Next Article in Journal
Dendritic Glycerol-Cholesterol Amphiphiles as Drug Delivery Systems: A Comparison between Monomeric and Polymeric Structures
Previous Article in Journal
IL-17 Inhibition: A Valid Therapeutic Strategy in the Management of Hidradenitis Suppurativa
Previous Article in Special Issue
Mechanistic Insight into the Early Stages of Toroidal Pore Formation by the Antimicrobial Peptide Smp24
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Topoisomeric Membrane-Active Peptides: A Review of the Last Two Decades

1
Department of Medicine and Life Sciences, Universitat Pompeu Fabra, 08003 Barcelona, Spain
2
Department of Biotechnology, University of Rijeka, 51000 Rijeka, Croatia
*
Author to whom correspondence should be addressed.
Pharmaceutics 2023, 15(10), 2451; https://doi.org/10.3390/pharmaceutics15102451
Submission received: 30 August 2023 / Revised: 6 October 2023 / Accepted: 10 October 2023 / Published: 12 October 2023
(This article belongs to the Special Issue State of the Art of Membrane Active Peptides)

Abstract

:
In recent decades, bioactive peptides have been gaining recognition in various biomedical areas, such as intracellular drug delivery (cell-penetrating peptides, CPPs) or anti-infective action (antimicrobial peptides, AMPs), closely associated to their distinct mode of interaction with biological membranes. Exploiting the interaction of membrane-active peptides with diverse targets (healthy, tumoral, bacterial or parasitic cell membranes) is opening encouraging prospects for peptides in therapeutics. However, ordinary peptides formed by L-amino acids are easily decomposed by proteases in biological fluids. One way to sidestep this limitation is to use topoisomers, namely versions of the peptide made up of D-amino acids in either canonic (enantio) or inverted (retroenantio) sequence. Rearranging peptide sequences in this fashion provides a certain degree of native structure mimicry that, in appropriate contexts, may deliver desirable biological activity while avoiding protease degradation. In this review, we will focus on recent accounts of membrane-active topoisomeric peptides with therapeutic applications as CPP drug delivery vectors, or as antimicrobial and anticancer candidates. We will also discuss the most common modes of interaction of these peptides with their membrane targets.

