Patch, J.A.; Barron, A.E. Helical peptoid mimics of magainin-2 amide. J. Am. Chem. Soc., 2003, 125, 12092-12093. Gabriel, J.G.; Som, A.; Madkour, E.A.; Eren, T.; Tew, N.T. Infectious disease: Connecting innate immunity to biocidal polymers. Mater. Sci. Eng., 2007, 57, 28-64. Cheng, R.P.; Gellman, S.H.; DeGrado, W.F. beta-Peptides: from structure to function. Chem. Rev., 2001, 101, 3219-3232. Arvidsson, P.I,; Ryder, Neil.S,; Weiss, H.M,; Gross, G,; Kretz, O,; Woessner, R,; Seebach, D. Antibiotic and hemolytic activity of a beta2/beta3 peptide capable of folding into a 12/10-helical secondary structure. ChemBioChem, 2003, 4 (12), 1345-1347. Appella, D.H.; Christianson, L.A.; Klein, D.A.; Powell, D.R.; Huang, X.; Joseph, J.; Barchi, J.; Gellman, S.H. Residue-based control of helix shape in -peptide oligomers. Nature, 1997, 387, 381-384. DeGrado, W.F,; Schneider, J.P,; Hamuro, Y . The twists and turns of beta-peptides. J. Pept. Res., 1999, 54 (3), 206-217. Liu, D.; DeGrado, W.F. De novo design, synthesis, and characterization of antimicrobial beta-peptides. J. Am. Chem. Soc., 2001, 123, 7553-7559. Tew, G.N.; Clements, D.; Tang, H.; Arnt, L.; Scott, R.W. Antimicrobial activity of an abiotic host defense peptide mimic. Biochim. Biophys. Acta Biomembr., 2006, 1758, 1387-1392. Seebach, D.; Matthews, J. L. -Peptides: a surprise at every turn. Chem. Commun., 1997, 1, 2015-2022. LePlae, P.R.; Fisk, J.D.; Porter, E.A.; Weisblum, B.; Gellman, S.H. Tolerance of acyclic residues in the beta-peptide 12-helix: access to diverse side-chain arrays for biological applications. J. Am. Chem. Soc., 2002, 124, 6820-6821. Wu, C.W.; Seurynck, S.L.; Lee, K.Y.C.; Barron, A.E. Helical Peptoid Mimics of Lung Surfactant Protein C. J. Chembiol., 2003, 10, 1057-1063. Miller, S.M.; Simon, R.J.; Ng, S.; Zuckermann, R.N.; Kerr, J.M.; Moos, W.H. Comparison of the proteolytic susceptibilities of homologous L-amino acid, D-amino acid, and N-substituted glycine peptide andpeptoid oligomers. Drug Dev. Res., 1995, 35, 20-32. Gibbons, J.A.; Hancock, A.A.; Vitt, C.R.; Knepper; S.; Buckner, S.A.; Brune, M.E.; Milicic, I.; Kerwin, J.F.; Richter, L.S.; Taylor, E.W., Spear, K.L.; Zuckermann, R.N.; Spellmeyer, D.C.; Braeckman, R.A.; Moos, W.H. Pharmacologic characterization of CHIR 2279, an N-substituted glycine peptoid with high-affinity binding for alpha 1-adrenoceptors. J. Pharmacol. Exp. Ther., 1996, 277, 885-899. Wu, C.W.; Sanborn, T.J.; Zuckermann, R.N.; Barron, A.E. Peptoid oligomers with alpha-chiral, aromatic side chains: effects of chain length on secondary structure. J. Am. Chem. Soc., 2001, 123, 29582963. Kirshenbaum, K.; Barron, A.E.; Armand, P.; Goldsmith, R.; Bradley, E.; Cohen, F.E.; Dill, K.A.; Zuckermann, R.N. Sequencespecific polypeptoids: a diverse family of heteropolymers with stable secondary structure. Proc. Natl. Acad. Sci., 1998, 95, 43034308. Shuey, S.W.; Delaney, W.J.; Shah, M.C.; Scialdone, M.A. Antimicrobial -peptoids by a block synthesis approach. Bioorg. Med. Chem. Lett., 2006, 16 (5), 1245-1248. Barron, A.E.; Zuckermann, R.N. Bioinspired polymeric materials: in-between proteins and plastics. Curr. Opin. Chem. Biol., 1999, 3, 681-687.

