Next Article in Journal
Subcutaneous Implantable Cardioverter Defibrillator: A Contemporary Overview
Previous Article in Journal
Obesity-Related Knee Osteoarthritis—Current Concepts
Previous Article in Special Issue
Phenotypic Characterization of Oral Mucor Species from Eurasian Vultures: Pathogenic Potential and Antimicrobial Ability
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Antimicrobial Peptides in Infectious Diseases and Beyond—A Narrative Review

by
Petros Ioannou
1,2,*,
Stella Baliou
2 and
Diamantis P. Kofteridis
1,2
1
School of Medicine, University of Crete, 71003 Heraklion, Greece
2
Internal Medicine, University Hospital of Heraklion, 71110 Heraklion, Greece
*
Author to whom correspondence should be addressed.
Life 2023, 13(8), 1651; https://doi.org/10.3390/life13081651
Submission received: 17 July 2023 / Revised: 26 July 2023 / Accepted: 27 July 2023 / Published: 28 July 2023

Abstract

:
Despite recent medical research and clinical practice developments, the development of antimicrobial resistance (AMR) significantly limits therapeutics for infectious diseases. Thus, novel treatments for infectious diseases, especially in this era of increasing AMR, are urgently needed. There is ongoing research on non-classical therapies for infectious diseases utilizing alternative antimicrobial mechanisms to fight pathogens, such as bacteriophages or antimicrobial peptides (AMPs). AMPs are evolutionarily conserved molecules naturally produced by several organisms, such as plants, insects, marine organisms, and mammals, aiming to protect the host by fighting pathogenic microorganisms. There is ongoing research regarding developing AMPs for clinical use in infectious diseases. Moreover, AMPs have several other non-medical applications in the food industry, such as preservatives, animal husbandry, plant protection, and aquaculture. This review focuses on AMPs, their origins, biology, structure, mechanisms of action, non-medical applications, and clinical applications in infectious diseases.

1. Introduction

Despite significant recent developments in medical research and clinical practice, therapeutics for infectious diseases are greatly limited by the emergence of antimicrobial resistance (AMR) [1,2,3]. AMR is one of the top threats in medicine and directly threatens society with increasing rates of morbidity, mortality, and healthcare costs [4,5,6,7,8,9,10]. For example, more than 2.8 million antimicrobial-resistant infections occur each year in the USA, resulting in more than 35,000 deaths [5]. AMR is also associated with a high cost for healthcare and society. For example, it is estimated that by 2050, the approximate annual worldwide cost of AMR could be between USD 300 billion and 1 trillion [7,11].
From a clinical perspective, infections by antibiotic-resistant bacteria develop in patients with previous antimicrobial use, recent hospitalization, or those who live in long-term care facilities [12,13,14,15,16]. The mechanism of development of AMR involves the exposure of bacteria in the patient’s microbiome to antibiotics, the elimination of bacteria that are sensitive to the used antibiotic, and the selection of those who are resistant [17,18,19]. Then, the resistant bacteria could be transmitted to other people via direct or indirect contact. This could lead to the spread of AMR in the hospital environment and long-term care facilities [12,13,14,15,16]. The development of antimicrobial resistance involves several mechanisms, such as limiting the uptake of antibiotics, the enzymatic inactivation of the drug, modification of the drug target, or the production of efflux pumps that will lead to the removal of the antibiotic [17,20,21]. Importantly, AMR is not only associated with using antimicrobials in humans. For example, an increased concentration of antimicrobials in industrially polluted water, untreated hospital effluent, municipal sewage, and in rivers and seas may also drive the development of AMR in environmental bacteria [22,23,24,25,26]. Beyond the selective pressure of antibiotics on susceptible microorganisms, other mechanisms are associated with the spread of AMR. These involve the mobilization and transfer of genetic elements from resistant microorganisms to susceptible ones [22,27]. The microorganisms spreading the AMR could be of nosocomial or environmental origin, such as those thriving in the external environment, in animals, or humans [22,27].
Among the more-common antimicrobial-resistant microorganisms are carbapenem-resistant Acinetobacter, carbapenem-resistant Enterobacteriaceae (CRE), methicillin-resistant Staphylococcus aureus (MRSA), and vancomycin-resistant Enterococcus (VRE) [5,28,29]. Currently, the clinical practice of infectious diseases and clinical microbiology is greatly limited because of AMR. The existing therapeutic approach involves using parenteral antimicrobials as single drugs or in combination, targeting the offending microorganism. Treatment is usually started empirically and modified according to culture and antimicrobial susceptibility results. Guidelines published by international and local infectious disease societies aid in selecting the suitable treatment, even for multidrug and extensively drug-resistant microorganisms [30,31,32].
However, in many cases, especially regarding pan-drug-resistant A. baumannii, Pseudomonas, or Klebsiella, no current available single antimicrobial agent retains in vitro susceptibility. In that case, treatment involves providing antimicrobial combinations that have in vitro synergy in an attempt to reduce morbidity and mortality [33,34,35]. Due to the mediocre activity of this approach, and the lack of viable antimicrobial treatment options, there is an unmet need for developing newer therapeutic options for these difficult-to-treat pathogens. Even though the antimicrobial pipeline contains several tested antimicrobials, very few complete all necessary steps toward approval for clinical practice. Even in that case, however, up until now, most of these antimicrobials belong to already existing antimicrobial classes [36,37]. Even so, unfortunately, resistance quickly develops, thus leaving few or no options for treatment [38,39,40]. To tackle antimicrobial resistance, the discovery of new antibiotics is not enough. Infection control practices, along with antimicrobial stewardship interventions, are also required. Infection control practices could lead to a reduction in the transmission of resistant microorganisms between humans. These measures are primarily in practice in healthcare settings, involving, among other things, appropriate guidance on disinfection and contact, droplet, or airborne isolation, depending on the pathogen [41,42]. On the other hand, antimicrobial stewardship interventions allow for the better use of antimicrobials by reducing the unnecessary exposure of patients to these drugs. The optimization of antimicrobial use through educational approaches, formulary restriction, solicited or even unsolicited infectious disease consultations, and prospective audits and feedback could lead to reduced AMR without compromising patient safety [43,44,45].
Novel treatments for infectious diseases in the era of increasing AMR are urgently needed. To that end, there is ongoing research in non-classical therapies for infectious diseases, utilizing other antimicrobial mechanisms to fight pathogens, such as bacteriophages or antimicrobial peptides [36,46,47,48,49]. This review focuses on antimicrobial peptides (AMPs) and the ongoing research regarding their development. Moreover, it provides a comprehensive overview of clinical studies where AMPs are used in the fight against infectious diseases, as well as other non-medical applications of AMPs.

2. Search Strategy

For the conduction of the present review, PubMed and Scopus were searched with the term ‘antimicrobial peptides’ until 20 June 2023 for original articles providing information on biology, clinical trials, and other medical and non-medical applications of AMPs, as well as for informative reviews on the topic. The resulting articles’ references were also searched to identify additional relevant articles. The extracted information included data on the biology of AMPs, their mechanism of action, non-medical applications, and clinical data from trials in humans.

3. Definitions and Origins of AMPs

AMPs are evolutionarily conserved molecules produced naturally by several organisms aiming to combat pathogenic microorganisms. They are an essential component of the innate immune system of several species, such as insects, amphibians, marine organisms, and even mammals. At the same time, plants and bacteria can also produce AMPs [50]. Indeed, invertebrates that lack an adaptive immune system mainly depend on AMPs to fight invading pathogens and eliminate infections [51,52]. Lysozyme, probably the first AMP ever described, was discovered about 100 years ago by Alexander Fleming; however, most of the discoveries in the field of AMPs were made after 1990 [53]. Until now, thousands of AMPs have been described, with some of them being natural and some of them being synthetic [54].
AMPs may be effective against various microorganisms, causing the rapid killing of Gram-positive and Gram-negative microorganisms, fungi, parasites, viruses, or even neoplastic cells. They may even be able to inhibit the growth of antimicrobial-resistant microorganisms with mechanisms that are distinct from those of currently used antimicrobials [55,56,57,58,59]. For example, persulcatusin isolated from ticks (Ixodes persulcatus) has antimicrobial activity against MRSA and vancomycin-resistant S. aureus (VRSA) [60,61]. Another example of AMPs that are clinically in use, and are of increasing importance in the era of AMR, is polymyxins—cationic polypeptides bound to fatty acids, [62]. They act by binding to the negatively charged part of the lipopolysaccharide (LPS), thus leading to permeability changes in the bacterial cell membrane and cell death [63,64]. Furthermore, many AMPs exert immunomodulatory activity, thereby promoting pathogen eradication in an indirect manner [65,66].

4. Structure and Function of AMPs

AMPs generally consist of 10 to 50 amino-acid residues (even though they could have as few as 5 and up to more than 100 amino-acid residues), have a positive net charge, and typically contain a significant number of hydrophobic residues (in general, about 50%) as well as hydrophilic segments, thus possessing amphiphilic characteristics [54,55,67,68,69,70]. They are classified according to their source and origin (from prokaryotes, protista, fungi, plants, and animals), their biological activity (against bacteria, viruses, parasites, fungi, insects, and tumors), their characteristics (such as charge and length), or other factors, such as their chemical modifications, their 3D structure, their binding, and their mechanism of action [71]. The classification of AMPs has to do with their secondary structure, according to which they are divided into three categories: a-helical, b-sheet, or extended/random-coil structure, with most AMPs belonging to the first two categories [56,68,70,72,73]. However, it is notable that the structure may depend on the identity of the AMP and the surrounding environment. Moreover, their structure may vary upon interaction with zwitterion and anionic membranes, thus making this classification problematic [74,75].
As stated initially in architecture, the following quote also applies to biology: “form ever follows function” [76]. Indeed, the mechanism of action of AMPs is closely related to their structural characteristics and physical features, which are mainly characterized by a net positive charge and hydrophobicity [70,77,78]. The presence of many positively charged amino-acid residues such as lysine and arginine provides a positive net charge to AMPs, allowing them to interact with the negatively charged outer cell membrane of bacteria while avoiding interaction with the uncharged eukaryotic cell membrane [79]. Hence, efforts to improve the activity of AMPs aim to replace amino-acid residues with positively charged ones [80,81]. An example is reflected in a study where cathelicidin-BF15 was modified by replacing amino-acid residues, leading to the production of ZY13—an AMP with improved antimicrobial activity, increased stability, and reduced hemolytic potential [82]. Mounting evidence suggests that increasing net cationic charge correlates to the increased antimicrobial activity of AMPs. However, this may also be associated with an increase in hemolytic activity [83,84,85,86,87]. On the other hand, the hydrophobicity of AMPs is a critical characteristic, as most of their amino-acid residues are hydrophobic. This allows AMPs to interact with cell membranes, even though it also adds to their hemolytic activity, as AMPs that are highly hydrophobic penetrate erythrocytic membranes to a higher extent [88,89].
The mechanism of action of AMPs has been a focus of research for many years, since a better understanding of the way these molecules act could allow for the optimization of their function by modifying their primary, secondary, and tertiary structure [70,79,90,91]. AMPs can be categorized based on their mechanism of action into those that kill by acting on the membrane causing disruption and those that act by non-membrane-disruptive mechanisms. The membrane targeting of AMPs can be accomplished through receptor association or without receptors. AMPs cross the membrane after achieving a particular threshold concentration and perturb bacterial membrane permeability through pore opening. More specifically, AMPs that act by membrane-disrupting mechanisms do so by forming a toroidal pore or a barrel-stave, or via a carpet-like mechanism, with all these causing membrane disruption that leads to the leak of intracellular contents, thus causing cell death (Figure 1) [70,92]. In all cases, AMPs accumulate and are appropriately organized on the target cell membrane, a process driven by their hydrophobic nature and the electrostatic interaction of their positive net charge with the negatively charged phospholipids on the cell membrane of the target cell [93]. The peptide–lipid ratio exerts a determining effect on the delivery of AMPs on the bacterial membrane. For example, AMPs are transferred in parallel into the bacterial membrane at a low peptide–lipid ratio [94].
In contrast, AMPs penetrate vertically into the bacterial membrane in the case of a high peptide–lipid ratio. In some instances, AMPs bind to receptors, exerting their efficacy in vitro in the nanomolar range [95]. Alternatively, AMPs display their bacterial potential through interactions with membrane components other than receptors in vitro at micromolar concentrations, thereby causing membrane disruption [96]. The toroidal pore mechanism disrupts the target cell membrane by perpendicularly inserting AMPs into the lipid bilayer. The carpet-like mechanism involves the accumulation of AMPs on the target cell membrane, leading to the penetration of the membrane. Finally, the barrel-stave mechanism, which is more typical for the AMPs that have a helical secondary structure, involves the formation of hydrophilic pores on the target cell membrane, which leads to the release of its intracellular content [51,93,97,98]. AMPs that act via non-membrane-disrupting mechanisms target microbial processes that do not involve the membrane and may resemble the action of classic antimicrobials, such as DNA synthesis, protein synthesis, specific enzyme inhibition, cell wall biosynthesis, inhibiting cell division, inhibiting RNA or DNA functions, and triggering apoptosis through reactive oxygen species (ROS) generation (Figure 1) [55,99,100,101,102].
Furthermore, beyond causing direct microorganism killing, AMPs may also have immunomodulatory properties (Figure 2). They are secreted by many different innate immune system cells, such as neutrophils and macrophages. They can intervene in the inflammatory milieu, leading to a more-controlled secretion of proinflammatory cytokines, preventing excessive tissue damage. At the same time, they also promote angiogenesis and reduce the excessive production of ROS [66,103,104,105]. More specifically, it has been suggested that AMPs are significant factors that attract and activate different populations of immune cells (e.g., polymorphonuclear cells or macrophages) through chemokines. Moreover, AMPs may suppress the expression of several pro-inflammatory cytokines, exert anti-endotoxin activity, and enhance the microbial killing mechanisms of immune cells, such as phagocytosis or the production of neutrophil extracellular traps (NETs) [106,107]. In other cases, AMPs exert their action through the alteration of signaling pathways such as p38, extracellular signal-regulated kinase 1/2 (Erk1/2), c-Jun N-terminal kinase JNK mitogen-activated protein (MAP)-kinases, Nuclear factor kappa-light-chain-enhancer of activated B cells (NF-kB), or Phosphoinositide 3-kinase (PI3K)/Protein kinase B (Akt), independently of bacterial ligand binding [108]. On the other hand, some AMPs, such as LL37, can induce auto-inflammation through the recognition of self-DNA by TLR9, self-RNA by TLR7/8, or double-stranded RNA by TLR3 [106]. In this regard, LL-37 has been reported to be harmful, allowing mast cell degranulation and consequent histamine release, which in turn leads to vasodilation [109] through the activation of signaling pathways such as PI3K/Akt, Erk1/2, and JNK [110]. Examples of the mechanism of action of AMPs are shown in Table 1 and in more detail in other studies [111,112].

5. The Physiological Role of AMPs

Physiologically, AMPs are either produced continuously or readily synthesized in response to a perceived microbial threat of the host by the upregulation of the corresponding genes. As an example, in humans, psoriasin, lactoferrin, and dermcidin are continuously present in the skin, while cathelicidin LL-37 is readily produced in response to infection [128,129]. In vertebrates, AMPs have the potential to kill microorganisms directly; however, they can also modulate the immune system by activating and recruiting cells of the immune system during infection [70,97,130]. Many different types of AMPs have been described from cells of the immune system, secretions, and epithelia of animals and amphibians [131,132]. More than 30 cathelicidins have been identified and described in mammals. They are stored in inactive forms in granules of the neutrophils so that after neutrophil activation, they can be released in an active state, contributing to microorganism eradication [133,134]. Being produced in other vertebrates, not only in mammals, several AMP classes other than cathelicidins have been described and have been reported to exert their mechanisms of action, such as modulation of the immune response, reduction of inflammation, and tissue damage, as well as the destruction of invading microorganisms [135,136,137,138,139]. In terms of antimicrobial activity, AMPs are very effective against bacteria, viruses, and fungi [139,140].
Insects and plants also produce AMPs, and these are of increased significance for these organisms since they lack an adaptive immune response and they only rely on the response of their innate immune system to combat invading pathogens [70]. Many AMPs have been isolated and described from insects, and more specifically from their hemolymphs, hemocytes, and epithelial cells [141,142,143]. In plants, leaves, seeds, roots, and tubers of plants have been found to produce AMPs, with defensins and thionins being some notable examples [103,144,145]. They display high thermal, proteolytic, and chemical stability due to the multiple disulfide bonds they contain due to the numerous cysteine residues [103,144,145,146]. The mechanism of action of AMPs produced by plants involves the disruption of the cell membrane and pore formation, the targeting of protein synthesis, and also the targeting of the DNA synthesis of pathogens [147]. Some of the AMPs produced by insects are defensins, cecropins, diptericins, and drosocins, and they exert similar activity against Gram-positive and Gram-negative microorganisms and fungi. One example of their antibacterial mechanism of action is associated with their cationic nature, which leads to the permeabilization of the bacterial cell membranes [148,149,150,151].
AMPs are also crucial for bacteria since their production allows them to adapt and survive in antagonistic environments where competition with other microorganisms in specific environmental conditions threatens their existence [97]. Many AMPs have been isolated from bacteria and present significant antimicrobial activity against Gram-positive, Gram-negative microorganisms and fungi. These AMPs, also called bacteriocins, are physiologically produced by those bacteria to inhibit or kill microorganisms that antagonize them without causing toxicity to the bacterium that produces them, thus providing a survival benefit [70,152].