Graphical Abstract

1. Introduction

Membranes are biological structures that keep the internal contents (cytoplasm, nucleus, etc.) of a cell separate from the external medium, thus ensuring cell integrity. The distinctive constitutive feature of biological membranes is a phospholipid bilayer matrix that proteins, drugs, or any other molecule directed toward an intracellular destination must cross to reach their target [1,2,3]. Membrane-active agents perturb the basic features (e.g., permeability, fluidity, etc.) of the membrane and have a significant impact on cell function (e.g., transport, signaling, and/or integrity). A growing body of recent research has focused on various types of agents capable of modulating membrane physical properties and performance [4]. Among these, peptides stand out as particularly promising on account of their structural versatility that often translates into favorable physicochemical and biochemical properties. High target specificity, low toxicity, and (most often) low immunogenicity underpin the prospects of peptides in therapeutic areas [5,6] such as anti-infectives [7,8], cancer [9,10,11], cardiovascular [12,13], and Alzheimer’s disease [14,15], among others.
Chirality is a key structural feature governing the action and specificity of bioactive peptides. In peptides, chirality stems from the presence of a stereogenic center at the α-carbon of every amino acid monomer (except Gly) of any peptide sequence (plus additional stereocenters at the β-carbons of Ile and Thr). While the vast majority of natural peptides are in an L-configuration at their α-carbon stereocenters, exceptions consisting of D-amino-acid-containing peptides in various families (arthropods, mollusks, amphibians, etc.) are known [16,17,18,19,20,21,22,23,24]. Peptides partially or totally made up of D-amino acids are predictably more stable in biological fluids than their all-L counterparts due to the resistance of peptide bonds with an adjoining α-carbon of a non-L configuration to protease hydrolysis [25,26,27]. This feature has been creatively exploited in recent times to boost the in vivo stability/performance of quite a few therapeutic peptides [28,29]. However, it is by no means a trivial tactic, as the switch from the L- to D-configuration at one or more stereocenters has an inevitable impact on the 3D structure of a peptide, which more often than not translates into partial or total loss of bioactivity. Tweaking peptide stereochemistry, therefore, inescapably requires significant effort in terms of, first, educated prediction and, second, the subsequent experimental evaluation of the biological outcomes of such structural manipulations [30].
Among the various approaches adopted by peptide medicinal chemists to overcome the in vivo vulnerability of bioactive peptides, this review focuses on tactics involving the conversion of bioactive peptide structures into their topoisomer versions. A generic definition of topoisomerism would correspond to composition-identical (hence isomeric) yet three-dimensionally distinct (hence topo-) variants of a molecule, with structural changes translating into different biological effects. In the peptide/protein field, this term has been preferentially applied to multiple disulfide peptides adopting different types of spatial folding as a result of alternative Cys pairings. Less frequently but also suitably, here and elsewhere [31], the term is used to describe all-D-amino acid counterparts of a peptide with either a conserved (the enantio version, abbreviated as e hereafter) or fully reversed (the retroenantio version, abbreviated as re) sequence relative to the canonical version. While all-D versions with, e.g., pairwise residue switches (or other maneuvers also preserving global composition) could arguably be included within the topoisomer category, it seems more fitting to restrict the term to full sequence inversion. Of particular interest for this work is the re modification (also named retro-inverso or retro-all-D by other authors, though herein we prefer re). As an example, Figure 1 shows the four possible arrangements a representative heptapeptide may adopt as far as sequence and chirality are concerned (the all-L but fully sequence-inverted retro version (denoted as r hereafter) are included for consistency). When the re analogue in a standard view (N-to-C-terminal from left to right) is flipped horizontally, i.e., a 180° rotation on an in-plane axis that produces a C-to-N (from left to right) arrangement, and then compared with the original peptide (all-L, N-to-C from left to right), the side chains in both structures adopt coincident orientations (Figure 1B), while the amide bond directions are reversed (NH-CO in the re vs. standard CO-NH in the parent) [25]. As peptide bioactivity largely involves side chain contacts, the side-chain-superimposable re version is a topological mimic of the parent peptide, with predictably better stability in biological media [32].