Anti-Infective Agents in Medicinal Chemistry, 2009, Vol. 8, No. 4 [149]

343

[128] [129]

[150]

[151]

[130]

[152]

[131]

[153]

[132] [133] [134] [135]

[154]

[155]

[136]

[156]

[137] [138]

[157]

[139] [140] [141]

[158]

[159] [160]

[142]

[161] [162] [163]

[143]

[144]

[164]

[165]

[145]

[166] [167]

[146]

[147] [148]

[168]

[169]

Wender, P.A.; Mitchell, D.J.; Pattabiraman, K.; Pelkey, E.T.; Steinman, L.; Rothbard, J.B. The design, synthesis, and evaluation of molecules that enable or enhance cellular uptake: peptoid molecular transporters. Proc. Natl. Acad. Sci., 2000, 97, 13003-13008. Zuckermann, R.N.; Kerr, J.M.; Kent, S.B.H.; Moos, W.H. Efficient Method for the Preparation of Peptoids [Oligo(N-substituted glycines)] by Submonomer Solid-Phase Synthesis.J. Am. Chem. Soc., 1992, 114, 10646-10647. Statz, A.R.; Meagher, R.J.; Barron, A.E.; Messersmith, P.B. New peptidomimetic polymers for antifouling surfaces. J. Am. Chem. Soc., 2005, 127, 7972-7973. Blondelle, S.E.; Lohner, K. Combinatorial libraries: a tool to design antimicrobial and antifungal peptide analogues having lytic specificities for structure-activity relationship studies. Bio. Polymers, 2000, 55, 74-87. Semetey, V.; Rognan, D.; Hemmerlin, C.; Graff, R., Briand; J.P.; Marraud, M.; Guichard, G. Stable helical secondary structure in short chain N,N'-linked oligoreas bearing proteinogenic side chain. Angew. Chem. Int. Ed. Engl., 2002, 41, 1893-1895. Hemmerlin, C.; Marraud, M.; Rognan, D.; Graff, R.; Semetey, V.; Briand; J.P.; Guichard, G. Helix forming oligoureas: temperaturedependent NMR, structure determination and circular dichroism of a nonomer with functionalised side chain. Helv. Chim. Acta, 2002, 85, 3692-3711. Violette, A.; Averlant-Petit, M.C.; Semetey, V.; Hemmerlin, C.; Casimir, R.; Graff, R.; Marraud, M.; Briand, J.P.; Rognan, D.; Guichard, G. N,N'-linked oligoureas as foldamers: chain length requirements for helix formation in protic solvent investigated by circular dichroism, NMR spectroscopy, and molecular dynamics. J. Am. Chem. Soc., 2005, 127, 2156-2164. Raguse, T.L.; Lai, J.R.; Gellman, S.H. Evidence that the -peptide 14-helix is stabilized by 3-residues with side-chain branching adjacent to the -carbon atom. Helv. Chim. Acta, 2002, 85, 41544164. Scheetz, T.; Bartlett, J.A.; Walters, J.D.; Schutte, B.C.; Casavant, T. L.; McCray, P. B. Genomics-based approaches to gene discovery in innate immunity. Immunol. Rev., 2002, 190, 137-145. Digilio, G; Bracco, C; Barbero, L; Chicco, D; Del Curto, MD; Esposito, P; Traversa, S; Aime, S. NMR conformational analysis of antide, a potent antagonist of the gonadotropin releasing hormone. J. Am. Chem. Soc., 2002, 124, 3431-3442. Patel, S.; Stott, I.P.; Bhakoo, M.; Elliott, P. Patenting computerdesigned peptides. J. Comp. Aided. Mol. Des., 1998, 12: 543-556. Lopez, S.F.; Kim, H.S.; Choi, E.C.; Delgado, M.; Granja, J.R.; Khasanov, A.; Kraehenbuehl, K.; Long, G.; Weinberger, D.A.; Wilcoxen, K.M.; Ghadiri, M.R. Antibacterial agents based on the cyclic D,L-alpha-peptide architecture. Nature, 2001, 412, 452-455. Soltani, S.; Keymanesh K.; Sardari, S. Evaluation of structural features of membrane acting antifungal peptides by artificial neural networks. J. Biol. Sci., 2008, 8, 834-845. Sardari, S.; Dezfulian, M. Cheminformatics in Anti-infective Agents Discovery. Mini Rev. Med. Chem., 2007, 7, 181-189. Hilpert, K.; Elliott, R. M.; Volkmer, R.; Henklein, P.; Donini, O.; Zhou, Q.; Winkler, F.H.D.; Hancock, W.R.E. Sequence requirements and an optimization strategy for short antimicrobial peptides. Chem. Biol., 2006, 13, 1101-1107. Jenssen, H.; Fjell, C.D.; Cherkasov, A.; Hancock, R.E. QSAR modeling and computer-aided design of antimicrobial peptides. J. Pept. Sci., 2007, 10, 1784-7019. Fernandez, M.; Caballero, J. Analysis of protegrin structurefunction-activity relationships: the structural characteristics important for antimicrobial activity using smoothed amino acid sequence descriptors. Mol. Simul., 2007, 33, 689-702. Goede, A.; Michalsky, E.; Schmidt, U.; Preissner, R. SuperMimic Fitting peptide mimetics into protein structures. BMC Bioinformatics, 2006, 7, 11. Fullbeck, M.; Michalsky, E.; Jaeger, I. S.; Henklein; P.; Kuhn, H.; Ruck-Braun, K.; Preissner, R. Design and biological evaluation of photo-switchable inhibitors. Genome Informatics, 2006, 17, 141151. Jones, R. M.; Boatman, D.; Semple, G.; Shin, Y. J.; Tamura, S. Y. Clinically validated peptides as templates for de novo peptidomimetic drug design at G-protein-coupled receptors. Cur. Opin. Pharmacol., 2003, 3, 530-543. Kritzer, J.A.; Tirado-Rives, J.; Hart, S.A.; Lear, J.D.; Jorgensen, W.L.; Schepartz, A. Relationship between side chain structure and