6. Applications of AMPs

6.1. Food Industry

One of the most promising applications of AMPs is their potential as alternative regimens to antimicrobial treatment. This will be described in detail later on. Other non-medical uses include their use in the food industry as preservatives [153,154]. Such examples include nisin, an AMP consisting of 34 amino-acid residues, which is isolated from Lactococcus lactis and used as a preservative in skim and whole-fat milk, cottage cheese, milk putting and other dairy products; enterocin, which is isolated from Enterococcus faecium, aureocin, an AMP isolated from S. aureus; bovicin, which is isolated from Streptococcus bovis; reuterin, an AMP isolated from Lactobacillus reuteri; and others that are mainly used as preservatives in dairy products [154]. Lactoferrin is another example of an AMP, commonly used as an antimicrobial agent in the USA to preserve meat products. At the same time, its derivative, lactoferricin, is an AMP with more potent antimicrobial activity and relative heat resistance, which provides higher food preservation capacity [155,156]. Another example of an AMP used as a food preservative is natamycin, commonly applied on the surface of cheese and salami-type sausages. Natamycin is produced by Streptomyces spp. and has potent antimicrobial activity against the majority of food-borne fungi even though it has no activity against bacteria or viruses [157]. Spheniscin is a defensin produced in the stomach of king penguins and preserves undigested food for future use in the egg incubation period. Thus, it was hypothesized that it could be of use as a food preservative [158].

6.2. Animal Husbandry

AMPs are also crucial in animal husbandry. Due to their potential for a strong antimicrobial effect with a low likelihood of resistance, it was considered that they could maintain sustainable livestock production. For example, the dietary supplementation of chickens with nisin changed the intestinal microbiome of chickens, leading to fewer counts of Enterobacterales and Bacteroides [159]. One step further, transgenic animals, exploiting genetic manipulation techniques, can produce AMPs that can optimize their production and reduce the possibility of infection. For example, the expression of bovine or human lactoferricin in the mammary glands of transgenic goats provided a wide range of antimicrobial activities [160]. Dietary supplementation with AMPs to reduce outbreaks of animal disease and improve growth performance and animal health in animal husbandry was proposed as an alternative to dietary supplementation with antimicrobials until it was banned by the European Union [156,161]. The addition of AMPs such as colistin E1, cipB-lactoferricin-lactoferrampin, and cecropin AD to the diets of weanling pigs led to an improvement of growth and health condition by beneficially altering the gut microbiome [162,163,164]. Another vital use of AMPs in the context of animal husbandry has to do with their antiviral activities, such as against porcine epidemic diarrhea virus, infectious bronchitis virus (IBV), porcine transmissible gastroenteritis virus, and influenza A [165,166,167,168]. Hence, AMPs could be used in animals to reduce the dangers associated with a viral infection, as exemplified in the case of the infection of chick embryos treated with swine intestine AMP, which has potent antiviral activity against IBV [166].

6.3. Plant Protection

Plant protection is another non-medical application of AMPs. AMPs could replace pesticides in the fight against plant pathogens and insects, which could allow for the reduction of environmental pollution and the minimization of risk for human health harms associated with the use of pesticides [169]. AMPs such as iseganan or pexiganan display important activity against phytopathogenic bacteria such as Pectobacterium spp., while other AMPs have important antifungal activities that could be used for plant protection [156,170]. The genetic manipulation of plants leading to the recombinant expression of AMPs in their bodies could lead to resistance to phytopathogens. For example, the gene expressing Alf-AFP defensin was introduced in the potato, leading to greater protection against Verticillium dahliae [171]. Furthermore, the expression of the mammalian AMP cecropin P1 in transgenic tobacco led to increased plant resistance towards E. carotovora, Pseudomonas marginata, and Pseudomonas syringae [172]. Finally, when an analog of magainin, MSI-99, was expressed in different plants, it provided protection against many bacterial and fungal phytopathogens [173,174,175].

6.4. Aquaculture

Finally, AMPs have also been used in aquaculture. Since fish and other sea products are essential components of the human diet, aquaculture is a growing finance sector. However, outbreaks of microbial disease can be associated with a significant delay in production and critical financial losses [156]. To that end, the synthetic AMP epinecidin-1 could hinder the growth of multiple bacteria that are known to be harmful to aquacultural organisms, such as Pasteurella multocida, E. coli, Aeromonas spp., and Vibrio spp. [176]. In another case, several AMPs like caerin 1.1, or sole NKLP27, had potent antimicrobial activity against most human and fish pathogenic bacteria. At the same time, they were also able to inactivate many fish viruses that are important in aquaculture [177].

7. Potential for the Utilization of AMPs in Infectious Diseases

AMPs could be used alone or in combination with conventional antimicrobials against microorganisms, including difficult-to-treat (DTT) microorganisms, such as multi-drug-resistant (MDR) or extensively drug-resistant (XDR) bacteria. They have the potential for rapidly killing microorganisms with a low possibility of the further development of resistance, which makes them ideal candidates for the fight against resistant pathogens [178,179,180]. An essential difference between AMPs and classical antimicrobials is that the latter target one single biological process. In contrast, AMPs usually exert multiple effects that may involve the disruption of the microbial cell membrane, transcription, translation, protein synthesis, and DNA synthesis [70,181,182,183]. Based on naturally occurring AMPs, synthetic AMPs with optimized pharmacological properties have been designed. These molecules already have increased potency against DTT microorganisms—an approach more effective than the current pharmacological approach of producing classic antimicrobial agents with broad-spectrum activity that is susceptible to the rapid development of AMR [70,182].

8. Examples of AMPs with Potential Interest in Infectious Diseases

Many AMPs have been extensively studied to identify whether they could be used in clinical practice in infectious diseases, either in their natural form or after modifications. AMPs belonging to the cathelicidin group are found in animals and humans. They could, theoretically, be suitable for clinical use due to their broad-spectrum antimicrobial activity against bacteria, viruses, and fungi and their immunomodulatory potential [133,184]. For example, cathelicidin LL-37 is one of the most-studied AMPs, exhibiting potent antimicrobial activity that spans bacteria and fungi [185]. LL-37 has been used as a basis for the design of synthetic AMPs with improved antimicrobial activity against DTT pathogens and biofilms [186,187]. For example, SAAP-148, an AMP subjected to modifications of the C-terminal chain of LL-37, has been reported to exert significant microbicidal activity against many DTT pathogens such as A. baumannii, P. aeruginosa, members of Enterobacterales and Gram-positive microorganisms such as E. faecium and S. aureus [188]. Notably, this AMP was also shown to be active against biofilms of some of these pathogens, such as A. baumannii, P. aeruginosa, and S. aureus, and was able to kill without the selection of resistance [188].
AMPs approved for clinical use nowadays due to their antimicrobial action include cyclic peptides such as polymyxins and gramicidins. Polymyxins were mainly used locally to treat ocular infections and also for the selective decontamination of the gastrointestinal tract. However, the increased AMR of Gram-negative microorganisms has led to a revival of the systematic use of this drug for DTT pathogens such as A. baumannii, P. aeruginosa, and Enterobacterales [64,189]. Gramicidins cause hemolysis to a significant extent, and due to this limitation, they are not used systemically but are primarily administered locally for the treatment of superficial wound infections, as well as throat, nasal and ocular infections [190]. Daptomycin, on the other hand, is a cyclic AMP already approved for systematic use, for the treatment of skin and soft tissue infections by Gram-positive bacteria [191].
Several other AMPs targeting multiple microorganisms, such as pexiganan and omiganan, are under evaluation in clinical trials. Pexiganan has been studied as a topical cream in phase III clinical trials for treating infection in diabetic patients with ulcers. Moreover, there are planned studies for the evaluation of its use in complicated skin and soft tissue infections [56,191,192]. Omiganan, an analog of indolicidin with activity against Gram-positive and Gram-negative bacteria and fungi, has been examined in clinical trials as a local treatment for infections associated with catheters, genital warts, and other conditions [56,191,193].
Other AMPs have also been studied in clinical studies of other conditions, related to their antimicrobial effect. For example, in a study in patients with periodontitis, AMP Nal-P-143 was found to be able to inhibit periodontal pathogens such as Fusobacterium nucleatum, Streptococcus gordonii, Treponema denticola, and Porphyromonas gingivalis and improve periodontal status [194].

9. Comparison of AMPs with Classic Antibiotics

There are several advantages of AMPs compared to classic antibiotics that encourage using them as therapeutic options. First of all, the primary clinical challenge of traditional antibiotics is the development of resistance. The potential of bacteria to convey resistance to AMPs is shallow, given that they are usually released during immune responses to combat microbial infections [195]. For example, a systematic study has reported that resistance against AMPs after point mutations and gene amplification was negligible. In contrast, cross-resistance between AMPs and antibiotic-resistant bacteria has not been observed [196]. Moreover, the mechanisms of the killing of AMPs are unique and may bypass AMR mechanisms. At the same time, their potential to activate the immune system provides an extra function not elaborated by classic antibiotics.
Another important antibacterial mechanism of AMPs is their biofilm inhibitory property [197]. AMPs hold great promise as a new-generation antimicrobial treatment. For these reasons, the Food and Drug Administration (FDA) has proceeded to approve three AMPs, while three more are being examined for potential clinical application [198]. Along with the above, several reports have highlighted that AMPs are less susceptible to developing host concentration-dependent toxicity. Significantly, AMPs can attenuate sepsis by neutralizing endotoxins, as shown in both in vitro and in vivo settings [199,200]. In contrast, antibiotics can potentiate septic shock due to the increased secretion of pathogen-associated molecular patterns (PAMPs), thereby aggravating infection severity [201,202]. Last but not least, the good thermal stability and water solubility of AMPs, their reduced bacterial mutagenesis, their subjection to bioengineering techniques, and their potential use in immunocompromised individuals constitute other advantages that support their clinical application compared to conventional antibiotics [54].
Despite the reported advantages of AMPs, several shortcomings of AMPs hinder their clinical application, such as high extraction costs [203], poor bioavailability [204], short half-lives [205], cytotoxicity, and lack of specificity [206]. Regarding AMPs’ toxicity, it is notable that toxicity can be either cellular or systemic [207]. The cellular cytotoxicity originates from their adopted amphiphilic structure because AMPs interact in a non-specific manner [51]. The systemic toxicity of AMPs can be attributed to the following: either sustained immune response or the uncontrolled function of the central nervous system due to their penetration of the blood–brain barrier or obstruction of blood vessels, leading to blood coagulation [207]. As a result, the topical administration of AMPs may be the safest mode of delivery to increase the benefit-to-harm ratio. Another last major limiting factor is their possible instability due to their degradation by proteases [208,209].
To address the challenges above, several strategies have been developed to improve efficacy. Chemical modifications have been proposed to circumvent the potential protease cleavage of AMPs. Specifically, the use of D-amino acids, cyclization, acetylation, dimerization, methylation, and peptidomimetics are some suggested chemical modifications of AMPs that sustain their antimicrobial potential while maintaining their stability [210,211,212,213]. On the other hand, the incorporation of AMPs into drug nano-delivery vehicles contributes to the excellent bioavailability of AMPs, as indicated by their low kidney clearance and their consequent increased retention [214], enabling their controlled release to target sites. Apart from this, nano-delivery systems offer improved pharmacokinetic properties, safety, and stability to AMPs, thereby increasing their efficacy [215,216]. In particular, several nanostructures have been deployed for the delivery of AMPs, including liposomes, micelles, dendrimers, liquid crystalline systems, hydrogels, polymeric nanoparticles, microspheres, metal nanocrystalline materials, carbon nanotubes, quantum dots, mesoporous silica nanoparticles and nano-fibers [217].

10. Combination of AMPs with Classic Antibiotics or Other Molecules

Due to the challenges encountered using a single antimicrobial, especially in pathogens with significant AMR [218,219], the combination of AMPs with either AMPs or antibiotics emerges as a promising solution (Table 2 and Table 3) [220]. The rationale for combining AMPs involves the additivity or synergy of their molecular mechanisms due to the opening of bacterial pores for extended periods, the blocking of pore repair, or disturbing bacterial intracellular functions to a greater extent. Similarly, AMPs could increase the therapeutic efficacy of antibiotics through the better delivery of antibiotics into bacterial cells, granting them access to their intracellular targets. The effectiveness of antimicrobial synergies may also involve targeting multiple independent signaling pathways in microorganisms. For example, combinations of antimicrobials present strong antimicrobial efficacy, circumventing the potential of bacterial drug resistance and possible concentration-associated toxicity [221]. Last, such synergistic activities of AMPs can also arise with other compounds or histones (Table 4 and Table 5). Several examples of AMP combinations that display a positive synergistic effect are described in Table 2, Table 3, Table 4 and Table 5.

11. Examples of AMP Use in Clinical Practice

Although AMPs are abundant, few have been selected through preclinical studies to proceed to clinical trials for potential use in humans. One such example is daptomycin. Daptomycin is a cyclic 13-member lipopeptide that is produced by Streptomyces roseosporus and was discovered in the 1980s [260]. The drug was not approved for clinical use initially since it was used twice daily, leading to muscle toxicity, despite its clinical efficacy. After changing the dose to once daily, the drug was successfully granted application, as the adverse events were drastically reduced [260]. Its mechanism of action is unclear compared to other antibiotics, but it resembles that of a classic AMP. It targets the bacterial cell membrane in a calcium-dependent manner, preferably at the division septum, and disrupts the architecture of the membrane after the oligomerization of the drug [261,262,263]. Randomized clinical trials have evaluated the activity of daptomycin in acute bacterial skin and skin structure infections (ABSSSIs) in a once-daily dose compared to the standard of care (vancomycin and antistaphylococcal penicillins) and showed they were comparable [264,265]. Thus, the FDA approved daptomycin for clinical use for the treatment of Gram-positive cocci, such as S. aureus (including strains resistant to methicillin), Enterococcus faecalis, and streptococci. In contrast, currently, this drug is used for many other infections caused by Gram-positive microorganisms [266,267,268].
Colistin is a cationic AMP and belongs to the family of polymyxins [269]. These compounds are among the oldest antibiotics, since they were discovered in 1947, and they were used for infections by Gram-negative microorganisms until other potent antimicrobials with fewer toxicities were discovered and licensed for clinical use, such as aminoglycosides [260]. Their mechanism of action involves the penetration of the cell membranes of Gram-negative bacteria, electrostatic interaction with phospholipids, and the disruption of membranes. Moreover, they can bind to the cell wall of lipopolysaccharide (LPS) and block some of its biological functions [270]. In recent decades, clinicians started using colistin again due to the emergence of XDR pathogens, such as P. aeruginosa or A. baumannii [33,269,271]. To that end, colistin has been successfully used in patients with highly resistant pathogens in many different infections and has shown adequate activity either as monotherapy or, more commonly, in the context of treatment with combinations of antimicrobials [272,273,274].
It is of note, though, that few AMPs are currently licensed for clinical use in humans. The majority of AMPs evaluated either in vitro or in vivo may not proceed to clinical studies, or if they do, a minority of them may be granted authorization for clinical use. Table 6 shows a limited number of examples of AMPs that have proceeded to clinical trials in humans, along with the main study results.

12. Conclusions

The increasing prevalence of AMR, with the high morbidity and mortality caused by antibiotic-resistant microorganisms, has led to the need to intensify the production of new antibiotics and identify new weapons in the fight against infectious diseases. To that end, AMPs emerge as valuable tools in infectious diseases, even though they have multiple other applications beyond medicine. Several AMPs are already in use in infectious diseases, such as daptomycin, colistin, and bacitracin; several others are under evaluation for potential use in the future. AMPs may have a vital role alone or in combination with other therapeutic options, such as classic antibiotics already in use. The ability of AMPs to directly kill bacteria without being affected by AMR adds an essential mechanism of action for clinicians caring for patients with infections. The results of studies involving AMPs with promising in vitro activity should be confirmed in well-designed, blind, randomized, controlled clinical trials that will evaluate their safety and efficacy in real-world situations, thus allowing their approval by regulatory authorities.