2. Origins of Topoisomer Peptides

Among the early attempts to explore the relationship between reversed chirality and biological activity, the work of Stewart and Wooley regarding all-D bradykinin [35,36] and of Vogler et al. regarding all-D-Val5-angiotensin II-Asp1-β-amide [35,36] is worth mentioning. In both instances, the e versions turned out to be completely inactive, and the outcome was rightfully interpreted in terms of a strict chiral requirement for productive receptor interaction.
The first successful report of a biologically active enantiomer was made in the seminal 1967 paper by Shemyakin et al. on enniatin B [37] (Figure 2), a cyclic depsipeptide (i.e., with both peptide and ester bonds) antibiotic produced by Fusarium fungi. The fact that synthetic enantio enniatin B was equipotent with the natural compound was rationalized as follows:
“Indeed, if one turns the formulas […] 60° in the plane of the figure, all the like asymmetric centers coincide, while each ester group will take the place of the N-methyl amide group and vice versa. […] There should therefore be a close matching of both these topochemically similar antipodes to the same stereoselective receptor…”.
By noting that a 60° rotation placed the side chains of both enantiomers pointing in the same direction (albeit with different –ester or amide– connecting units in each case), the authors were paving the way to the re concept of peptides topologically superimposable with the parent isomer by simple maneuvers (e.g., 60° rotation in enniatin; or 180° on an in-the-plane axis, switching N- and C-termini), and likely to be biologically active as recognizable by a chiral receptor yet plausibly longer-lived in biological media due to their D-amino acid residue contents [38].
In another significant paper published in 1969 [39], the same authors expanded their proposal to other peptide structures. Thus, [Gly5,10] gramicidin S, an antimicrobial peptide (AMP) active against Gram-positive bacteria, was compared with its re isomer and found to have comparable activity. In that same paper, Shemyakin et al. [39] recommended N- and C-terminal blocking (acetyl and carboxamide, respectively) for a linear re peptide to retain the activity of its parent, by attenuating electrostaticc effects of the end groups and equalizing hydrophobicity.
In later years, similar principles of e and re switching were generalized for AMPs by the Merrifield group [38,40,41,42,43] and others [44,45], highlighting the importance of factors such as overall positive charge, amphipathicity, and serum stability for effective antimicrobial effect [46,47,48,49,50,51,52]. Again focusing on peptide–membrane interactions, in 1994, Prochiantz et al. were able to demonstrate that cell-penetrating peptides (CPPs) can enter cells independently of their chiral configuration [53]. The internalization of canonic and e versions of 16-residue penetratin was studied and found to be comparable, with both peptides appearing to cross the membrane through energy-independent mechanisms [53,54]. This finding has been corroborated for other CPP sequences in recent years [55,56,57,58], including CPP shuttles crossing crucial therapeutic boundaries such as the blood–brain barrier [59].
Finally, another area where topoisomeric approaches have been explored is that of peptide-based vaccines, where van Regenmortel et al. [60] demonstrated the antigenic mimicry between native peptide antigens and their re counterparts. In contrast to the substantial amount of work in the AMP and CPP fields, the literature does not show that these pioneering efforts have inspired a significant follow-up.
Generalizing topoisomer strategies like those successfully described above for AMPs and CPPs to other peptide therapeutic areas such as peptide hormones has not been straightforward, underscoring the fact that, in many other cases, stereochemically stringent interactions are at play, which can be contrasted with the fairly relaxed chiral rules of engagement for membrane-active peptides.