344 Anti-Infective Agents in Medicinal Chemistry, 2009, Vol. 8, No. 4 14-helix stability of beta-peptides in water. J. Am. Chem. Soc., 2005, 127, 167-178. Koerber, S.C.; Rizo, J.; Struthers, R.S.; Rivier, J.E. C. Consensus bioactive conformation of cyclic GnRH antagonists defined by NMR and molecular modeling. J. Med. Chem., 2000, 43, 819-828. Digilio, G.; Bracco, C.; Barbero, L.; Chicco, D.; Del Curto, M. D. ; Esposito, P.; Traversa. S.; Aime, S. NMR conformational analysis of antide, a potent antagonist of the gonadotropin releasing hormone. J. Am. Chem. Soc., 2002, 124, 3431-3442. Cho, N.; Harada, M.; Imaeda, T.; Imada, T.; Matsumoto, H.; Hayase, Y.; Sasaki, S.; Furuya, S.; Suzuki, N.; Okubo, S. Discovery of a novel, potent, and orally active nonpeptide antagonist of the human luteinizing hormone-releasing hormone (LHRH) receptor. J. Med. Chem., 1998, 41, 4190-4195. Neil, J.M.; Andruszkiewicz, R.; Gupta, S.; Milewski, S.; Payne, J.W. Structure-activity relationships for a series of peptidomimetic antimicrobial prodrugs containing glutamine analogues. J. Antimicrob. Chemother., 2003, 51, 821-831. Sugg, E.E. Annu. Nonpeptide agonists for peptide receptors: lessons from ligands. Rep. Med. Chem., 1997, 32, 277-283. Bianchini, G.; Aschi, M.; Cavicchio, G.; Crucianelli, M.; Preziuso, S.; Gallina, G.; Nastari, A.; Gavuzzod, E.; Mazzaa, F. Design, modelling, synthesis and biological evaluation of peptidomimetic phosphinates as inhibitors of matrix metalloproteinases MMP-2 and MMP-8. Bioorg. Med. Chem., 2005, 13, 4740-4749. Araya, E.; Rodriguez, A.; Rubio, J.; Spada, A.; Joglar, J.; Llebaria, A.; Lagunas, C.; Fernandez, A.G.; Spisanie, S.; Perez, J.J. Synthesis and evaluation of diverse analogs of amygdalin as potential peptidomimetics of peptide T. Bioorg. Med. Chem. Lett., 2005, 15, 1493-1496. Gonzalez, M.P.; Caballero, J.; Tundidor-Camba, A.; Helguerab, A.M.; Fernandezc, M. Modeling of farnesyltransferase inhibition by some thiol and non-thiol peptidomimetic inhibitors using genetic neural networks and RDF approaches. Bioorg. Med. Chem., 2006, 14, 200-213. Zhang, J.; Liu, L.; Guo, Y.; Yang, J.; Wei, L.; Zhong, D.; Wang, J. Design and Synthesis of a Novel Peptidomimetic Inhibitor of Caspase-3. Chem. Res. Chin. Univ., 2006, 22, 225-228. Nantermet, P.G.; Burgey, C.S.; Robinson, K.A.; Pellicore, J.M.; Newton, C.N.; Deng, J.Z.; Selnick, H.G.; Lewis, S.D.; Lucas, B.J.;