Author Contributions

Conceptualization, P.I.; methodology, P.I.; software, P.I.; validation, S.B. and D.P.K.; writing—original draft preparation, P.I. and S.B.; writing—review and editing, D.P.K.; visualization, P.I.; supervision, P.I.; project administration, P.I. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Murray, C.J.L.; Ikuta, K.S.; Sharara, F.; Swetschinski, L.; Aguilar, G.R.; Gray, A.; Han, C.; Bisignano, C.; Rao, P.; Wool, E.; et al. Antimicrobial Resistance Collaborators Global Burden of Bacterial Antimicrobial Resistance in 2019: A Systematic Analysis. Lancet 2022, 399, 629–655. [Google Scholar] [CrossRef] [PubMed]
  2. Mestrovic, T.; Aguilar, G.R.; Swetschinski, L.R.; Ikuta, K.S.; Gray, A.P.; Weaver, N.D.; Han, C.; E Wool, E.; Hayoon, A.G.; Hay, S.I.; et al. The Burden of Bacterial Antimicrobial Resistance in the WHO European Region in 2019: A Cross-Country Systematic Analysis. Lancet Public Health 2022, 7, e897–e913. [Google Scholar] [CrossRef] [PubMed]
  3. Tarín-Pelló, A.; Suay-García, B.; Pérez-Gracia, M.-T. Antibiotic Resistant Bacteria: Current Situation and Treatment Options to Accelerate the Development of a New Antimicrobial Arsenal. Expert Rev. Anti Infect. Ther. 2022, 20, 1095–1108. [Google Scholar] [CrossRef] [PubMed]
  4. EClinicalMedicine. Antimicrobial Resistance: A Top Ten Global Public Health Threat. EClinicalMedicine 2021, 41, 101221. [Google Scholar] [CrossRef]
  5. Centers for Disease Control and Prevention. Antibiotic Resistance Threats in the United States; Centers for Disease Control and Prevention: Atlanta, GA, USA, 2019. [Google Scholar]
  6. Colomb-Cotinat, M.; Lacoste, J.; Brun-Buisson, C.; Jarlier, V.; Coignard, B.; Vaux, S. Estimating the Morbidity and Mortality Associated with Infections Due to Multidrug-Resistant Bacteria (MDRB), France, 2012. Antimicrob. Resist. Infect. Control 2016, 5, 56. [Google Scholar] [CrossRef] [Green Version]
  7. Dadgostar, P. Antimicrobial Resistance: Implications and Costs. Infect. Drug Resist. 2019, 12, 3903–3910. [Google Scholar] [CrossRef] [Green Version]
  8. Pulingam, T.; Parumasivam, T.; Gazzali, A.M.; Sulaiman, A.M.; Chee, J.Y.; Lakshmanan, M.; Chin, C.F.; Sudesh, K. Antimicrobial Resistance: Prevalence, Economic Burden, Mechanisms of Resistance and Strategies to Overcome. Eur. J. Pharm. Sci. 2022, 170, 106103. [Google Scholar] [CrossRef]
  9. Kaier, K.; Wilson, C.; Chalkley, M.; Davey, P.G.; Suetens, C.; Grundmann, H.; de Kraker, M.; Schumacher, M.; Wolkewitz, M.; Frank, U. Health and Economic Impacts of Antibiotic Resistance in European Hospitals—Outlook on the BURDEN Project. Infection 2008, 36, 492–494. [Google Scholar] [CrossRef]
  10. Naylor, N.R.; Atun, R.; Zhu, N.; Kulasabanathan, K.; Silva, S.; Chatterjee, A.; Knight, G.M.; Robotham, J.V. Estimating the Burden of Antimicrobial Resistance: A Systematic Literature Review. Antimicrob. Resist. Infect. Control 2018, 7, 58. [Google Scholar] [CrossRef] [Green Version]
  11. Chokshi, A.; Sifri, Z.; Cennimo, D.; Horng, H. Global Contributors to Antibiotic Resistance. J. Glob. Infect. Dis. 2019, 11, 36–42. [Google Scholar] [CrossRef]
  12. Jian, Z.; Zeng, L.; Xu, T.; Sun, S.; Yan, S.; Yang, L.; Huang, Y.; Jia, J.; Dou, T. Antibiotic Resistance Genes in Bacteria: Occurrence, Spread, and Control. J. Basic Microbiol. 2021, 61, 1049–1070. [Google Scholar] [CrossRef] [PubMed]
  13. Lerminiaux, N.A.; Cameron, A.D.S. Horizontal Transfer of Antibiotic Resistance Genes in Clinical Environments. Can. J. Microbiol. 2019, 65, 34–44. [Google Scholar] [CrossRef]
  14. Munita, J.M.; Arias, C.A. Mechanisms of Antibiotic Resistance. Microbiol. Spectr. 2016, 4, 464–473. [Google Scholar] [CrossRef] [Green Version]
  15. Darby, E.M.; Trampari, E.; Siasat, P.; Gaya, M.S.; Alav, I.; Webber, M.A.; Blair, J.M.A. Molecular Mechanisms of Antibiotic Resistance Revisited. Nat. Rev. Microbiol. 2022, 21, 280–295. [Google Scholar] [CrossRef] [PubMed]
  16. Vandenbroucke-Grauls, C.M.J.E.; Kluytmans, J.A.J.W. Tracing the Origins of Antibiotic Resistance. Nat. Med. 2022, 28, 638–640. [Google Scholar] [CrossRef]
  17. Reygaert, W.C. An Overview of the Antimicrobial Resistance Mechanisms of Bacteria. AIMS Microbiol. 2018, 4, 482–501. [Google Scholar] [CrossRef]
  18. Sandoval-Motta, S.; Aldana, M. Adaptive Resistance to Antibiotics in Bacteria: A Systems Biology Perspective. Wiley Interdiscip. Rev. Syst. Biol. Med. 2016, 8, 253–267. [Google Scholar] [CrossRef] [PubMed]
  19. Blanquart, F.; Lehtinen, S.; Lipsitch, M.; Fraser, C. The Evolution of Antibiotic Resistance in a Structured Host Population. J. R. Soc. Interface 2018, 15, 20180040. [Google Scholar] [CrossRef] [Green Version]
  20. Uddin, T.M.; Chakraborty, A.J.; Khusro, A.; Zidan, B.R.M.; Mitra, S.; Emran, T.B.; Dhama, K.; Ripon, M.K.H.; Gajdács, M.; Sahibzada, M.U.K.; et al. Antibiotic Resistance in Microbes: History, Mechanisms, Therapeutic Strategies and Future Prospects. J. Infect. Public. Health 2021, 14, 1750–1766. [Google Scholar] [CrossRef]
  21. Christaki, E.; Marcou, M.; Tofarides, A. Antimicrobial Resistance in Bacteria: Mechanisms, Evolution, and Persistence. J. Mol. Evol. 2020, 88, 26–40. [Google Scholar] [CrossRef]
  22. Larsson, D.G.J.; Flach, C.-F. Antibiotic Resistance in the Environment. Nat. Rev. Microbiol. 2022, 20, 257–269. [Google Scholar] [CrossRef] [PubMed]
  23. Karkman, A.; Do, T.T.; Walsh, F.; Virta, M.P.J. Antibiotic-Resistance Genes in Waste Water. Trends Microbiol. 2018, 26, 220–228. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Pazda, M.; Kumirska, J.; Stepnowski, P.; Mulkiewicz, E. Antibiotic Resistance Genes Identified in Wastewater Treatment Plant Systems—A Review. Sci. Total Environ. 2019, 697, 134023. [Google Scholar] [CrossRef] [PubMed]
  25. Lin, Z.; Yuan, T.; Zhou, L.; Cheng, S.; Qu, X.; Lu, P.; Feng, Q. Impact Factors of the Accumulation, Migration and Spread of Antibiotic Resistance in the Environment. Environ. Geochem. Health 2021, 43, 1741–1758. [Google Scholar] [CrossRef] [PubMed]
  26. Shi, X.; Xia, Y.; Wei, W.; Ni, B.-J. Accelerated Spread of Antibiotic Resistance Genes (ARGs) Induced by Non-Antibiotic Conditions: Roles and Mechanisms. Water Res. 2022, 224, 119060. [Google Scholar] [CrossRef]
  27. Bengtsson-Palme, J.; Kristiansson, E.; Larsson, D.G.J. Environmental Factors Influencing the Development and Spread of Antibiotic Resistance. FEMS Microbiol. Rev. 2018, 42, fux053. [Google Scholar] [CrossRef]
  28. De Oliveira, D.M.P.; Forde, B.M.; Kidd, T.J.; Harris, P.N.A.; Schembri, M.A.; Beatson, S.A.; Paterson, D.L.; Walker, M.J. Antimicrobial Resistance in ESKAPE Pathogens. Clin. Microbiol. Rev. 2020, 33, e00181-19. [Google Scholar] [CrossRef]
  29. Mancuso, G.; Midiri, A.; Gerace, E.; Biondo, C. Bacterial Antibiotic Resistance: The Most Critical Pathogens. Pathogens 2021, 10, 1310. [Google Scholar] [CrossRef]
  30. Paul, M.; Carrara, E.; Retamar, P.; Tängdén, T.; Bitterman, R.; Bonomo, R.A.; de Waele, J.; Daikos, G.L.; Akova, M.; Harbarth, S.; et al. European Society of Clinical Microbiology and Infectious Diseases (ESCMID) Guidelines for the Treatment of Infections Caused by Multidrug-Resistant Gram-Negative Bacilli (Endorsed by European Society of Intensive Care Medicine). Clin. Microbiol. Infect. 2022, 28, 521–547. [Google Scholar] [CrossRef]
  31. Tamma, P.D.; Aitken, S.L.; Bonomo, R.A.; Mathers, A.J.; van Duin, D.; Clancy, C.J. Infectious Diseases Society of America Guidance on the Treatment of Extended-Spectrum β-Lactamase Producing Enterobacterales (ESBL-E), Carbapenem-Resistant Enterobacterales (CRE), and Pseudomonas Aeruginosa with Difficult-to-Treat Resistance (DTR-P. Aeruginosa). Clin. Infect. Dis. 2021, 72, e169–e183. [Google Scholar] [CrossRef]
  32. Magiorakos, A.-P.; Srinivasan, A.; Carey, R.B.; Carmeli, Y.; Falagas, M.E.; Giske, C.G.; Harbarth, S.; Hindler, J.F.; Kahlmeter, G.; Olsson-Liljequist, B.; et al. Multidrug-Resistant, Extensively Drug-Resistant and Pandrug-Resistant Bacteria: An International Expert Proposal for Interim Standard Definitions for Acquired Resistance. Clin. Microbiol. Infect. 2012, 18, 268–281. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Karakonstantis, S.; Ioannou, P.; Samonis, G.; Kofteridis, D.P. Systematic Review of Antimicrobial Combination Options for Pandrug-Resistant Acinetobacter Baumannii. Antibiotics 2021, 10, 1344. [Google Scholar] [CrossRef] [PubMed]
  34. Karakonstantis, S.; Kritsotakis, E.I.; Gikas, A. Treatment Options for K. Pneumoniae, P. Aeruginosa and A. Baumannii Co-Resistant to Carbapenems, Aminoglycosides, Polymyxins and Tigecycline: An Approach Based on the Mechanisms of Resistance to Carbapenems. Infection 2020, 48, 835–851. [Google Scholar] [CrossRef] [PubMed]
  35. Karakonstantis, S.; Gikas, A.; Astrinaki, E.; Kritsotakis, E.I. Excess Mortality Due to Pandrug-Resistant Acinetobacter Baumannii Infections in Hospitalized Patients. J. Hosp. Infect. 2020, 106, 447–453. [Google Scholar] [CrossRef] [PubMed]
  36. Theuretzbacher, U.; Outterson, K.; Engel, A.; Karlén, A. The Global Preclinical Antibacterial Pipeline. Nat. Rev. Microbiol. 2020, 18, 275–285. [Google Scholar] [CrossRef] [Green Version]
  37. Butler, M.S.; Gigante, V.; Sati, H.; Paulin, S.; Al-Sulaiman, L.; Rex, J.H.; Fernandes, P.; Arias, C.A.; Paul, M.; Thwaites, G.E.; et al. Analysis of the Clinical Pipeline of Treatments for Drug-Resistant Bacterial Infections: Despite Progress, More Action Is Needed. Antimicrob. Agents Chemother. 2022, 66, e0199121. [Google Scholar] [CrossRef]
  38. Malik, S.; Kaminski, M.; Landman, D.; Quale, J. Cefiderocol Resistance in Acinetobacter Baumannii: Roles of β-Lactamases, Siderophore Receptors, and Penicillin Binding Protein 3. Antimicrob. Agents Chemother. 2020, 64, e01221-20. [Google Scholar] [CrossRef]
  39. Wang, Y.; Wang, J.; Wang, R.; Cai, Y. Resistance to Ceftazidime-Avibactam and Underlying Mechanisms. J. Glob. Antimicrob. Resist. 2020, 22, 18–27. [Google Scholar] [CrossRef]
  40. Gaibani, P.; Giani, T.; Bovo, F.; Lombardo, D.; Amadesi, S.; Lazzarotto, T.; Coppi, M.; Rossolini, G.M.; Ambretti, S. Resistance to Ceftazidime/Avibactam, Meropenem/Vaborbactam and Imipenem/Relebactam in Gram-Negative MDR Bacilli: Molecular Mechanisms and Susceptibility Testing. Antibiotics 2022, 11, 628. [Google Scholar] [CrossRef]
  41. Mitchell, B.G.; Fasugba, O.; Russo, P.L. Where Is the Strength of Evidence? A Review of Infection Prevention and Control Guidelines. J. Hosp. Infect. 2020, 105, 242–251. [Google Scholar] [CrossRef]
  42. Masterton, R.G.; Mifsud, A.J.; Rao, G.G. Hospital Isolation Precautions Working Group Review of Hospital Isolation and Infection Control Precautions. J. Hosp. Infect. 2003, 54, 171–173. [Google Scholar] [CrossRef] [PubMed]
  43. Rice, L.B. Antimicrobial Stewardship and Antimicrobial Resistance. Med. Clin. N. Am. 2018, 102, 805–818. [Google Scholar] [CrossRef] [PubMed]
  44. Lanckohr, C.; Bracht, H. Antimicrobial Stewardship. Curr. Opin. Crit. Care 2022, 28, 551–556. [Google Scholar] [CrossRef]
  45. Dyar, O.J.; Huttner, B.; Schouten, J.; Pulcini, C. ESGAP (ESCMID Study Group for Antimicrobial stewardshiP) What Is Antimicrobial Stewardship? Clin. Microbiol. Infect. 2017, 23, 793–798. [Google Scholar] [CrossRef] [Green Version]
  46. Dedrick, R.M.; Guerrero-Bustamante, C.A.; Garlena, R.A.; Russell, D.A.; Ford, K.; Harris, K.; Gilmour, K.C.; Soothill, J.; Jacobs-Sera, D.; Schooley, R.T.; et al. Engineered Bacteriophages for Treatment of a Patient with a Disseminated Drug-Resistant Mycobacterium Abscessus. Nat. Med. 2019, 25, 730–733. [Google Scholar] [CrossRef] [PubMed]
  47. Schmidt, C. Phage Therapy’s Latest Makeover. Nat. Biotechnol. 2019, 37, 581–586. [Google Scholar] [CrossRef]
  48. Leuck, A.-M.; Ahmad, A.; Dunny, G.M. Combination Bacteriophage and Antibiotic Treatment for In Vitro Enterococcal Biofilms on Left Ventricular Assist Device Drivelines. Open Forum Infect. Dis. 2016, 3, 2048. [Google Scholar] [CrossRef] [Green Version]
  49. Ioannou, P.; Baliou, S.; Samonis, G. Bacteriophages in Infectious Diseases and Beyond—A Narrative Review. Antibiotics 2023, 12, 1012. [Google Scholar] [CrossRef]
  50. Hazam, P.K.; Goyal, R.; Ramakrishnan, V. Peptide Based Antimicrobials: Design Strategies and Therapeutic Potential. Prog. Biophys. Mol. Biol. 2019, 142, 10–22. [Google Scholar] [CrossRef]
  51. Brogden, K.A. Antimicrobial Peptides: Pore Formers or Metabolic Inhibitors in Bacteria? Nat. Rev. Microbiol. 2005, 3, 238–250. [Google Scholar] [CrossRef]
  52. Gueguen, Y.; Bernard, R.; Julie, F.; Paulina, S.; Delphine, D.-G.; Franck, V.; Philippe, B.; Evelyne, B. Oyster Hemocytes Express a Proline-Rich Peptide Displaying Synergistic Antimicrobial Activity with a Defensin. Mol. Immunol. 2009, 46, 516–522. [Google Scholar] [CrossRef] [Green Version]
  53. Lemaitre, B.; Nicolas, E.; Michaut, L.; Reichhart, J.M.; Hoffmann, J.A. The Dorsoventral Regulatory Gene Cassette Spätzle/Toll/Cactus Controls the Potent Antifungal Response in Drosophila Adults. Cell 1996, 86, 973–983. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Moghadam, M.T.; Mojtahedi, A.; Moghaddam, M.M.; Fasihi-Ramandi, M.; Mirnejad, R. Rescuing Humanity by Antimicrobial Peptides against Colistin-Resistant Bacteria. Appl. Microbiol. Biotechnol. 2022, 106, 3879–3893. [Google Scholar] [CrossRef] [PubMed]
  55. Hancock, R.E.W.; Sahl, H.-G. Antimicrobial and Host-Defense Peptides as New Anti-Infective Therapeutic Strategies. Nat. Biotechnol. 2006, 24, 1551–1557. [Google Scholar] [CrossRef] [PubMed]
  56. Mahlapuu, M.; Björn, C.; Ekblom, J. Antimicrobial Peptides as Therapeutic Agents: Opportunities and Challenges. Crit. Rev. Biotechnol. 2020, 40, 978–992. [Google Scholar] [CrossRef]