3. Membrane-Active Peptides and Their Mechanisms of Interaction

Peptide–membrane interactions arise from distinctive sequence motifs within peptides that correspond to specific physicochemical properties. Membrane-active peptides typically possess structural features (electrical charge, amphipathicity, etc.) that foster—or prevent—interaction with membrane bilayers of diverse lipid compositions. These membrane-active peptides have been categorized into two main groups based on their mode of interaction: AMPs and anticancer peptides (ACP), both functioning as membrane disruptors, and CPPs, which translocate across the membrane to enter the cell [61]. Interestingly, these groups of peptides often exhibit similar properties, with instances where AMP/ACPs overlap with CPPs, and vice versa [62].
As mentioned, AMPs and ACPs exert their main biological effects by disrupting or modulating the integrity and function of cell membranes [63,64]. Despite the 20,000-plus hitherto catalogued AMP/ACP sequences (from natural sources and/or the result of man-made design and production), some aspects of their mechanisms of action still stimulate research and controversy [65,66]. There is a consensus that the ability to disrupt the membrane is not only dependent on the type of AMP/ACP sequence and its conformation but also on the membrane composition of the target cells. In most typical scenarios, cationic AMP/ACPs interact with anionic phospholipid head groups on the outer leaflet of bacterial, fungal, or protozoan membranes through electrostatic and hydrophobic interactions [67]. Upon reaching a concentration threshold, the peptides fuse into the membrane and damage its structure [47,48] via processes such as those depicted in Figure 3.
Unlike AMPs and ACPs, membrane-active CPPs enter cells via innocuously translocating membranes, without ensuing cell death [69]. This makes CPPs valuable shuttles for crossing cell membranes or other physiological barriers and intracellularly delivering diverse payloads [70,71]. The precise mechanisms of cell entry, again subject to some controversy, have nonetheless been recognized to be influenced by factors such as physico-chemical properties [72], cargo [73], concentration [74], cell type [75], temperature [76], and the environmental status of the cell [74]. There is also agreement that CPP–membrane interaction occurs primarily via either passive diffusion or endocytic or non-endocytic pathways (Figure 4).
Passive transport is an energy-independent process that relies on the concentration gradient of a substance. It is commonly observed for ions and small molecules, including some peptides, and it is driven by the natural tendency of particles to move down concentration gradients until equilibrium is reached. This movement occurs through mechanisms such as facilitated diffusion [77,78,79,80] and osmosis [77,81,82,83].
Endocytosis, an energy-dependent process, involves the engulfment of molecules on a membrane to gain cell entry [84]. CPPs employ endocytosis as a frequent mechanism of internalization [84]. Endocytosis encompasses three types of processes: phagocytosis, pinocytosis, and receptor-mediated endocytosis. Pinocytosis occurs in various cell types and involves the uptake of extracellular fluid, while phagocytosis is specific to macrophages and leukocytes and entails the internalization of diverse substances. Receptor-mediated endocytosis occurs when molecular shuttles bind to specific receptors, such as in the LDL-mediated transport of cholesterol into cells. A graphical representation of these processes is shown in Figure 4 [84,85,86].
Figure 4. Two distinct pathways for the internalization of cell-penetrating peptides (CPPs), adapted from [87]. Endocytosis (left) involves the internalization of molecules through a process where the cell membrane engulfs them. It encompasses different subtypes, such as phagocytosis and pinocytosis, all involving expenditure of energy by the cell to internalize payloads. Non-endocytic routes (right) rely on specific interactions between the molecules involved and the cell membrane or other cellular components to facilitate uptake.
Figure 4. Two distinct pathways for the internalization of cell-penetrating peptides (CPPs), adapted from [87]. Endocytosis (left) involves the internalization of molecules through a process where the cell membrane engulfs them. It encompasses different subtypes, such as phagocytosis and pinocytosis, all involving expenditure of energy by the cell to internalize payloads. Non-endocytic routes (right) rely on specific interactions between the molecules involved and the cell membrane or other cellular components to facilitate uptake.
Pharmaceutics 15 02451 g004
Uptake to the cytosol via energy-independent, non-endocytic pathways can be broadly categorized into two groups depending on whether or not pore formation occurs [69,74,88]. In the former case, CPPs disrupt the lipid bilayer membrane after a critical concentration is reached [89,90,91], much like the mechanism used by AMPs to disrupt bacterial membranes (see Figure 3) [48,92,93,94,95]. Additionally, certain circumstances allow CPPs and pathogenic amyloid peptides to use this mechanism for cell entry and activity (see Figure 4) [90]. Transmembrane pore mechanisms include the toroidal pore [92,96] and the barrel-stave model [48,92,96,97]. In contrast, non-pore mechanisms involve the carpet [92,96] and membrane-thinning models, among others [92,98].

4. Membrane-Active Topoisomeric Peptides

In designing effective membrane-active peptides, not only membrane activity but also stability in biological fluids are crucial considerations. Many promising membrane-active peptide candidates have seen their progress toward adoption on the market hampered by excessively high susceptibility to protease degradation [99,100]. As mentioned in the introduction, D-amino-acid-containing analogs of bioactive peptides, in particular those typifying e and re topoisomeric versions, may constitute viable strategies for curtailing peptide clearance via body proteases and ensuring longer systemic survival. In the following sections, we discuss in more detail examples of topoisomers within the AMP, CPP, and ACP families developed to this end.