Moradi et al. Krueger, J.A. P2 pyridine N-oxide thrombin inhibitors: a novel peptidomimetic scaffold. Bioorg. Med. Chem. Lett., 2005, 15, 27712775. Smith, A.B.; Cho, Y.S.; Pettitb, G.R.; Hirschmanna, R. Design, synthesis, and evaluation of azepine-based cryptophycin mimetics. Tetrahedron, 2003, 59, 6991-7009. Larsen, S.D.; Stevens, F.G.; Lindberg, T.J.; Bodnar, P.M., OSullivan, T.J.; Schostarez, H.J.; Palazuk, B.J.; Bleasdaleb, J.E. Modification of the N-terminus of peptidomimetic protein tyrosine phosphatase 1B (PTP1B) inhibitors: identification of analogues with cellular activity. Bioorg. Med. Chem. Lett., 2003, 13, 971-975. Turkson, J.; Kim, J.S.; Zhang, S.; Yuan, J.; Huang, M.; Glenn, M.; Haura, E.; Sebti, S,; Hamilton, A.D.; Jove, R. Novel peptidomimetic inhibitors of signal transducer and activator of transcription 3 dimerization and biological activity. Mol. Cancer Ther., 2004, 3 (3), p.261-269. Heerding, D.A.; Abruzzese, M.; Alberts, D.; Burgess, J.; Callahan, J.F.; Huffman, W.F. Novel peptidomimetic hematoregulatory compounds. Bioorg. Med. Chem. Lett., 2000,10, 531-534. Hippenmeyer, P.J.; Ruminski, P.G.; Rico, J.G.; Lu, H.S.; Griggs, D.W. Adenovirus inhibition by peptidomimetic integrin antagonists. Antivir. Res., 2002, 55, 169-178. Zhong, H.; Carlson, H.A. Computational studies and peptidomimetic design for the human p53-MDM2 complex. Proteins, 2005, 58, 222-234. Kadono, S.; Sakamoto, A.; Kikuchi, Y.; Oheda, M.; Yabuta, N.; Yoshihashi, K.; Kitazawa, T. Structure-based design of P3 moieties in the peptide mimetic factor VIIa inhibitor. Biochem. Biophy. Res. Comm., 2005, 327, 589-596. Soltani, S.; Keymanesh, K.; Sardari, S. In Silico analysis of antifungal peptides (AFPs): Determining the lead template sequence of potent antifungal peptides. Expert Opin. Drug. Discov., 2007, 2, 837-847. Marshall, N.J.; Andruszkiewicz, R.; Gupta, S.; Milewski, S.; Payne, J.W. Structure-activity relationships for a series of peptidomimetic antimicrobial prodrugs containing glutamine analogues. J. Antimicrob. Chemother., 2003, 51, 821-831. Goodman, M; Shao, H. Peptidomimetic building blocks for drug discovery: an overview. Pure. Appl. Chem., 1996, 68, 1303-1308.

[170]

[180] [181]

[171]

[172]

[182]

[173]

[183]

[174] [175]

[184] [185]

[176]

[186]

[187]

[177]

[188]

[178] [179]

[189]

Received: June 11, 2009

Revised: August 01, 2009

Accepted: August 01, 2009

Potrebbero piacerti anche