  57. Giovati, L.; Ciociola, T.; Magliani, W.; Conti, S. Antimicrobial Peptides with Antiprotozoal Activity: Current State and Future Perspectives. Future Med. Chem. 2018, 10, 2569–2572. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Brice, D.C.; Diamond, G. Antiviral Activities of Human Host Defense Peptides. Curr. Med. Chem. 2020, 27, 1420–1443. [Google Scholar] [CrossRef]
  59. Fernández de Ullivarri, M.; Arbulu, S.; Garcia-Gutierrez, E.; Cotter, P.D. Antifungal Peptides as Therapeutic Agents. Front. Cell Infect. Microbiol. 2020, 10, 105. [Google Scholar] [CrossRef]
  60. Miyoshi, N.; Isogai, E.; Hiramatsu, K.; Sasaki, T. Activity of Tick Antimicrobial Peptide from Ixodes Persulcatus (Persulcatusin) against Cell Membranes of Drug-Resistant Staphylococcus Aureus. J. Antibiot. 2017, 70, 142–146. [Google Scholar] [CrossRef] [Green Version]
  61. Miyoshi, N.; Saito, T.; Ohmura, T.; Kuroda, K.; Suita, K.; Ihara, K.; Isogai, E. Functional Structure and Antimicrobial Activity of Persulcatusin, an Antimicrobial Peptide from the Hard Tick Ixodes Persulcatus. Parasites Vectors 2016, 9, 85. [Google Scholar] [CrossRef] [Green Version]
  62. Landman, D.; Georgescu, C.; Martin, D.A.; Quale, J. Polymyxins Revisited. Clin. Microbiol. Rev. 2008, 21, 449–465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Velkov, T.; Thompson, P.E.; Nation, R.L.; Li, J. Structure—Activity Relationships of Polymyxin Antibiotics. J. Med. Chem. 2010, 53, 1898–1916. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Falagas, M.E.; Kasiakou, S.K. Colistin: The Revival of Polymyxins for the Management of Multidrug-Resistant Gram-Negative Bacterial Infections. Clin. Infect. Dis. 2005, 40, 1333–1341. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Yeung, A.T.Y.; Gellatly, S.L.; Hancock, R.E.W. Multifunctional Cationic Host Defence Peptides and Their Clinical Applications. Cell Mol. Life Sci. 2011, 68, 2161–2176. [Google Scholar] [CrossRef]
  66. Hancock, R.E.W.; Nijnik, A.; Philpott, D.J. Modulating Immunity as a Therapy for Bacterial Infections. Nat. Rev. Microbiol. 2012, 10, 243–254. [Google Scholar] [CrossRef]
  67. Chen, C.H.; Lu, T.K. Development and Challenges of Antimicrobial Peptides for Therapeutic Applications. Antibiotics 2020, 9, 24. [Google Scholar] [CrossRef] [Green Version]
  68. Pasupuleti, M.; Schmidtchen, A.; Malmsten, M. Antimicrobial Peptides: Key Components of the Innate Immune System. Crit. Rev. Biotechnol. 2012, 32, 143–171. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Lei, J.; Sun, L.; Huang, S.; Zhu, C.; Li, P.; He, J.; Mackey, V.; Coy, D.H.; He, Q. The Antimicrobial Peptides and Their Potential Clinical Applications. Am. J. Transl. Res. 2019, 11, 3919–3931. [Google Scholar]
  70. Mwangi, J.; Hao, X.; Lai, R.; Zhang, Z.-Y. Antimicrobial Peptides: New Hope in the War against Multidrug Resistance. Zool. Res. 2019, 40, 488–505. [Google Scholar] [CrossRef]
  71. Chung, P.Y.; Khanum, R. Antimicrobial Peptides as Potential Anti-Biofilm Agents against Multidrug-Resistant Bacteria. J. Microbiol. Immunol. Infect. 2017, 50, 405–410. [Google Scholar] [CrossRef]
  72. Nguyen, L.T.; Haney, E.F.; Vogel, H.J. The Expanding Scope of Antimicrobial Peptide Structures and Their Modes of Action. Trends Biotechnol. 2011, 29, 464–472. [Google Scholar] [CrossRef]
  73. Takahashi, D.; Shukla, S.K.; Prakash, O.; Zhang, G. Structural Determinants of Host Defense Peptides for Antimicrobial Activity and Target Cell Selectivity. Biochimie 2010, 92, 1236–1241. [Google Scholar] [CrossRef]
  74. Travkova, O.G.; Moehwald, H.; Brezesinski, G. The Interaction of Antimicrobial Peptides with Membranes. Adv. Colloid Interface Sci. 2017, 247, 521–532. [Google Scholar] [CrossRef]
  75. Priyadarshini, D.; Ivica, J.; Separovic, F.; de Planque, M.R.R. Characterisation of Cell Membrane Interaction Mechanisms of Antimicrobial Peptides by Electrical Bilayer Recording. Biophys. Chem. 2022, 281, 106721. [Google Scholar] [CrossRef] [PubMed]
  76. Sullivan, L. The Tall Office Building Artistically Considered. Lippincott’s Mag. 1896, 57, 403–409. [Google Scholar]
  77. Matsuzaki, K. Why and How Are Peptide-Lipid Interactions Utilized for Self-Defense? Magainins and Tachyplesins as Archetypes. Biochim. Biophys. Acta 1999, 1462, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Dathe, M.; Wieprecht, T. Structural Features of Helical Antimicrobial Peptides: Their Potential to Modulate Activity on Model Membranes and Biological Cells. Biochim. Biophys. Acta 1999, 1462, 71–87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Guilhelmelli, F.; Vilela, N.; Albuquerque, P.; da Derengowski, L.S.; Silva-Pereira, I.; Kyaw, C.M. Antibiotic Development Challenges: The Various Mechanisms of Action of Antimicrobial Peptides and of Bacterial Resistance. Front. Microbiol. 2013, 4, 353. [Google Scholar] [CrossRef] [Green Version]
  80. Yeaman, M.R.; Yount, N.Y. Mechanisms of Antimicrobial Peptide Action and Resistance. Pharmacol. Rev. 2003, 55, 27–55. [Google Scholar] [CrossRef] [Green Version]
  81. Arias, M.; Piga, K.B.; Hyndman, M.E.; Vogel, H.J. Improving the Activity of Trp-Rich Antimicrobial Peptides by Arg/Lys Substitutions and Changing the Length of Cationic Residues. Biomolecules 2018, 8, 19. [Google Scholar] [CrossRef] [Green Version]
  82. Jin, L.; Bai, X.; Luan, N.; Yao, H.; Zhang, Z.; Liu, W.; Chen, Y.; Yan, X.; Rong, M.; Lai, R.; et al. A Designed Tryptophan- and Lysine/Arginine-Rich Antimicrobial Peptide with Therapeutic Potential for Clinical Antibiotic-Resistant Candida Albicans Vaginitis. J. Med. Chem. 2016, 59, 1791–1799. [Google Scholar] [CrossRef] [PubMed]
  83. Gagnon, M.-C.; Strandberg, E.; Grau-Campistany, A.; Wadhwani, P.; Reichert, J.; Bürck, J.; Rabanal, F.; Auger, M.; Paquin, J.-F.; Ulrich, A.S. Influence of the Length and Charge on the Activity of α-Helical Amphipathic Antimicrobial Peptides. Biochemistry 2017, 56, 1680–1695. [Google Scholar] [CrossRef] [PubMed]
  84. Hong, S.Y.; Park, T.G.; Lee, K.H. The Effect of Charge Increase on the Specificity and Activity of a Short Antimicrobial Peptide. Peptides 2001, 22, 1669–1674. [Google Scholar] [CrossRef]
  85. Lyu, Y.; Yang, Y.; Lyu, X.; Dong, N.; Shan, A. Antimicrobial Activity, Improved Cell Selectivity and Mode of Action of Short PMAP-36-Derived Peptides against Bacteria and Candida. Sci. Rep. 2016, 6, 27258. [Google Scholar] [CrossRef] [Green Version]
  86. Chen, Y.; Mant, C.T.; Farmer, S.W.; Hancock, R.E.W.; Vasil, M.L.; Hodges, R.S. Rational Design of Alpha-Helical Antimicrobial Peptides with Enhanced Activities and Specificity/Therapeutic Index. J. Biol. Chem. 2005, 280, 12316–12329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Jiang, Z.; Vasil, A.I.; Hale, J.D.; Hancock, R.E.W.; Vasil, M.L.; Hodges, R.S. Effects of Net Charge and the Number of Positively Charged Residues on the Biological Activity of Amphipathic Alpha-Helical Cationic Antimicrobial Peptides. Biopolymers 2008, 90, 369–383. [Google Scholar] [CrossRef]
  88. Chen, Y.; Guarnieri, M.T.; Vasil, A.I.; Vasil, M.L.; Mant, C.T.; Hodges, R.S. Role of Peptide Hydrophobicity in the Mechanism of Action of Alpha-Helical Antimicrobial Peptides. Antimicrob. Agents Chemother. 2007, 51, 1398–1406. [Google Scholar] [CrossRef] [Green Version]
  89. Tachi, T.; Epand, R.F.; Epand, R.M.; Matsuzaki, K. Position-Dependent Hydrophobicity of the Antimicrobial Magainin Peptide Affects the Mode of Peptide-Lipid Interactions and Selective Toxicity. Biochemistry 2002, 41, 10723–10731. [Google Scholar] [CrossRef]
  90. Hancock, R.E.; Lehrer, R. Cationic Peptides: A New Source of Antibiotics. Trends Biotechnol. 1998, 16, 82–88. [Google Scholar] [CrossRef]
  91. Ulm, H.; Wilmes, M.; Shai, Y.; Sahl, H.-G. Antimicrobial Host Defensins—Specific Antibiotic Activities and Innate Defense Modulation. Front. Immunol. 2012, 3, 249. [Google Scholar] [CrossRef] [Green Version]
  92. Huang, Y.; Huang, J.; Chen, Y. Alpha-Helical Cationic Antimicrobial Peptides: Relationships of Structure and Function. Protein Cell 2010, 1, 143–152. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Falanga, A.; Lombardi, L.; Franci, G.; Vitiello, M.; Iovene, M.R.; Morelli, G.; Galdiero, M.; Galdiero, S. Marine Antimicrobial Peptides: Nature Provides Templates for the Design of Novel Compounds against Pathogenic Bacteria. Int. J. Mol. Sci. 2016, 17, 785. [Google Scholar] [CrossRef] [Green Version]
  94. Abrunhosa, F.; Faria, S.; Gomes, P.; Tomaz, I.; Pessoa, J.C.; Andreu, D.; Bastos, M. Interaction and Lipid-Induced Conformation of Two Cecropin−Melittin Hybrid Peptides Depend on Peptide and Membrane Composition. J. Phys. Chem. B 2005, 109, 17311–17319. [Google Scholar] [CrossRef] [Green Version]
  95. Breukink, E.; Wiedemann, I.; van Kraaij, C.; Kuipers, O.P.; Sahl, H.-G.; de Kruijff, B. Use of the Cell Wall Precursor Lipid II by a Pore-Forming Peptide Antibiotic. Science 1999, 286, 2361–2364. [Google Scholar] [CrossRef] [Green Version]
  96. Shai, Y. Mode of Action of Membrane Active Antimicrobial Peptides. Biopolymers 2002, 66, 236–248. [Google Scholar] [CrossRef] [PubMed]
  97. Kumar, P.; Kizhakkedathu, J.N.; Straus, S.K. Antimicrobial Peptides: Diversity, Mechanism of Action and Strategies to Improve the Activity and Biocompatibility In Vivo. Biomolecules 2018, 8, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Rodríguez-Baño, J.; del Toro, M.D.; López-Méndez, J.; Mutters, N.T.; Pascual, A. Minimum Requirements in Infection Control. Clin. Microbiol. Infect. 2015, 21, 1072–1076. [Google Scholar] [CrossRef] [Green Version]
  99. Yan, J.; Wang, K.; Dang, W.; Chen, R.; Xie, J.; Zhang, B.; Song, J.; Wang, R. Two Hits Are Better than One: Membrane-Active and DNA Binding-Related Double-Action Mechanism of NK-18, a Novel Antimicrobial Peptide Derived from Mammalian NK-Lysin. Antimicrob. Agents Chemother. 2013, 57, 220–228. [Google Scholar] [CrossRef] [Green Version]
  100. Park, C.B.; Kim, H.S.; Kim, S.C. Mechanism of Action of the Antimicrobial Peptide Buforin II: Buforin II Kills Microorganisms by Penetrating the Cell Membrane and Inhibiting Cellular Functions. Biochem. Biophys. Res. Commun. 1998, 244, 253–257. [Google Scholar] [CrossRef] [Green Version]
  101. Lee, B.; Hwang, J.S.; Lee, D.G. Antibacterial Action of Lactoferricin B like Peptide against Escherichia Coli: Reactive Oxygen Species-induced Apoptosis-like Death. J. Appl. Microbiol. 2020, 129, 287–295. [Google Scholar] [CrossRef]
  102. Hancock, R.E.W.; Rozek, A. Role of Membranes in the Activities of Antimicrobial Cationic Peptides. FEMS Microbiol. Lett. 2002, 206, 143–149. [Google Scholar] [CrossRef]
  103. Hilchie, A.L.; Wuerth, K.; Hancock, R.E.W. Immune Modulation by Multifaceted Cationic Host Defense (Antimicrobial) Peptides. Nat. Chem. Biol. 2013, 9, 761–768. [Google Scholar] [CrossRef] [PubMed]
  104. Afacan, N.J.; Yeung, A.T.Y.; Pena, O.M.; Hancock, R.E.W. Therapeutic Potential of Host Defense Peptides in Antibiotic-Resistant Infections. Curr. Pharm. Des. 2012, 18, 807–819. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Nijnik, A.; Hancock, R. Host Defence Peptides: Antimicrobial and Immunomodulatory Activity and Potential Applications for Tackling Antibiotic-Resistant Infections. Emerg. Health Threat. J. 2009, 2, e1. [Google Scholar] [CrossRef]
  106. Zhang, L.; Gallo, R.L. Antimicrobial Peptides. Curr. Biol. 2016, 26, R14–R19. [Google Scholar] [CrossRef]
  107. Coorens, M.; Scheenstra, M.R.; Veldhuizen, E.J.A.; Haagsman, H.P. Interspecies Cathelicidin Comparison Reveals Divergence in Antimicrobial Activity, TLR Modulation, Chemokine Induction and Regulation of Phagocytosis. Sci. Rep. 2017, 7, 40874. [Google Scholar] [CrossRef] [Green Version]
  108. Scott, M.G.; Davidson, D.J.; Gold, M.R.; Bowdish, D.; Hancock, R.E.W. The Human Antimicrobial Peptide LL-37 Is a Multifunctional Modulator of Innate Immune Responses. J. Immunol. 2002, 169, 3883–3891. [Google Scholar] [CrossRef] [Green Version]
  109. Niyonsaba, F.; Iwabuchi, K.; Someya, A.; Hirata, M.; Matsuda, H.; Ogawa, H.; Nagaoka, I. A Cathelicidin Family of Human Antibacterial Peptide LL-37 Induces Mast Cell Chemotaxis. Immunology 2002, 106, 20–26. [Google Scholar] [CrossRef] [PubMed]
  110. Yu, Y.; Zhang, Y.; Zhang, Y.; Lai, Y.; Chen, W.; Xiao, Z.; Zhang, W.; Jin, M.; Yu, B. LL-37-Induced Human Mast Cell Activation through G Protein-Coupled Receptor MrgX2. Int. Immunopharmacol. 2017, 49, 6–12. [Google Scholar] [CrossRef]
  111. Le, C.-F.; Fang, C.-M.; Sekaran, S.D. Intracellular Targeting Mechanisms by Antimicrobial Peptides. Antimicrob. Agents Chemother. 2017, 61, e02340-16. [Google Scholar] [CrossRef] [Green Version]
  112. Luo, Y.; Song, Y. Mechanism of Antimicrobial Peptides: Antimicrobial, Anti-Inflammatory and Antibiofilm Activities. Int. J. Mol. Sci. 2021, 22, 11401. [Google Scholar] [CrossRef] [PubMed]
  113. Rahnamaeian, M.; Cytryńska, M.; Zdybicka-Barabas, A.; Vilcinskas, A. The Functional Interaction between Abaecin and Pore-Forming Peptides Indicates a General Mechanism of Antibacterial Potentiation. Peptides 2016, 78, 17–23. [Google Scholar] [CrossRef] [PubMed]
  114. He, K.; Ludtke, S.J.; Heller, W.T.; Huang, H.W. Mechanism of Alamethicin Insertion into Lipid Bilayers. Biophys. J. 1996, 71, 2669–2679. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Li, W.-F.; Ma, G.-X.; Zhou, X.-X. Apidaecin-Type Peptides: Biodiversity, Structure-Function Relationships and Mode of Action. Peptides 2006, 27, 2350–2359. [Google Scholar] [CrossRef]
  116. Boman, H.G.; Agerberth, B.; Boman, A. Mechanisms of Action on Escherichia Coli of Cecropin P1 and PR-39, Two Antibacterial Peptides from Pig Intestine. Infect. Immun. 1993, 61, 2978–2984. [Google Scholar] [CrossRef] [Green Version]