4.1. AMP Topoisomers for Facing the AMR Challenge

The persistent overuse of antibiotics in both preclinical and clinical settings has contributed to an alarming rise in antimicrobial resistance (AMR). The cursory prescription of broad-spectrum antibiotics to patients with suspected infections resulted in the unwarranted use of, e.g., over 30% of prescribed antibiotics in the USA alone in 2014, with adverse effects in up to 20% of patients [101,102]. The World Health Organization (WHO) warned that, without intervention, the global consequences of AMR may be devastating, with a projected annual death toll of up to 10 million people by 2050 [103]. This prediction surpasses the combined deaths caused by cancer (8.2 million) and diabetes (1.5 million) [104].
The COVID-19 pandemic has exacerbated this issue [105,106,107,108]. According to a USA antimicrobial resistance report, 80% of patients hospitalized with COVID-19 received antibiotics [109] even though only a minority (17.6%) had a confirmed bacterial infection [110]. The financial burden on governments in relation to eventually tackling the AMRs ensuing from these practices will likely be huge. For instance, 6 out of the 18 most-worrying AMR threats have fetched up a staggering USD 4.6 bn annual bill in the USA alone [109]. Not surprisingly, deaths attributed to AMR are projected to escalate even more rapidly due to the aforementioned factors [111].
A major challenge in the AMR struggle revolves around the six ESKAPE bacteria (Enterococcus faecium, Staphylococcus aureus, Klebsiella pneumoniae, Acinetobacter baumannii, Pseudomonas aeruginosa, and Enterobacter spp.) [112,113,114], a group of highly virulent pathogens with an extraordinary ability to defeat antibiotic activity. In this context, AMPs are promising contenders [115,116,117] due to their rapid action at low micromolar concentrations; their broad spectrum, encompassing both Gram-positive and -negative bacteria, fungi, and viruses [92,97]; and their mechanisms of action, with a much lower tendency to develop resistance compared to conventional antibiotics [118,119].
AMPs were originally isolated from natural sources. Gramicidin S [120], extracted from the soil bacterium Bacillus brevis [121,122], was first reported in 1939. The finding of AMPs in prokaryotes raised the question of whether eukaryotes also produced AMPs against infections, particularly plants or insects lacking an immune system.
A substance in wheat flour found to be lethal to bread yeast was first described in 1896 [123], but it took some 80 years until it was isolated in a pure form (purothionin) from wheat endosperm in 1972 and shown to inhibit bacterial growth [124]. In 1962, a paper described antibacterial activity in the skin secretion of the Bombina variegata frog [125]; this activity was later identified as corresponding to the AMP bombinin [126]. Subsequent reports of eukaryotic AMPs included the cecropins (1981) of the Hyalophora cecropia moth related by Boman et al. [127,128] and, also in the 1980s, the α-defensins of rabbits [129,130,131] and humans [132] reported by Lehrer et al. and the magainins [133] from the Xenopus laevis frog reported by Zasloff et al. in 1987; this was followed by an ever-growing stream of AMPs from diverse sources such as β-defensins and θ-defensins from immune cells [134,135] or the first anionic (Asp rich) AMP reported in the mid-1990s [136]. At present, AMPs have been found in all types of organisms, including plants [93,137], animals [138,139,140,141,142], and bacteria [143], and they have been have widely acknowledged as ideal candidates for tackling AMR [144,145,146,147,148].
Utilizing topoisomers as natural mimics of AMPs to develop improved candidates is a sensible strategy with which to combat AMR. Among AMP topoisomers, re versions have a high degree of structural resemblance to natural AMPs and are therefore promising candidates. Table 1 lists the contributions to the AMP topoisomer repertoire over the last decades.
In the 1990s, the Merrifield laboratory [42] described the e version of the natural AMPs cecropin A, melittin, and magainin 2 amide, as well as cecropin–melittin hybrids, and showed that the antimicrobial activity of the D-enantiomeric versions was equivalent to that of L-parental peptides. They also found that the mechanisms of action did not require a specific chiral receptor [42]. More recently, Kumar et al. studied the e and re forms of peptide 73, a derivative of aurein2.2 (GLFDIVKKVVGAL) [149,150]. Both e73 and re73 versions exhibited activity similar to that of peptide 73, including efficacy against S. aureus in a cutaneous infection model [150].
In a study by Lynn et al., the e and re forms of BMAP-28, a bovine cathelicidin AMP, were tested for their activity against Leishmania parasites. Both topoisomers effectively reduced both promastigote and amastigote forms of L. major [151,152]. In contrast, canonic BMAP-28 was ineffective due to degradation by the parasite metalloproteinase GP63.
Another bovine AMP, the 13-amino-acid indolicidin, isolated from neutrophil granules [153], was also studied in its r, e, and re versions [44]. While all the peptides exhibited antimicrobial activity comparable to that of natural indolicidin, those incorporating D-amino acids were advantageous due to their lower hemolytic activity.
Crotalicidin (Ctn), a cathelicidin AMP derived from South American pit vipers, exhibits both antibacterial and anticancer properties. Falcão et al. dissected Ctn and showed that a Ctn [15-34] fragment had similar activity but much better serum stability when compared to the parent peptide [154,155,156]. More recently [31], we investigated the e, r, and re topoisomers of Ctn and Ctn [15-34] and showed that while Ctn topoisomers underwent a 50% reduction in antimicrobial activity compared to the L-form, activity and improved serum stability were maintained for Ctn [15-34] topoisomers.
Neubauer et al. investigated the activity of the AMPs aurein 1.2, CAMEL, citropin 1.1, omiganan, pexiganan, and temporin A along with their r analogues. With the exception of r-omiganan, the retro analogues exhibited reduced activity compared to their native counterparts. The authors attributed the lower antimicrobial efficacy observed to a relatively higher hydrophilicity in comparison to the natural peptides. This study was useful in charting the limitations of retro analogues in antimicrobial applications, with hydrophobicity and hemolytic activity being presented as particularly relevant issues [157].