  117. Cudic, M.; Bulet, P.; Hoffmann, R.; Craik, D.J.; Otvos, L. Chemical Synthesis, Antibacterial Activity and Conformation of Diptericin, an 82-Mer Peptide Originally Isolated from Insects. Eur. J. Biochem. 1999, 266, 549–558. [Google Scholar] [CrossRef]
  118. Chileveru, H.R.; Lim, S.A.; Chairatana, P.; Wommack, A.J.; Chiang, I.-L.; Nolan, E.M. Visualizing Attack of Escherichia Coli by the Antimicrobial Peptide Human Defensin 5. Biochemistry 2015, 54, 1767–1777. [Google Scholar] [CrossRef] [Green Version]
  119. Puri, S.; Edgerton, M. How Does It Kill?: Understanding the Candidacidal Mechanism of Salivary Histatin 5. Eukaryot. Cell 2014, 13, 958–964. [Google Scholar] [CrossRef] [Green Version]
  120. Yoneyama, F.; Imura, Y.; Ichimasa, S.; Fujita, K.; Zendo, T.; Nakayama, J.; Matsuzaki, K.; Sonomoto, K. Lacticin Q, a Lactococcal Bacteriocin, Causes High-Level Membrane Permeability in the Absence of Specific Receptors. Appl. Environ. Microbiol. 2009, 75, 538–541. [Google Scholar] [CrossRef] [Green Version]
  121. Cruz-Chamorro, L.; Puertollano, M.A.; Puertollano, E.; de Cienfuegos, G.A.; de Pablo, M.A. In Vitro Biological Activities of Magainin Alone or in Combination with Nisin. Peptides 2006, 27, 1201–1209. [Google Scholar] [CrossRef]
  122. Piers, K.L.; Brown, M.H.; Hancock, R.E. Improvement of Outer Membrane-Permeabilizing and Lipopolysaccharide-Binding Activities of an Antimicrobial Cationic Peptide by C-Terminal Modification. Antimicrob. Agents Chemother. 1994, 38, 2311–2316. [Google Scholar] [CrossRef] [Green Version]
  123. Ruhr, E.; Sahl, H.G. Mode of Action of the Peptide Antibiotic Nisin and Influence on the Membrane Potential of Whole Cells and on Cytoplasmic and Artificial Membrane Vesicles. Antimicrob. Agents Chemother. 1985, 27, 841–845. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Sugiarto, H.; Yu, P.-L. Mechanisms of Action of Ostrich Beta-Defensins against Escherichia Coli. FEMS Microbiol. Lett. 2007, 270, 195–200. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Katsu, T.; Nakao, S.; Iwanaga, S. Mode of Action of an Antimicrobial Peptide, Tachyplesin I, on Biomembranes. Biol. Pharm. Bull. 1993, 16, 178–181. [Google Scholar] [CrossRef] [Green Version]
  126. Dash, R.; Bhattacharjya, S. Thanatin: An Emerging Host Defense Antimicrobial Peptide with Multiple Modes of Action. Int. J. Mol. Sci. 2021, 22, 1522. [Google Scholar] [CrossRef] [PubMed]
  127. Koo, S.P.; Yeaman, M.R.; Nast, C.C.; Bayer, A.S. The Cytoplasmic Membrane Is a Primary Target for the Staphylocidal Action of Thrombin-Induced Platelet Microbicidal Protein. Infect. Immun. 1997, 65, 4795–4800. [Google Scholar] [CrossRef]
  128. Ganz, T. The Role of Antimicrobial Peptides in Innate Immunity. Integr. Comp. Biol. 2003, 43, 300–304. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Harder, J.; Schröder, J.-M.; Gläser, R. The Skin Surface as Antimicrobial Barrier: Present Concepts and Future Outlooks. Exp. Dermatol. 2013, 22, 1–5. [Google Scholar] [CrossRef]
  130. Mahlapuu, M.; Håkansson, J.; Ringstad, L.; Björn, C. Antimicrobial Peptides: An Emerging Category of Therapeutic Agents. Front. Cell. Infect. Microbiol. 2016, 6, 194. [Google Scholar] [CrossRef] [Green Version]
  131. Lu, Y.; Ma, Y.; Wang, X.; Liang, J.; Zhang, C.; Zhang, K.; Lin, G.; Lai, R. The First Antimicrobial Peptide from Sea Amphibian. Mol. Immunol. 2008, 45, 678–681. [Google Scholar] [CrossRef]
  132. Wang, Z.; Wang, X.; Wang, J. Recent Advances in Antibacterial and Antiendotoxic Peptides or Proteins from Marine Resources. Mar. Drugs 2018, 16, 57. [Google Scholar] [CrossRef] [Green Version]
  133. Kościuczuk, E.M.; Lisowski, P.; Jarczak, J.; Strzałkowska, N.; Jóźwik, A.; Horbańczuk, J.; Krzyżewski, J.; Zwierzchowski, L.; Bagnicka, E. Cathelicidins: Family of Antimicrobial Peptides. A Review. Mol. Biol. Rep. 2012, 39, 10957–10970. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Treffers, C.; Chen, L.; Anderson, R.C.; Yu, P.-L. Isolation and Characterisation of Antimicrobial Peptides from Deer Neutrophils. Int. J. Antimicrob. Agents 2005, 26, 165–169. [Google Scholar] [CrossRef]
  135. Brogden, K.A.; Ackermann, M.; McCray, P.B.; Tack, B.F. Antimicrobial Peptides in Animals and Their Role in Host Defences. Int. J. Antimicrob. Agents 2003, 22, 465–478. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Zhang, Y.; Zou, A.; Manchu, R.; Zhou, Y.; Wang, S. Purification and Antimicrobial Activity of Antimicrobial Protein from Brown-Spotted Grouper, Epinephelus Fario. Zool. Res. 2009, 29, 627–632. [Google Scholar] [CrossRef]
  137. Haney, E.F.; Hancock, R.E.W. Peptide Design for Antimicrobial and Immunomodulatory Applications. Biopolymers 2013, 100, 572–583. [Google Scholar] [CrossRef] [Green Version]
  138. Maróti, G.; Kereszt, A.; Kondorosi, E.; Mergaert, P. Natural Roles of Antimicrobial Peptides in Microbes, Plants and Animals. Res. Microbiol. 2011, 162, 363–374. [Google Scholar] [CrossRef]
  139. Oppenheim, J.J.; Biragyn, A.; Kwak, L.W.; Yang, D. Roles of Antimicrobial Peptides Such as Defensins in Innate and Adaptive Immunity. Ann. Rheum. Dis. 2003, 62 (Suppl. S2), ii17–ii21. [Google Scholar] [CrossRef]
  140. Qi, R.-H.; Chen, Y.; Guo, Z.-L.; Zhang, F.; Fang, Z.; Huang, K.; Yu, H.-N.; Wang, Y.-P. Identification and Characterization of Two Novel Cathelicidins from the Frog Odorrana Livida. Zool. Res. 2019, 40, 94–101. [Google Scholar] [CrossRef] [Green Version]
  141. Brown, S.E.; Howard, A.; Kasprzak, A.B.; Gordon, K.H.; East, P.D. A Peptidomics Study Reveals the Impressive Antimicrobial Peptide Arsenal of the Wax Moth Galleria Mellonella. Insect Biochem. Mol. Biol. 2009, 39, 792–800. [Google Scholar] [CrossRef]
  142. Hancock, R.E.W.; Brown, K.L.; Mookherjee, N. Host Defence Peptides from Invertebrates—Emerging Antimicrobial Strategies. Immunobiology 2006, 211, 315–322. [Google Scholar] [CrossRef]
  143. Bulet, P.; Stöcklin, R. Insect Antimicrobial Peptides: Structures, Properties and Gene Regulation. Protein Pept. Lett. 2005, 12, 3–11. [Google Scholar] [CrossRef]
  144. Tam, J.P.; Wang, S.; Wong, K.H.; Tan, W.L. Antimicrobial Peptides from Plants. Pharmaceuticals 2015, 8, 711–757. [Google Scholar] [CrossRef] [Green Version]
  145. Stotz, H.U.; Thomson, J.G.; Wang, Y. Plant Defensins: Defense, Development and Application. Plant Signal Behav. 2009, 4, 1010–1012. [Google Scholar] [CrossRef] [Green Version]
  146. Hammami, R.; Ben Hamida, J.; Vergoten, G.; Fliss, I. PhytAMP: A Database Dedicated to Antimicrobial Plant Peptides. Nucleic Acids Res. 2009, 37, D963–D968. [Google Scholar] [CrossRef] [Green Version]
  147. Nawrot, R.; Barylski, J.; Nowicki, G.; Broniarczyk, J.; Buchwald, W.; Goździcka-Józefiak, A. Plant Antimicrobial Peptides. Folia Microbiol 2014, 59, 181–196. [Google Scholar] [CrossRef] [Green Version]
  148. Mylonakis, E.; Podsiadlowski, L.; Muhammed, M.; Vilcinskas, A. Diversity, Evolution and Medical Applications of Insect Antimicrobial Peptides. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2016, 371, 20150290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Rozgonyi, F.; Szabo, D.; Kocsis, B.; Ostorházi, E.; Abbadessa, G.; Cassone, M.; Wade, J.D.; Otvos, L. The Antibacterial Effect of a Proline-Rich Antibacterial Peptide A3-APO. Curr. Med. Chem. 2009, 16, 3996–4002. [Google Scholar] [CrossRef]
  150. Ciumac, D.; Gong, H.; Hu, X.; Lu, J.R. Membrane targeting cationic antimicrobial peptides. J. Colloid. Interface Sci. 2019, 537, 163–185. [Google Scholar] [CrossRef] [PubMed]
  151. Wu, Q.; Patočka, J.; Kuča, K. Insect Antimicrobial Peptides, a Mini Review. Toxins 2018, 10, 461. [Google Scholar] [CrossRef] [PubMed]
  152. Yang, S.-C.; Lin, C.-H.; Sung, C.T.; Fang, J.-Y. Antibacterial Activities of Bacteriocins: Application in Foods and Pharmaceuticals. Front. Microbiol. 2014, 5, 241. [Google Scholar] [CrossRef] [Green Version]
  153. Mills, S.; Serrano, L.M.; Griffin, C.; O’Connor, P.M.; Schaad, G.; Bruining, C.; Hill, C.; Ross, R.P.; Meijer, W.C. Inhibitory Activity of Lactobacillus Plantarum LMG P-26358 against Listeria Innocua When Used as an Adjunct Starter in the Manufacture of Cheese. Microb. Cell Fact. 2011, 10 (Suppl. S1), S7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Silva, C.C.G.; Silva, S.P.M.; Ribeiro, S.C. Application of Bacteriocins and Protective Cultures in Dairy Food Preservation. Front. Microbiol. 2018, 9, 594. [Google Scholar] [CrossRef] [PubMed]
  155. Galanakis, C. Food Quality and Shelf Life; Elsevier: Amsterdam, The Netherlands, 2019; ISBN 978-0-12-817190-5. [Google Scholar]
  156. Erdem Büyükkiraz, M.; Kesmen, Z. Antimicrobial Peptides (AMPs): A Promising Class of Antimicrobial Compounds. J. Appl. Microbiol. 2022, 132, 1573–1596. [Google Scholar] [CrossRef] [PubMed]
  157. Elsser-Gravesen, D.; Elsser-Gravesen, A. Biopreservatives. Adv. Biochem. Eng. Biotechnol. 2014, 143, 29–49. [Google Scholar] [CrossRef]
  158. Thouzeau, C.; Le Maho, Y.; Froget, G.; Sabatier, L.; Le Bohec, C.; Hoffmann, J.A.; Bulet, P. Spheniscins, Avian Beta-Defensins in Preserved Stomach Contents of the King Penguin, Aptenodytes Patagonicus. J. Biol. Chem. 2003, 278, 51053–51058. [Google Scholar] [CrossRef] [Green Version]
  159. Józefiak, D.; Kierończyk, B.; Juśkiewicz, J.; Zduńczyk, Z.; Rawski, M.; Długosz, J.; Sip, A.; Højberg, O. Dietary Nisin Modulates the Gastrointestinal Microbial Ecology and Enhances Growth Performance of the Broiler Chickens. PLoS ONE 2013, 8, e85347. [Google Scholar] [CrossRef] [Green Version]
  160. Zhang, J.X.; Zhang, S.F.; Wang, T.D.; Guo, X.J.; Hu, R.L. Mammary Gland Expression of Antibacterial Peptide Genes to Inhibit Bacterial Pathogens Causing Mastitis. J. Dairy Sci. 2007, 90, 5218–5225. [Google Scholar] [CrossRef] [Green Version]
  161. Hao, H.; Cheng, G.; Iqbal, Z.; Ai, X.; Hussain, H.I.; Huang, L.; Dai, M.; Wang, Y.; Liu, Z.; Yuan, Z. Benefits and Risks of Antimicrobial Use in Food-Producing Animals. Front. Microbiol. 2014, 5, 288. [Google Scholar] [CrossRef] [Green Version]
  162. Cutler, S.A.; Lonergan, S.M.; Cornick, N.; Johnson, A.K.; Stahl, C.H. Dietary Inclusion of Colicin E1 Is Effective in Preventing Postweaning Diarrhea Caused by F18-Positive Escherichia Coli in Pigs. Antimicrob. Agents Chemother. 2007, 51, 3830–3835. [Google Scholar] [CrossRef] [Green Version]
  163. Tang, Z.; Yin, Y.; Zhang, Y.; Huang, R.; Sun, Z.; Li, T.; Chu, W.; Kong, X.; Li, L.; Geng, M.; et al. Effects of Dietary Supplementation with an Expressed Fusion Peptide Bovine Lactoferricin-Lactoferrampin on Performance, Immune Function and Intestinal Mucosal Morphology in Piglets Weaned at Age 21 d. Br. J. Nutr. 2009, 101, 998–1005. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Wu, S.; Zhang, F.; Huang, Z.; Liu, H.; Xie, C.; Zhang, J.; Thacker, P.A.; Qiao, S. Effects of the Antimicrobial Peptide Cecropin AD on Performance and Intestinal Health in Weaned Piglets Challenged with Escherichia Coli. Peptides 2012, 35, 225–230. [Google Scholar] [CrossRef]
  165. Guo, N.; Zhang, B.; Hu, H.; Ye, S.; Chen, F.; Li, Z.; Chen, P.; Wang, C.; He, Q. Caerin1.1 Suppresses the Growth of Porcine Epidemic Diarrhea Virus In Vitro via Direct Binding to the Virus. Viruses 2018, 10, 507. [Google Scholar] [CrossRef] [Green Version]
  166. Sun, Q.; Wang, K.; She, R.; Ma, W.; Peng, F.; Jin, H. Swine Intestine Antimicrobial Peptides Inhibit Infectious Bronchitis Virus Infectivity in Chick Embryos. Poult. Sci. 2010, 89, 464–469. [Google Scholar] [CrossRef] [PubMed]
  167. Liang, X.; Zhang, X.; Lian, K.; Tian, X.; Zhang, M.; Wang, S.; Chen, C.; Nie, C.; Pan, Y.; Han, F.; et al. Antiviral Effects of Bovine Antimicrobial Peptide against TGEV in Vivo and in Vitro. J. Vet. Sci. 2020, 21, e80. [Google Scholar] [CrossRef]
  168. Hsieh, I.-N.; Hartshorn, K.L. The Role of Antimicrobial Peptides in Influenza Virus Infection and Their Potential as Antiviral and Immunomodulatory Therapy. Pharmaceuticals 2016, 9, 53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Rekha; Naik, S.N.; Prasad, R. Pesticide Residue in Organic and Conventional Food-Risk Analysis. J. Chem. Health Saf. 2006, 13, 12–19. [Google Scholar] [CrossRef]
  170. Kamysz, W.; Krolicka, A.; Bogucka, K.; Ossowski, T.; Lukasiak, J.; Lojkowska, E. Antibacterial Activity of Synthetic Peptides Against Plant Pathogenic Pectobacterium Species. J. Phytopathol. 2005, 153, 313–317. [Google Scholar] [CrossRef]
  171. Gao, A.G.; Hakimi, S.M.; Mittanck, C.A.; Wu, Y.; Woerner, B.M.; Stark, D.M.; Shah, D.M.; Liang, J.; Rommens, C.M. Fungal Pathogen Protection in Potato by Expression of a Plant Defensin Peptide. Nat. Biotechnol. 2000, 18, 1307–1310. [Google Scholar] [CrossRef]
  172. Zakharchenko, N.S.; Rukavtsova, E.B.; Gudkov, A.T.; Bur’ianov, I.I. [Enhanced resistance to phytopathogenic bacteria in transgenic tobacco plants with synthetic gene of antimicrobial peptide cecropin P1]. Genetika 2005, 41, 1445–1452. [Google Scholar] [CrossRef]
  173. Alan, A.R.; Blowers, A.; Earle, E.D. Expression of a Magainin-Type Antimicrobial Peptide Gene (MSI-99) in Tomato Enhances Resistance to Bacterial Speck Disease. Plant Cell Rep. 2004, 22, 388–396. [Google Scholar] [CrossRef]
  174. Vidal, J.R.; Kikkert, J.R.; Malnoy, M.A.; Wallace, P.G.; Barnard, J.; Reisch, B.I. Evaluation of Transgenic “Chardonnay” (Vitis Vinifera) Containing Magainin Genes for Resistance to Crown Gall and Powdery Mildew. Transgenic Res. 2006, 15, 69–82. [Google Scholar] [CrossRef] [PubMed]
  175. Chakrabarti, A.; Ganapathi, T.R.; Mukherjee, P.K.; Bapat, V.A. MSI-99, a Magainin Analogue, Imparts Enhanced Disease Resistance in Transgenic Tobacco and Banana. Planta 2003, 216, 587–596. [Google Scholar] [CrossRef]