The last two examples serve to emphasize the need for caution when proposing and/or implementing topoisomeric approaches in AMPs (or, for that matter, in bioactive peptides of any type), as the biological outcomes of sequence inversion (r versions), D-amino acid replacement (e versions), or the combination of both (re versions) are arguably non-innocuous given their impact on D structure and ultimately activity [31,99].
Table 1. Examples of recent AMP topoisomers.
Table 1. Examples of recent AMP topoisomers.
Name aSequence bTopoisomer ClassActive AgainstObservations cReferences
Retro-indolicinRRWPWWPWKWPLIrG+ (S. aureus), G− (E. coli)Same MIC values.[44]
Inverso-indolicinilpwkwpwwpwreG+ (S. aureus), G− (E. coli)Same MIC values and increased stability.[44]
Retroinverso-indolicinrrwpwwpwkwplireG+ (S. aureus), G− (E. coli)Same MIC values and increased stability.[44]
Retro-[Trp4,6,8,9,11Phe]-indolicidinRRFPFFPFKFPLIrG+ (S. aureus), G− (E. coli)Same MIC values.[44]
Inverso-[Trp4,6,8,9,11Phe]-indolicidinilpfkfpffpfrreG+ (S. aureus), G− (E. coli)Same MIC values and increased stability.[44]
Retroinverso-[Trp4,6,8,9,11Phe]-indolicidinrrfpffpfkfplireG+ (S. aureus), G− (E. coli)Same MIC values and increased stability.[44]
D-V681kwksflktfksavktvlhtalkaisseG+/G−Same MIC values and increased stability.[158]
D-V13KDkwksflktfksakktvlhtalkaisseG+/G−Enhanced AMP activity and increased stability.[158]
D-BMAP-28 *GGlrslGrkilrawkkyGpiivpiiriGeLeishmania major (protozoa)Enhanced AMP activity.[151]
RI-BMAP-28 *GiriipviipGykkwarlikrGlsrlGGreLeishmania major (protozoa)Enhanced AMP activity.[151]
D-Ano-NH2GllkriktlleG+/G−Same MIC values.[159]
D-GL13KGkiiklkaslklleG+Enhanced AMP activity.[160,161]
retro-HHC10WRIWKWWRKrG+/G−Same MIC values.[162]
inverso-HHC10krwwkwirweG+/G−Same MIC values and increased stability.[162]
retro-inverso-HHC10wriwkwwrkreG+/G−Same MIC values.[162]
inverso-CysHHC10ckrwwkwirweG+/G−Same MIC values.[162,163,164]
IK8-all DirikirikeG+/G−Enhanced AMP activity.[165]
IK12-all DirvkirvkirvkeG+/G−Enhanced AMP activity.[165]
D-MPIidwkklldaakqileG+/G−Same MIC values.[166]
r-CAMELLVKLVAGIKKFLKWKrG+/G−Curtailed AMP activity[157]
r-citropin 1.1LGGIVSAVKKIVDFLGrG+/G−Curtailed AMP activity.[157]
r-omigananKRRWPWWPWRLIrG+/G−Enhanced AMP activity.[157,167]
r-pexigananKKLIKVFAKGFKKAKKLFKGIGrG−Same MIC values.[157]
r-temporin ALIGSLVRGILPLFrG+Curtailed AMP activity[157]
D-RR4wlrrikawlrrikaeG−Same MIC values.[168]
RI-73lwGvwrrvidwlrreG+ (S. aureus)Same MIC values.[150]
D2Dkk(1nal)fk(1nal)knleeG+/G−Enhanced AMP activity.[169]
(ri)-r(P)ApoBSPro *GsllkvprkpspiifklkGpklavhpreG−Same MIC values and increased stability.[170]
Ctn retro *FPITVGIVMPKKFIKKLRKKVSKKVKKFFKKFRKrG−Curtailed AMP activity[31]
Ctn enantio *krfkkffkkvkksvkkrlkkifkkpmviGvtipfeG−Curtailed AMP activity and increased stability.[31]
Ctn retroenantio *fpitvGivmpkkfikklrkkvskkvkkffkkfrkreG−Curtailed AMP activity and increased stability.[31]
Ctn[15-34] retro *FPITVGIVMPKKFIKKLRKKrG−Same MIC values.[31]
Ctn[15-34] enantio *kkrlkkifkkpmviGvtipfeG−Same MIC values and increased stability.[31]
Ctn[15-34] retroenantio *fpitvGivmpkkfikklrkkreG−Same MIC values and increased stability.[31]
D-CaerinGllsvlGsvakhvlphvvpviaehleG+Enhanced AMP activity.[171]
a AMP name provided in the publication; b uppercase letters represent L-amino acids, while lowercase letters represent D-amino acids; c relative to original peptide; * peptides described as both AMPs and ACPs. Abbreviations: G+, Gram-positive; G−, Gram-negative; 1nal; 3-(1-naphthyl)-D-alanine; e, enantio; r, retro; re, retroenantio.