  176. Yin, Z.-X.; He, W.; Chen, W.-J.; Yan, J.-H.; Yang, J.-N.; Chan, S.-M.; He, J.-G. Cloning, Expression and Antimicrobial Activity of an Antimicrobial Peptide, Epinecidin-1, from the Orange-Spotted Grouper, Epinephelus Coioides. Aquaculture 2006, 253, 204–211. [Google Scholar] [CrossRef]
  177. León, R.; Ruiz, M.; Valero, Y.; Cárdenas, C.; Guzman, F.; Vila, M.; Cuesta, A. Exploring Small Cationic Peptides of Different Origin as Potential Antimicrobial Agents in Aquaculture. Fish Shellfish Immunol. 2020, 98, 720–727. [Google Scholar] [CrossRef] [PubMed]
  178. Deslouches, B.; Di, Y.P. Antimicrobial Peptides: A Potential Therapeutic Option for Surgical Site Infections. Clin. Surg. 2017, 2, 1740. [Google Scholar]
  179. Loeffler, J.M.; Nelson, D.; Fischetti, V.A. Rapid Killing of Streptococcus Pneumoniae with a Bacteriophage Cell Wall Hydrolase. Science 2001, 294, 2170–2172. [Google Scholar] [CrossRef] [PubMed]
  180. Van ’t Hof, W.; Veerman, E.C.; Helmerhorst, E.J.; Amerongen, A.V. Antimicrobial Peptides: Properties and Applicability. Biol. Chem. 2001, 382, 597–619. [Google Scholar] [CrossRef]
  181. Gee, M.L.; Burton, M.; Grevis-James, A.; Hossain, M.A.; McArthur, S.; Palombo, E.A.; Wade, J.D.; Clayton, A.H.A. Imaging the Action of Antimicrobial Peptides on Living Bacterial Cells. Sci. Rep. 2013, 3, 1557. [Google Scholar] [CrossRef] [Green Version]
  182. Hurdle, J.G.; O’Neill, A.J.; Chopra, I.; Lee, R.E. Targeting Bacterial Membrane Function: An Underexploited Mechanism for Treating Persistent Infections. Nat. Rev. Microbiol. 2011, 9, 62–75. [Google Scholar] [CrossRef] [Green Version]
  183. Reddy, K.V.R.; Yedery, R.D.; Aranha, C. Antimicrobial Peptides: Premises and Promises. Int. J. Antimicrob. Agents 2004, 24, 536–547. [Google Scholar] [CrossRef]
  184. Gennaro, R.; Skerlavaj, B.; Romeo, D. Purification, Composition, and Activity of Two Bactenecins, Antibacterial Peptides of Bovine Neutrophils. Infect. Immun. 1989, 57, 3142–3146. [Google Scholar] [CrossRef] [Green Version]
  185. Björstad, A.; Askarieh, G.; Brown, K.L.; Christenson, K.; Forsman, H.; Onnheim, K.; Li, H.-N.; Teneberg, S.; Maier, O.; Hoekstra, D.; et al. The Host Defense Peptide LL-37 Selectively Permeabilizes Apoptotic Leukocytes. Antimicrob. Agents Chemother. 2009, 53, 1027–1038. [Google Scholar] [CrossRef] [Green Version]
  186. Dürr, U.H.N.; Sudheendra, U.S.; Ramamoorthy, A. LL-37, the Only Human Member of the Cathelicidin Family of Antimicrobial Peptides. Biochim. Biophys. Acta 2006, 1758, 1408–1425. [Google Scholar] [CrossRef] [Green Version]
  187. Xhindoli, D.; Pacor, S.; Benincasa, M.; Scocchi, M.; Gennaro, R.; Tossi, A. The Human Cathelicidin LL-37—A Pore-Forming Antibacterial Peptide and Host-Cell Modulator. Biochim. Biophys. Acta 2016, 1858, 546–566. [Google Scholar] [CrossRef]
  188. De Breij, A.; Riool, M.; Cordfunke, R.A.; Malanovic, N.; de Boer, L.; Koning, R.I.; Ravensbergen, E.; Franken, M.; van der Heijde, T.; Boekema, B.K.; et al. The Antimicrobial Peptide SAAP-148 Combats Drug-Resistant Bacteria and Biofilms. Sci. Transl. Med. 2018, 10, eaan4044. [Google Scholar] [CrossRef] [Green Version]
  189. Dijkmans, A.C.; Wilms, E.B.; Kamerling, I.M.C.; Birkhoff, W.; Ortiz-Zacarías, N.V.; van Nieuwkoop, C.; Verbrugh, H.A.; Touw, D.J. Colistin: Revival of an Old Polymyxin Antibiotic. Ther. Drug Monit. 2015, 37, 419–427. [Google Scholar] [CrossRef]
  190. Stevenson, C.L. Advances in Peptide Pharmaceuticals. Curr. Pharm. Biotechnol. 2009, 10, 122–137. [Google Scholar] [CrossRef]
  191. Sierra, J.M.; Fusté, E.; Rabanal, F.; Vinuesa, T.; Viñas, M. An Overview of Antimicrobial Peptides and the Latest Advances in Their Development. Expert Opin. Biol. Ther. 2017, 17, 663–676. [Google Scholar] [CrossRef]
  192. Lipsky, B.A.; Holroyd, K.J.; Zasloff, M. Topical versus Systemic Antimicrobial Therapy for Treating Mildly Infected Diabetic Foot Ulcers: A Randomized, Controlled, Double-Blinded, Multicenter Trial of Pexiganan Cream. Clin. Infect. Dis. 2008, 47, 1537–1545. [Google Scholar] [CrossRef] [Green Version]
  193. Rijsbergen, M.; Rijneveld, R.; Todd, M.; Feiss, G.L.; Kouwenhoven, S.T.P.; Quint, K.D.; van Alewijk, D.C.J.G.; de Koning, M.N.C.; Klaassen, E.S.; Burggraaf, J.; et al. Results of Phase 2 Trials Exploring the Safety and Efficacy of Omiganan in Patients with Human Papillomavirus-Induced Genital Lesions. Br. J. Clin. Pharmacol. 2020, 86, 2133–2143. [Google Scholar] [CrossRef]
  194. Wang, H.; Ai, L.; Zhang, Y.; Cheng, J.; Yu, H.; Li, C.; Zhang, D.; Pan, Y.; Lin, L. The Effects of Antimicrobial Peptide Nal-P-113 on Inhibiting Periodontal Pathogens and Improving Periodontal Status. Biomed. Res. Int. 2018, 2018, 1805793. [Google Scholar] [CrossRef] [Green Version]
  195. Mangoni, M.L.; McDermott, A.M.; Zasloff, M. Antimicrobial Peptides and Wound Healing: Biological and Therapeutic Considerations. Exp. Dermatol. 2016, 25, 167–173. [Google Scholar] [CrossRef] [Green Version]
  196. Spohn, R.; Daruka, L.; Lázár, V.; Martins, A.; Vidovics, F.; Grézal, G.; Méhi, O.; Kintses, B.; Számel, M.; Jangir, P.K.; et al. Integrated Evolutionary Analysis Reveals Antimicrobial Peptides with Limited Resistance. Nat. Commun. 2019, 10, 4538. [Google Scholar] [CrossRef] [Green Version]
  197. Volejníková, A.; Melicherčík, P.; Nešuta, O.; Vaňková, E.; Bednárová, L.; Rybáček, J.; Čeřovský, V. Antimicrobial Peptides Prevent Bacterial Biofilm Formation on the Surface of Polymethylmethacrylate Bone Cement. J. Med. Microbiol. 2019, 68, 961–972. [Google Scholar] [CrossRef]
  198. Zhang, Q.-Y.; Yan, Z.-B.; Meng, Y.-M.; Hong, X.-Y.; Shao, G.; Ma, J.-J.; Cheng, X.-R.; Liu, J.; Kang, J.; Fu, C.-Y. Antimicrobial Peptides: Mechanism of Action, Activity and Clinical Potential. Mil. Med. Res. 2021, 8, 48. [Google Scholar] [CrossRef]
  199. Hancock, R.E.W.; Scott, M.G. The Role of Antimicrobial Peptides in Animal Defenses. Proc. Natl. Acad. Sci. USA 2000, 97, 8856–8861. [Google Scholar] [CrossRef]
  200. Brandenburg, K.; Heinbockel, L.; Correa, W.; Lohner, K. Peptides with Dual Mode of Action: Killing Bacteria and Preventing Endotoxin-Induced Sepsis. Biochim. Biophys. Acta 2016, 1858, 971–979. [Google Scholar] [CrossRef]
  201. Brandenburg, K. Lack of New Antiinfective Agents: Passing into the Pre-Antibiotic Age? WJBC 2015, 6, 71. [Google Scholar] [CrossRef]
  202. Brandenburg, K.; Andrä, J.; Garidel, P.; Gutsmann, T. Peptide-Based Treatment of Sepsis. Appl. Microbiol. Biotechnol. 2011, 90, 799–808. [Google Scholar] [CrossRef]
  203. Dutta, P.; Das, S. Mammalian Antimicrobial Peptides: Promising Therapeutic Targets Against Infection and Chronic Inflammation. CTMC 2015, 16, 99–129. [Google Scholar] [CrossRef] [PubMed]
  204. Wang, C.; Hong, T.; Cui, P.; Wang, J.; Xia, J. Antimicrobial peptides towards clinical application: Delivery and formulation. Adv. Drug Deliv. Rev. 2021, 175, 113818. [Google Scholar] [CrossRef] [PubMed]
  205. Atefyekta, S.; Blomstrand, E.; Rajasekharan, A.K.; Svensson, S.; Trobos, M.; Hong, J.; Webster, T.J.; Thomsen, P.; Andersson, M. Antimicrobial Peptide-Functionalized Mesoporous Hydrogels. ACS Biomater. Sci. Eng. 2021, 7, 1693–1702. [Google Scholar] [CrossRef]
  206. Boto, A.; Pérez de la Lastra, J.; González, C. The Road from Host-Defense Peptides to a New Generation of Antimicrobial Drugs. Molecules 2018, 23, 311. [Google Scholar] [CrossRef] [Green Version]
  207. Li, J.; Koh, J.-J.; Liu, S.; Lakshminarayanan, R.; Verma, C.S.; Beuerman, R.W. Membrane Active Antimicrobial Peptides: Translating Mechanistic Insights to Design. Front. Neurosci. 2017, 11, 73. [Google Scholar] [CrossRef] [Green Version]
  208. Greco, I.; Molchanova, N.; Holmedal, E.; Jenssen, H.; Hummel, B.D.; Watts, J.L.; Håkansson, J.; Hansen, P.R.; Svenson, J. Correlation between Hemolytic Activity, Cytotoxicity and Systemic in Vivo Toxicity of Synthetic Antimicrobial Peptides. Sci. Rep. 2020, 10, 13206. [Google Scholar] [CrossRef] [PubMed]
  209. Svendsen, J.S.M.; Grant, T.M.; Rennison, D.; Brimble, M.A.; Svenson, J. Very Short and Stable Lactoferricin-Derived Antimicrobial Peptides: Design Principles and Potential Uses. Acc. Chem. Res. 2019, 52, 749–759. [Google Scholar] [CrossRef]
  210. Yurkova, M.S.; Zenin, V.A.; Sadykhov, E.G.; Fedorov, A.N. Dimerization of Antimicrobial Peptide Polyphemusin I into One Polypeptide Chain: Theoretical and Practical Consequences. Appl. Biochem. Microbiol. 2020, 56, 893–897. [Google Scholar] [CrossRef]
  211. Irazazabal, L.N.; Porto, W.F.; Ribeiro, S.M.; Casale, S.; Humblot, V.; Ladram, A.; Franco, O.L. Selective Amino Acid Substitution Reduces Cytotoxicity of the Antimicrobial Peptide Mastoparan. Biochim. Biophys. Acta 2016, 1858, 2699–2708. [Google Scholar] [CrossRef]
  212. Zhao, Y.; Zhang, M.; Qiu, S.; Wang, J.; Peng, J.; Zhao, P.; Zhu, R.; Wang, H.; Li, Y.; Wang, K.; et al. Antimicrobial Activity and Stability of the D-Amino Acid Substituted Derivatives of Antimicrobial Peptide Polybia-MPI. AMB Express 2016, 6, 122. [Google Scholar] [CrossRef] [Green Version]
  213. Kindrachuk, J.; Scruten, E.; Attah-Poku, S.; Bell, K.; Potter, A.; Babiuk, L.A.; Griebel, P.J.; Napper, S. Stability, Toxicity, and Biological Activity of Host Defense Peptide BMAP28 and Its Inversed and Retro-Inversed Isomers. Biopolymers 2011, 96, 14–24. [Google Scholar] [CrossRef]
  214. Carratalá, J.V.; Serna, N.; Villaverde, A.; Vázquez, E.; Ferrer-Miralles, N. Nanostructured Antimicrobial Peptides: The Last Push towards Clinics. Biotechnol. Adv. 2020, 44, 107603. [Google Scholar] [CrossRef] [PubMed]
  215. Umerska, A.; Cassisa, V.; Matougui, N.; Joly-Guillou, M.-L.; Eveillard, M.; Saulnier, P. Antibacterial Action of Lipid Nanocapsules Containing Fatty Acids or Monoglycerides as Co-Surfactants. Eur. J. Pharm. Biopharm. 2016, 108, 100–110. [Google Scholar] [CrossRef]
  216. Radaic, A.; de Jesus, M.B.; Kapila, Y.L. Bacterial Anti-Microbial Peptides and Nano-Sized Drug Delivery Systems: The State of the Art toward Improved Bacteriocins. J. Control. Release 2020, 321, 100–118. [Google Scholar] [CrossRef] [PubMed]
  217. Tang, Z.; Ma, Q.; Chen, X.; Chen, T.; Ying, Y.; Xi, X.; Wang, L.; Ma, C.; Shaw, C.; Zhou, M. Recent Advances and Challenges in Nanodelivery Systems for Antimicrobial Peptides (AMPs). Antibiotics 2021, 10, 990. [Google Scholar] [CrossRef] [PubMed]
  218. Rajasekaran, G.; Kim, E.Y.; Shin, S.Y. LL-37-Derived Membrane-Active FK-13 Analogs Possessing Cell Selectivity, Anti-Biofilm Activity and Synergy with Chloramphenicol and Anti-Inflammatory Activity. Biochim. Biophys. Acta 2017, 1859, 722–733. [Google Scholar] [CrossRef]
  219. Yu, G.; Baeder, D.Y.; Regoes, R.R.; Rolff, J. Combination Effects of Antimicrobial Peptides. Antimicrob. Agents Chemother. 2016, 60, 1717–1724. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  220. Sheard, D.E.; O’Brien-Simpson, N.M.; Wade, J.D.; Separovic, F. Combating Bacterial Resistance by Combination of Antibiotics with Antimicrobial Peptides. Pure Appl. Chem. 2019, 91, 199–209. [Google Scholar] [CrossRef]
  221. Xu, X.; Xu, L.; Yuan, G.; Wang, Y.; Qu, Y.; Zhou, M. Synergistic Combination of Two Antimicrobial Agents Closing Each Other’s Mutant Selection Windows to Prevent Antimicrobial Resistance. Sci. Rep. 2018, 8, 7237. [Google Scholar] [CrossRef]
  222. Jangra, M.; Raka, V.; Nandanwar, H. In Vitro Evaluation of Antimicrobial Peptide Tridecaptin M in Combination with Other Antibiotics against Multidrug Resistant Acinetobacter Baumannii. Molecules 2020, 25, 3255. [Google Scholar] [CrossRef]
  223. Oo, T.Z.; Cole, N.; Garthwaite, L.; Willcox, M.D.P.; Zhu, H. Evaluation of Synergistic Activity of Bovine Lactoferricin with Antibiotics in Corneal Infection. J. Antimicrob. Chemother. 2010, 65, 1243–1251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Nuding, S.; Frasch, T.; Schaller, M.; Stange, E.F.; Zabel, L.T. Synergistic Effects of Antimicrobial Peptides and Antibiotics against Clostridium Difficile. Antimicrob. Agents Chemother. 2014, 58, 5719–5725. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Jahangiri, A.; Neshani, A.; Mirhosseini, S.A.; Ghazvini, K.; Zare, H.; Sedighian, H. Synergistic Effect of Two Antimicrobial Peptides, Nisin and P10 with Conventional Antibiotics against Extensively Drug-Resistant Acinetobacter Baumannii and Colistin-Resistant Pseudomonas Aeruginosa Isolates. Microb. Pathog. 2021, 150, 104700. [Google Scholar] [CrossRef] [PubMed]
  226. Portelinha, J.; Angeles-Boza, A.M. The Antimicrobial Peptide Gad-1 Clears Pseudomonas Aeruginosa Biofilms under Cystic Fibrosis Conditions. ChemBioChem 2021, 22, 1646–1655. [Google Scholar] [CrossRef]
  227. Dosler, S.; Mataraci, E. In Vitro Pharmacokinetics of Antimicrobial Cationic Peptides Alone and in Combination with Antibiotics against Methicillin Resistant Staphylococcus Aureus Biofilms. Peptides 2013, 49, 53–58. [Google Scholar] [CrossRef]
  228. Mataraci, E.; Dosler, S. In Vitro Activities of Antibiotics and Antimicrobial Cationic Peptides Alone and in Combination against Methicillin-Resistant Staphylococcus Aureus Biofilms. Antimicrob. Agents Chemother. 2012, 56, 6366–6371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Chen, Y.; Wu, J.; Cheng, H.; Dai, Y.; Wang, Y.; Yang, H.; Xiong, F.; Xu, W.; Wei, L. Anti-Infective Effects of a Fish-Derived Antimicrobial Peptide Against Drug-Resistant Bacteria and Its Synergistic Effects with Antibiotic. Front. Microbiol. 2020, 11, 602412. [Google Scholar] [CrossRef]
  230. Song, M.; Liu, Y.; Huang, X.; Ding, S.; Wang, Y.; Shen, J.; Zhu, K. A Broad-Spectrum Antibiotic Adjuvant Reverses Multidrug-Resistant Gram-Negative Pathogens. Nat. Microbiol. 2020, 5, 1040–1050. [Google Scholar] [CrossRef]
  231. Zarkotou, O.; Pournaras, S.; Tselioti, P.; Dragoumanos, V.; Pitiriga, V.; Ranellou, K.; Prekates, A.; Themeli-Digalaki, K.; Tsakris, A. Predictors of Mortality in Patients with Bloodstream Infections Caused by KPC-Producing Klebsiella Pneumoniae and Impact of Appropriate Antimicrobial Treatment. Clin. Microbiol. Infect. 2011, 17, 1798–1803. [Google Scholar] [CrossRef] [Green Version]
  232. Bassetti, M.; Righi, E. New Antibiotics and Antimicrobial Combination Therapy for the Treatment of Gram-Negative Bacterial Infections. Curr. Opin. Crit. Care 2015, 21, 402–411. [Google Scholar] [CrossRef]
  233. Antonoplis, A.; Zang, X.; Huttner, M.A.; Chong, K.K.L.; Lee, Y.B.; Co, J.Y.; Amieva, M.R.; Kline, K.A.; Wender, P.A.; Cegelski, L. A Dual-Function Antibiotic-Transporter Conjugate Exhibits Superior Activity in Sterilizing MRSA Biofilms and Killing Persister Cells. J. Am. Chem. Soc. 2018, 140, 16140–16151. [Google Scholar] [CrossRef]
  234. Reinhardt, A.; Neundorf, I. Design and Application of Antimicrobial Peptide Conjugates. IJMS 2016, 17, 701. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Liu, J.; Chen, F.; Wang, X.; Peng, H.; Zhang, H.; Wang, K.-J. The Synergistic Effect of Mud Crab Antimicrobial Peptides Sphistin and Sph12−38 with Antibiotics Azithromycin and Rifampicin Enhances Bactericidal Activity Against Pseudomonas Aeruginosa. Front. Cell. Infect. Microbiol. 2020, 10, 572849. [Google Scholar] [CrossRef] [PubMed]
  236. Wu, X.; Li, Z.; Li, X.; Tian, Y.; Fan, Y.; Yu, C.; Zhou, B.; Liu, Y.; Xiang, R.; Yang, L. Synergistic Effects of Antimicrobial Peptide DP7 Combined with Antibiotics against Multidrug-Resistant Bacteria. DDDT 2017, 11, 939–946. [Google Scholar] [CrossRef] [Green Version]
  237. Akbari, R.; Hakemi-Vala, M.; Pashaie, F.; Bevalian, P.; Hashemi, A.; Pooshang Bagheri, K. Highly Synergistic Effects of Melittin with Conventional Antibiotics Against Multidrug-Resistant Isolates of Acinetobacter Baumannii and Pseudomonas Aeruginosa. Microb. Drug Resist. 2019, 25, 193–202. [Google Scholar] [CrossRef] [PubMed]
  238. Naghmouchi, K.; Le Lay, C.; Baah, J.; Drider, D. Antibiotic and Antimicrobial Peptide Combinations: Synergistic Inhibition of Pseudomonas Fluorescens and Antibiotic-Resistant Variants. Res. Microbiol. 2012, 163, 101–108. [Google Scholar] [CrossRef]
  239. Kampshoff, F.; Willcox, M.D.P.; Dutta, D. A Pilot Study of the Synergy between Two Antimicrobial Peptides and Two Common Antibiotics. Antibiotics 2019, 8, 60. [Google Scholar] [CrossRef] [Green Version]
  240. Ruden, S.; Rieder, A.; Chis Ster, I.; Schwartz, T.; Mikut, R.; Hilpert, K. Synergy Pattern of Short Cationic Antimicrobial Peptides Against Multidrug-Resistant Pseudomonas Aeruginosa. Front. Microbiol. 2019, 10, 2740. [Google Scholar] [CrossRef] [Green Version]
  241. Choi, H.; Lee, D.G. Synergistic Effect of Antimicrobial Peptide Arenicin-1 in Combination with Antibiotics against Pathogenic Bacteria. Res. Microbiol. 2012, 163, 479–486. [Google Scholar] [CrossRef]
  242. Lee, H.; Lee, D.G. Novel Approaches for Efficient Antifungal Drug Action. J. Microbiol. Biotechnol. 2018, 28, 1771–1781. [Google Scholar] [CrossRef]
  243. Hanson, M.A.; Dostálová, A.; Ceroni, C.; Poidevin, M.; Kondo, S.; Lemaitre, B. Synergy and Remarkable Specificity of Antimicrobial Peptides in Vivo Using a Systematic Knockout Approach. eLife 2019, 8, e44341. [Google Scholar] [CrossRef] [PubMed]
  244. Dobson, A.J.; Purves, J.; Kamysz, W.; Rolff, J. Comparing Selection on S. Aureus between Antimicrobial Peptides and Common Antibiotics. PLoS ONE 2013, 8, e76521. [Google Scholar] [CrossRef] [Green Version]
  245. Rahnamaeian, M.; Cytryńska, M.; Zdybicka-Barabas, A.; Dobslaff, K.; Wiesner, J.; Twyman, R.M.; Zuchner, T.; Sadd, B.M.; Regoes, R.R.; Schmid-Hempel, P.; et al. Insect Antimicrobial Peptides Show Potentiating Functional Interactions against Gram-Negative Bacteria. Proc. R. Soc. B. 2015, 282, 20150293. [Google Scholar] [CrossRef] [PubMed]
  246. Casciaro, B.; Loffredo, M.R.; Luca, V.; Verrusio, W.; Cacciafesta, M.; Mangoni, M.L. Esculentin-1a Derived Antipseudomonal Peptides: Limited Induction of Resistance and Synergy with Aztreonam. PPL 2019, 25, 1155–1162. [Google Scholar] [CrossRef]
  247. Magana, M.; Pushpanathan, M.; Santos, A.L.; Leanse, L.; Fernandez, M.; Ioannidis, A.; Giulianotti, M.A.; Apidianakis, Y.; Bradfute, S.; Ferguson, A.L.; et al. The Value of Antimicrobial Peptides in the Age of Resistance. Lancet Infect. Dis. 2020, 20, e216–e230. [Google Scholar] [CrossRef]
  248. Glattard, E.; Salnikov, E.S.; Aisenbrey, C.; Bechinger, B. Investigations of the Synergistic Enhancement of Antimicrobial Activity in Mixtures of Magainin 2 and PGLa. Biophys. Chem. 2016, 210, 35–44. [Google Scholar] [CrossRef] [PubMed]
  249. Zerweck, J.; Strandberg, E.; Kukharenko, O.; Reichert, J.; Bürck, J.; Wadhwani, P.; Ulrich, A.S. Molecular Mechanism of Synergy between the Antimicrobial Peptides PGLa and Magainin 2. Sci. Rep. 2017, 7, 13153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  250. Ilyas, H.; Kim, J.; Lee, D.; Malmsten, M.; Bhunia, A. Structural Insights into the Combinatorial Effects of Antimicrobial Peptides Reveal a Role of Aromatic–Aromatic Interactions in Antibacterial Synergism. J. Biol. Chem. 2019, 294, 14615–14633. [Google Scholar] [CrossRef]
  251. Najjar, M.B.; Kashtanov, D.; Chikindas, M.L. Epsilon-Poly-l-Lysine and Nisin A Act Synergistically against Gram-Positive Food-Borne Pathogens Bacillus Cereus and Listeria Monocytogenes. Lett. Appl. Microbiol. 2007, 45, 13–18. [Google Scholar] [CrossRef]
  252. Field, D.; Seisling, N.; Cotter, P.D.; Ross, R.P.; Hill, C. Synergistic Nisin-Polymyxin Combinations for the Control of Pseudomonas Biofilm Formation. Front. Microbiol. 2016, 7, 1713. [Google Scholar] [CrossRef] [Green Version]
  253. Westerhoff, H.V.; Zasloff, M.; Rosner, J.L.; Hendler, R.W.; Waal, A.; Gomes, A.V.; Jongsma, A.P.M.; Riethorst, A.; Juretic, D. Functional Synergism of the Magainins PGLa and Magainin-2 in Escherichia Coli, Tumor Cells and Liposomes. Eur. J. Biochem. 1995, 228, 257–264. [Google Scholar] [CrossRef] [PubMed]
  254. Yan, H.; Hancock, R.E.W. Synergistic Interactions between Mammalian Antimicrobial Defense Peptides. Antimicrob. Agents Chemother. 2001, 45, 1558–1560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Zdybicka-Barabas, A.; Mak, P.; Klys, A.; Skrzypiec, K.; Mendyk, E.; Fiołka, M.J.; Cytryńska, M. Synergistic Action of Galleria Mellonella Anionic Peptide 2 and Lysozyme against Gram-Negative Bacteria. Biochim. Biophys. Acta 2012, 1818, 2623–2635. [Google Scholar] [CrossRef] [Green Version]
  256. Chen, X.; Niyonsaba, F.; Ushio, H.; Okuda, D.; Nagaoka, I.; Ikeda, S.; Okumura, K.; Ogawa, H. Synergistic Effect of Antibacterial Agents Human Beta-Defensins, Cathelicidin LL-37 and Lysozyme against Staphylococcus Aureus and Escherichia Coli. J. Dermatol. Sci. 2005, 40, 123–132. [Google Scholar] [CrossRef] [PubMed]
  257. Zhao, X.; Zhen, Z.; Wang, X.; Guo, N. Synergy of a Combination of Nisin and Citric Acid against Staphylococcus Aureus and Listeria Monocytogenes. Food Addit. Contam. Part. A 2017, 34, 2058–2068. [Google Scholar] [CrossRef]
  258. Ruden, S.; Hilpert, K.; Berditsch, M.; Wadhwani, P.; Ulrich, A.S. Synergistic Interaction between Silver Nanoparticles and Membrane-Permeabilizing Antimicrobial Peptides. Antimicrob. Agents Chemother. 2009, 53, 3538–3540. [Google Scholar] [CrossRef] [Green Version]
  259. Doolin, T.; Amir, H.M.; Duong, L.; Rosenzweig, R.; Urban, L.A.; Bosch, M.; Pol, A.; Gross, S.P.; Siryaporn, A. Mammalian Histones Facilitate Antimicrobial Synergy by Disrupting the Bacterial Proton Gradient and Chromosome Organization. Nat. Commun. 2020, 11, 3888. [Google Scholar] [CrossRef]
  260. Bennett, J.E.; Dolin, E.; Blaser, M.J. Mandell, Douglas, and Bennett’s Principles and Practice of Infectious Diseases, 9th ed.; Elsevier: Philadelphia, PA, USA, 2019. [Google Scholar]
  261. Hobbs, J.K.; Miller, K.; O’Neill, A.J.; Chopra, I. Consequences of Daptomycin-Mediated Membrane Damage in Staphylococcus Aureus. J. Antimicrob. Chemother. 2008, 62, 1003–1008. [Google Scholar] [CrossRef] [Green Version]
  262. Müller, A.; Wenzel, M.; Strahl, H.; Grein, F.; Saaki, T.N.V.; Kohl, B.; Siersma, T.; Bandow, J.E.; Sahl, H.-G.; Schneider, T.; et al. Daptomycin Inhibits Cell Envelope Synthesis by Interfering with Fluid Membrane Microdomains. Proc. Natl. Acad. Sci. USA 2016, 113, E7077–E7086. [Google Scholar] [CrossRef]
  263. Pogliano, J.; Pogliano, N.; Silverman, J.A. Daptomycin-Mediated Reorganization of Membrane Architecture Causes Mislocalization of Essential Cell Division Proteins. J. Bacteriol. 2012, 194, 4494–4504. [Google Scholar] [CrossRef]
  264. Konychev, A.; Heep, M.; Moritz, R.K.C.; Kreuter, A.; Shulutko, A.; Fierlbeck, G.; Bouylout, K.; Pathan, R.; Trostmann, U.; Chaves, R.L. Safety and Efficacy of Daptomycin as First-Line Treatment for Complicated Skin and Soft Tissue Infections in Elderly Patients: An Open-Label, Multicentre, Randomized Phase IIIb Trial. Drugs Aging 2013, 30, 829–836. [Google Scholar] [CrossRef] [PubMed]
  265. Aikawa, N.; Kusachi, S.; Mikamo, H.; Takesue, Y.; Watanabe, S.; Tanaka, Y.; Morita, A.; Tsumori, K.; Kato, Y.; Yoshinari, T. Efficacy and Safety of Intravenous Daptomycin in Japanese Patients with Skin and Soft Tissue Infections. J. Infect. Chemother. 2013, 19, 447–455. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  266. Telles, J.P.; Cieslinski, J.; Tuon, F.F. Daptomycin to Bone and Joint Infections and Prosthesis Joint Infections: A Systematic Review. Braz. J. Infect. Dis. 2019, 23, 191–196. [Google Scholar] [CrossRef]
  267. Jaber, R.H.; Beahm, N.P. Daptomycin for the Treatment of Acute Bacterial Meningitis: A Narrative Review. Int. J. Antimicrob. Agents 2023, 61, 106770. [Google Scholar] [CrossRef]
  268. Rosanova, M.T.; Bes, D.; Serrano-Aguilar, P.; Sberna, N.; Herrera-Ramos, E.; Lede, R.L. Efficacy and Safety of Daptomycin: Systematic Review and Meta-Analysis. Ther. Adv. Infect. Dis. 2019, 6, 2049936119886465. [Google Scholar] [CrossRef] [Green Version]
  269. Li, J.; Nation, R.L.; Turnidge, J.D.; Milne, R.W.; Coulthard, K.; Rayner, C.R.; Paterson, D.L. Colistin: The Re-Emerging Antibiotic for Multidrug-Resistant Gram-Negative Bacterial Infections. Lancet Infect. Dis. 2006, 6, 589–601. [Google Scholar] [CrossRef]
  270. From, A.H.; Fong, J.S.; Good, R.A. Polymyxin B Sulfate Modification of Bacterial Endotoxin: Effects on the Development of Endotoxin Shock in Dogs. Infect. Immun. 1979, 23, 660–664. [Google Scholar] [CrossRef]
  271. Kofteridis, D.P.; Andrianaki, A.M.; Maraki, S.; Mathioudaki, A.; Plataki, M.; Alexopoulou, C.; Ioannou, P.; Samonis, G.; Valachis, A. Treatment Pattern, Prognostic Factors, and Outcome in Patients with Infection Due to Pan-Drug-Resistant Gram-Negative Bacteria. Eur. J. Clin. Microbiol. Infect. Dis. 2020, 39, 965–970. [Google Scholar] [CrossRef]
  272. Jung, S.Y.; Lee, S.H.; Lee, S.Y.; Yang, S.; Noh, H.; Chung, E.K.; Lee, J.I. Antimicrobials for the Treatment of Drug-Resistant Acinetobacter Baumannii Pneumonia in Critically Ill Patients: A Systemic Review and Bayesian Network Meta-Analysis. Crit. Care 2017, 21, 319. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  273. Liu, J.; Shu, Y.; Zhu, F.; Feng, B.; Zhang, Z.; Liu, L.; Wang, G. Comparative Efficacy and Safety of Combination Therapy with High-Dose Sulbactam or Colistin with Additional Antibacterial Agents for Multiple Drug-Resistant and Extensively Drug-Resistant Acinetobacter Baumannii Infections: A Systematic Review and Network Meta-Analysis. J. Glob. Antimicrob. Resist. 2021, 24, 136–147. [Google Scholar] [CrossRef] [PubMed]
  274. Betrosian, A.P.; Frantzeskaki, F.; Xanthaki, A.; Douzinas, E.E. Efficacy and Safety of High-Dose Ampicillin/Sulbactam vs. Colistin as Monotherapy for the Treatment of Multidrug Resistant Acinetobacter Baumannii Ventilator-Associated Pneumonia. J. Infect. 2008, 56, 432–436. [Google Scholar] [CrossRef] [PubMed]
  275. Fowler, V.G.; Boucher, H.W.; Corey, G.R.; Abrutyn, E.; Karchmer, A.W.; Rupp, M.E.; Levine, D.P.; Chambers, H.F.; Tally, F.P.; Vigliani, G.A.; et al. Daptomycin versus Standard Therapy for Bacteremia and Endocarditis Caused by Staphylococcus Aureus. N. Engl. J. Med. 2006, 355, 653–665. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  276. Miller, J.L.; Slatkin, M.H.; Johnson, B.A. Local Use of Bacitracin. J. Investig. Dermatol. 1948, 10, 179–188. [Google Scholar] [CrossRef] [PubMed]
  277. Grönberg, A.; Mahlapuu, M.; Ståhle, M.; Whately-Smith, C.; Rollman, O. Treatment with LL-37 Is Safe and Effective in Enhancing Healing of Hard-to-Heal Venous Leg Ulcers: A Randomized, Placebo-Controlled Clinical Trial. Wound Repair. Regen. 2014, 22, 613–621. [Google Scholar] [CrossRef]
  278. Daley, P.; Louie, T.; Lutz, J.E.; Khanna, S.; Stoutenburgh, U.; Jin, M.; Adedoyin, A.; Chesnel, L.; Guris, D.; Larson, K.B.; et al. Surotomycin versus Vancomycin in Adults with Clostridium Difficile Infection: Primary Clinical Outcomes from the Second Pivotal, Randomized, Double-Blind, Phase 3 Trial. J. Antimicrob. Chemother. 2017, 72, 3462–3470. [Google Scholar] [CrossRef]
  279. Niemeyer-van der Kolk, T.; van der Wall, H.; Hogendoorn, G.K.; Rijneveld, R.; Luijten, S.; van Alewijk, D.C.J.G.; van den Munckhof, E.H.A.; de Kam, M.L.; Feiss, G.L.; Prens, E.P.; et al. Pharmacodynamic Effects of Topical Omiganan in Patients with Mild to Moderate Atopic Dermatitis in a Randomized, Placebo-Controlled, Phase II Trial. Clin. Transl. Sci. 2020, 13, 994–1003. [Google Scholar] [CrossRef]
  280. Peek, N.F.A.W.; Nell, M.J.; Brand, R.; Jansen-Werkhoven, T.; van Hoogdalem, E.J.; Verrijk, R.; Vonk, M.J.; Wafelman, A.R.; Valentijn, A.R.P.M.; Frijns, J.H.M.; et al. Ototopical Drops Containing a Novel Antibacterial Synthetic Peptide: Safety and Efficacy in Adults with Chronic Suppurative Otitis Media. PLoS ONE 2020, 15, e0231573. [Google Scholar] [CrossRef] [Green Version]
  281. Mercer, D.K.; Robertson, J.C.; Miller, L.; Stewart, C.S.; O’Neil, D.A. NP213 (Novexatin®): A Unique Therapy Candidate for Onychomycosis with a Differentiated Safety and Efficacy Profile. Med. Mycol. 2020, 58, 1064–1072. [Google Scholar] [CrossRef] [Green Version]
  282. Mullane, K.; Lee, C.; Bressler, A.; Buitrago, M.; Weiss, K.; Dabovic, K.; Praestgaard, J.; Leeds, J.A.; Blais, J.; Pertel, P. Multicenter, Randomized Clinical Trial to Compare the Safety and Efficacy of LFF571 and Vancomycin for Clostridium Difficile Infections. Antimicrob. Agents Chemother. 2015, 59, 1435–1440. [Google Scholar] [CrossRef] [Green Version]
  283. Corey, R.; Naderer, O.J.; O’Riordan, W.D.; Dumont, E.; Jones, L.S.; Kurtinecz, M.; Zhu, J.Z. Safety, Tolerability, and Efficacy of GSK1322322 in the Treatment of Acute Bacterial Skin and Skin Structure Infections. Antimicrob. Agents Chemother. 2014, 58, 6518–6527. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The mechanisms of action of AMPs. AMPs may act through the permeabilization of bacterial cytoplasmic membranes in three different ways. In the carpet model, most AMPs are located on the membrane, causing the perturbation of its integrity and the formation of micelles. In the barrel-stave model, the AMPs are inserted into the membrane, causing the formation of a pore with the hydrophilic side facing the interior of the pore and the hydrophobic side towards the lipid core of the membrane. A transmembrane pore model involves the formation of a pore, as in the toroidal-pore model, but the pore is also lined with the hydrophilic side of the AMPs and the head parts of the membrane phospholipids. AMP: antimicrobial peptide.
Figure 1. The mechanisms of action of AMPs. AMPs may act through the permeabilization of bacterial cytoplasmic membranes in three different ways. In the carpet model, most AMPs are located on the membrane, causing the perturbation of its integrity and the formation of micelles. In the barrel-stave model, the AMPs are inserted into the membrane, causing the formation of a pore with the hydrophilic side facing the interior of the pore and the hydrophobic side towards the lipid core of the membrane. A transmembrane pore model involves the formation of a pore, as in the toroidal-pore model, but the pore is also lined with the hydrophilic side of the AMPs and the head parts of the membrane phospholipids. AMP: antimicrobial peptide.
Life 13 01651 g001
Figure 2. Graphical representation showing the immunomodulatory function of AMPs. Apart from the direct killing mechanisms of AMPs, they can kill pathogens through the induction of innate and adaptive immunity. AMPs can induce cells involved in innate immunity, such as macrophages (Mφs), polymorphonuclear cells (PMNs), and mast cells. AMP: antimicrobial peptide; DCs: dendritic cells; Mφs: macrophages; PRR: pathogen recognition receptor; PMN: polymorphonuclear cell; ROS: reactive oxygen species; NETs: neutrophil extracellular traps.
Figure 2. Graphical representation showing the immunomodulatory function of AMPs. Apart from the direct killing mechanisms of AMPs, they can kill pathogens through the induction of innate and adaptive immunity. AMPs can induce cells involved in innate immunity, such as macrophages (Mφs), polymorphonuclear cells (PMNs), and mast cells. AMP: antimicrobial peptide; DCs: dendritic cells; Mφs: macrophages; PRR: pathogen recognition receptor; PMN: polymorphonuclear cell; ROS: reactive oxygen species; NETs: neutrophil extracellular traps.
Life 13 01651 g002
Table 1. Classification of antimicrobial peptides based on the mechanism of action.
Table 1. Classification of antimicrobial peptides based on the mechanism of action.
NameSourceMechanism of ActionRef
AbaecinBumblebeeInhibits DnaK[113]
AlamethicinFungus Trichoderma virideBarrel-stave transmembrane pore model[114]
Apidaecin Hb1aHoneybeeInhibits protein biosynthesis by targeting ribosomes, inhibits DnaK and GroEL, inhibits ABC transport system, binds LPS[115]
Buforin IIEnzymatic cleavage of buforin I (from Asian toad Bufo bufo gargarizans)Inhibits DNA, inhibits RNA[100]
Cecropin P1Pig intestineCarpet model/Detergent-like mode non-membrane pore model[116]
DiptericinHemolymph of injured Sarcophaga peregrine larvaInhibits septation[117]
HD5 (oxidized)Human small intestineInhibits cell division[118]
Histatin-5Human salivaInhibits proteases, inhibits MMP-2 and MMP-9, inhibits generation of ROS[119]
Lacticin QLactococcus lactis QU 5Toroidal transmembrane pore model[120]
Magainin 1African clawed frog Xenopus laevisInhibits energy metabolism proteins, inhibits amino-acid metabolism[121]
MBI-27Derived from part of silk moth cecropin and bee melittin peptidesInhibits LPS[122]
NisinLactococcus lactisInhibits lipid II in peptidoglycan biosynthesis[123]
Ostricacin-1Ostrich defensinInhibits DNA[124]
TachyplesinHorseshoe crab hemocytesBinds DNA minor groove[125]
ThanatinPodisus maculiventrisAgglutination non-membrane pore model[126]
tPMP-1Platelet granulesActivation of autolytic enzyme[127]
AMPs: antimicrobial peptides; LPS: lipopolysaccharide; ROS: reactive oxygen species. This table is not exhaustive of all AMPs or all mechanisms of action.
Table 2. Examples of AMP combinations with antibiotics that exert positive synergistic effects.
Table 2. Examples of AMP combinations with antibiotics that exert positive synergistic effects.
AMPAntibioticTargetRef
Tridecaptin MRifampicin, vancomycin, ceftazidimeA. baumannii[222]
Lactoferricin Ciprofloxacin, ceftazidimeP. aeruginosa[223]
LL-37, HBD3Tigecycline, moxifloxacin, piperacillin/tazobactam, meropenemC. difficile[224]
P10Ceftazidime, doripenemA. baumannii and P. aeruginosa[225]
Gad-1Kanamycin, ciprofloxacinP. aeruginosa[226]
NisinPenicillin, chloramphenicol, ciprofloxacin, indolicidin, or azithromycinS. aureus[227,228]
SAAP-148Demeclocycline hydrochloride (DMCT)P. aeruginosa[229]
(SLAP)-S25Cefepime, colistin, ofloxacin, rifampicin, tetracycline, and vancomycinmultidrug-resistant Gram-negative pathogens[230]
Colistintigecycline, carbapenem, gentamicinKlebsiella KPC[231,232]
OctaarginineVancomycinbiofilms of S. aureus[233,234]
Sphistin, Sph12−38Rifampicin, azithromycinP. aeruginosa[235]
DP7Azithromycin, vancomycinS. aureus, P. aeruginosa, A. baumannii, E. coli[236]
P10Ceftazidime, doripenemA. baumannii, colistin-resistant P. aeruginosa[225]
MelittinDoripenem and ceftazidime A. baumannii and P. aeruginosa[237]
LL 17–29ChloramphenicolS. aureus, P. aeruginosa[218]
Nisin Z, pediocin, or colistinPenicillin, ampicillin, or rifampicinP. fluorescens[238]
MelamineCiprofloxacin, fluoroquinoloneP. aeruginosa[239]
Indolicidin, polymyxin BTobramycin, gentamycin, and amikacinP. aeruginosa[240]
Arenicin-1Ampicillin, erythromycin, and chloramphenicolS. aureus, S. epidermis, P. aeruginosa, and E. coli[241]
AMP: antimicrobial peptide.
Table 3. Examples of AMP combinations that exert positive synergistic effects.
Table 3. Examples of AMP combinations that exert positive synergistic effects.
AMPSynergistic AMP MoleculeTargetRef
HsAFP1RsAFP2 and RsAFP1Biofilm cells[242]
DiptericinsAttacinsP. burhodogranariea[243]
PexigananMelittinS. aureus[244]
AbaecinHymenoptaecinGram-negative bacteria[245]
Esculentin-1aAztreonamP. aeruginosa[246]
PGLaMagainin 2E. coli[247,248,249]
VG16KRKPKYE28plant pathogens[250]
Nisin AEpsilon-poly-L-lysineBacillus cereus and L. monocytogenes[251]
NisinColistin Biofilms of P. aeruginosa[225,252]
Nisin ColistinA. baumannii, colistin-resistant P. aeruginosa[225]
Magainin-2Peptidyl-glycylleucine-carboxyamideE. coli[253]
Indolicidin, LL-37BactenecinP. aeruginosa and E. coli[254]
IndolicidinBactenecinP. aeruginosa and E. coli[254]
Protegrin 1LL-37E. faecalis[254]
BactenecinLL-37E. faecalis[254]
Protegrin1BactenecinE. faecalis[254]
Apidaecin, pexigananLL 19–27E. coli[219]
Galleria mellonella anionic peptide 2LysozymeGram-negative bacteria[255]
B-defensin, LL-37LysozymeS. aureus, E. coli[256]
AMP: antimicrobial peptide.
Table 4. Examples of AMP combinations with other compounds that exert positive synergistic effects.
Table 4. Examples of AMP combinations with other compounds that exert positive synergistic effects.
AMPSynergistic CompoundTargetRef
NisinCitric acidS. aureus and L. monocytogenes[257]
Nisin AEpsilon-poly-L-lysineGram-positive food-borne pathogens Bacillus cereus and L. monocytogenes[251]
Polymyxin B, Gramicidin SSilver nitrate, silver nanoparticlesGram-negative bacteria[258]
AMP: antimicrobial peptide.
Table 5. Examples of AMP combinations with histones that exert positive synergistic effects.
Table 5. Examples of AMP combinations with histones that exert positive synergistic effects.
AMPHistoneTargetRef
LL-37 and magainin-2H2A and H3Gram-positive and Gram-negative bacteria[259]
Polymyxin BH2AE. coli[259]
AMP: antimicrobial peptide.
Table 6. Examples of antimicrobial peptides in human clinical trials.
Table 6. Examples of antimicrobial peptides in human clinical trials.
StudyPopulationInterventionComparatorOutcome
Fowler et al., 2006 [275]—Phase III RCTPatients with S. aureus bacteremia with or without endocarditisIV treatment with the AMP daptomycinIV treatment with low-dose gentamicin plus either an antistaphylococcal penicillin or vancomycinDaptomycin was non-inferior to standard-of-care. Treatment success rates were similar in subgroups of patients with complicated bacteremia, right-sided endocarditis, and methicillin-resistant S. aureus
Miller et al., 1948 [276]—non-randomized study130 patients with superficial infections of the skin and 35 patients with secondary skin infectionsBacitracin applied locallyNoneCure rate higher than 50% in superficial skin infections and 100% in secondary skin infections
NCT05340790—Phase I RCTHealthy female volunteersDose 1 to 5 of AMP PL-18 vaginal suppositoriesPlacebo doses 1 to 5 of vaginal suppositoriesSafety assessment (recruiting)
Gronberg et al., 2014 [277]—RCTAdult patients with hard-to-heal venous leg ulcersRepeated doses of LL-37 applied locallyRepeated doses of placebo applied locallySafe and well tolerated. Significant early healing of ulcers
Daley et al., 2017 [278]—Phase III RCTAdult patients with Clostridioides difficile infectionAMP surotomycin orallyVancomycin orallyNon-inferior but non-superior to vancomycin for clinical response
Lipsky et al., 2008 [192]—Phase III RCTAdult diabetic patients with infected wounds at the lower extremitiesAMP pexiganan locally and oral placeboLocal placebo and oral ofloxacinPexiganan is comparable to oral ofloxacin for mildly infected diabetic ulcers
Niemeyer-van der Kolk et al., 2020 [279]—Phase II RCTPatients 18–65 years old with mild-to-moderate atopic dermatitisAMP Omiganan locallyVehicle gel locally (placebo)Treatment with the AMP improved dysbiosis in mild to moderate atopic dermatitis patients
Small improvements in clinical scores were detected
Peek et al., 2020 [280]—Phase II RCTAdults with chronic suppurative otitis media resistant to antibiotic therapyAMP P60.4Ac locally with ear dropsVehicle locally with ear drops (placebo)Safe and well-tolerated treatment. Significantly higher treatment success than placebo
NCT00231153, Phase III RCTPatients with central venous cathetersAMP Omiganan 1% gel local application at the catheter insertion sitePovidone-Iodine 10% local application at the catheter insertion siteFailed to show adequate efficacy in catheter-associated infections
NCT04767321, Phase I/II RCTAdults 18–65 years old with persistent carriage of S. aureusNasal application of the AMP LTX-109Nasal application of placeboSafety, tolerability, and microbial eradication—recruitment completed
Mercer et al., 2020 [281]—Phase I and Phase II RCTs12, 48, and 47 patients with onychomycosis of the toenailLocal application of NP213Local application of placeboNP213 clinical safety profile. Positive patient-reported outcomes
Mullane et al., 2015 [282]—Phase II RCT72 patients with Clostridioides difficile infectionLFF571 orallyVancomycin orallyThe rate of clinical cure was non-inferior to that of vancomycin. Similar 30-day sustained cure rates.
More adverse events
for LFF571.
Corey et al., 2014 [283]—Phase IIa RCT84 adult patients with ABSSSIGSK1322322 orallyLinezolid orallyClinical success in the ITT population and the per-protocol population were 67 and 91% in the GSK1322322-treated group and 89 and 100% in the linezolid-treated group
ABSSSI: acute bacterial skin and skin structure infection; AMP: antimicrobial peptide; ITT: intention-to-treat; IV: intravenous; RCT: randomized controlled trial; This table may not be exhaustive of all antimicrobial peptides that are currently approved or in clinical trials.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ioannou, P.; Baliou, S.; Kofteridis, D.P. Antimicrobial Peptides in Infectious Diseases and Beyond—A Narrative Review. Life 2023, 13, 1651. https://doi.org/10.3390/life13081651

AMA Style

Ioannou P, Baliou S, Kofteridis DP. Antimicrobial Peptides in Infectious Diseases and Beyond—A Narrative Review. Life. 2023; 13(8):1651. https://doi.org/10.3390/life13081651

Chicago/Turabian Style

Ioannou, Petros, Stella Baliou, and Diamantis P. Kofteridis. 2023. "Antimicrobial Peptides in Infectious Diseases and Beyond—A Narrative Review" Life 13, no. 8: 1651. https://doi.org/10.3390/life13081651

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

Article Metrics

Back to TopTop