4.2. CPP Topoisomers and Drug Delivery Challenges

CPPs, also known as Trojan horse, protein translocation domains (PTD), or membrane translocation sequences (MTS), are peptides with the ability to traverse membranes —including barriers such as the gastrointestinal barrier [172], the blood–placental barrier [173], or the highly restricted blood–brain barrier (BBB) [174]—and deliver therapeutically active payloads across the boundary. CPPs are about 5–30 residues long, usually linear, (although cyclic versions have been described [27,69]), and mostly composed of L-amino acids, with some sequences including D-residues or L,D-combinations [69]. CPPs have been classified as cationic [175,176,177,178], amphipathic [179,180,181], hydrophobic [182,183,184], or anionic [185,186], with cargoes that include small molecules (drugs and dyes), proteins, nanoparticles, or genetic material [71].
Although most CPPs are considered safe, their (mostly) cationic nature may pose toxicity issues to some organs and tissues; hence it is important to develop CPPs safer for humans. Moreover, many CPPs degrade fast in biological fluids, proteolytic stability thus being critical for their efficacy as drug delivery vehicles [187,188]. The design of improved CPP platforms has addressed these issues by means of in vivo toxicity screens [189] and/or resorting to topoisomers (e or re versions) to avoid protease degradation. Some recent examples are shown in Table 2.
The first CPP identified was a fragment of the trans-activator of transcription (Tat) protein [190]. A detailed study of the protein defined Tat [48–60] as the most effective fragment [69]. Wender et al. studied a shortened version, Tat [49–57], as well as the corresponding e and re versions and found that these topoisomers were more effective than the canonic L-version in terms of entering Jurkat cells [191]. Similarly, Seisel et al. investigated the e and re versions of the Tat [48–60] sequence using iCal36, a peptide intended for patients with cystic fibrosis, as cargo, with the re topoisomer found to be most appropriate [192].
Another CPP discovered soon after Tat was penetratin, derived from the DNA binding domain of the Antennapedia protein [69]. Nielsen et al. showed that oral coadministration of insulin with L- or D-penetratin lowers blood glucose levels significantly more than insulin alone. D-penetratin was most effective due to its lower level of degradation by proteases [193]. Similar results in terms of blood glucose reduction were reported by Kamei et al. [194] with respect to D-PenetraMax, another e penetratin topoisomeric analogue.
Following the discovery of Tat and penetratin, oligoarginines such as R6 or R8 have also been recognized as effective CPPs [177]. Studies of e/re topoisomeric poly-Arg peptides again showed their effectiveness in penetration. For instance, Garcia et al. showed that r6 and r8 in combination with lauric acid enhanced insulin transport through the gastrointestinal tract by around 30–40% in Caco-2/HT-29 cells [195]. Similar work conducted by Kamei et al. [196] showed that r8 improved on R8 with regard to intravenous co-administration with insulin.
A rather interesting example of CPP topoisomerism is DAngiopep, the re version of Angiopep-2, an artificial peptide that crosses the BBB [197]. By combining DAngiopep with nanoparticles and a dye, effective BBB penetration and glioma targeting within the brain was demonstrated. This approach could prove particularly valuable in facilitating tumor identification during surgical procedures [198]. Another relevant BBB-crossing CPP, the THR peptide, discovered through phage display, can successfully interact with the human transferrin receptor. The protease vulnerability of the THR peptide [199] has been successfully overcome by its re version [28], used to transport nanoparticles inside the brain [200].
Another example worth mentioning is mastoparan (MP), a 14-residue peptide isolated from the venom of Vespula lewisii. Topoisomers of MP and its MitP analogue (with α-aminoisobutyric acid instead of alanine in position 10) were shown to be effective CPPs, with the e version of MP displaying the highest translocation efficacy and protease resistance [201].
To investigate D- and L-CPP entrance in cells, Verdurmen et al. compared the effect of three peptides (hLf, penetratin, and nonaarginine) in their canonic and e forms on three different cell lines (HeLa, Mc57 fibrosarcoma, and Jurkat T) [202]. They observed distinct differences in uptake efficiency between the two enantiomers at low concentrations, which could be attributed to a two-step internalization process. A first step was binding to heparan sulphates (HS) [203], the receptors of Arg-rich CPPs (especially of L-versions [202,204]), followed by internalization via endocytosis. Notably, the presence of HS appeared to hinder the efficiency of the second step for D-CPPs at lower concentrations. In contrast, at higher concentrations, D-enantiomers became more efficient, as the dominant mechanism was direct penetration [187]. These findings underscore the importance of stereochemistry, mechanisms of action, and applied concentration when studying CPPs.
Another interesting example is DCDX, the re version of CDX, a 16-residue peptide derived from the II loop of snake neurotoxin candoxin [205], with higher transcytosis observed in BBB models compared to the protease labile L-version [206]. Han et al. proved that DCDX combined with liposomes crosses the BBB in vitro, following an energy-dependent lipid raft/caveolae- and clathrin-dependent pathway [207].
Yet another example of a successful topoisomer CPP engineered into a druggable candidate, the 16-residue peptide wliymyayvaGilkrw (DRT-017), was developed by our group. It embodies the re version of a transmembrane (TM5) motif of the CB1 cannabinoid receptor (CB1R) fused with the e version of a BBB shuttle. The co-administration of the peptide and a cannabinoid preserves THC-induced analgesia but minimizes side effects (i.e., cognitive impairment) by restricting, both in vitro and in vivo, the formation of a heterodimer between CB1R and the serotonin 5HT2A receptor responsible for the unwanted side effect [208].
Table 2. Examples of recent CPP topoisomers.
Table 2. Examples of recent CPP topoisomers.
Name aSequence bCargoTopoisomer ClassIn Vitro Cell Lines TestedIn Vivo ModelsReferences
Tat-DGrkkrrqrrrppqPeptides, small moleculeseCaco-2, ATCC, HTB-37, Calu-3, ATCC, HTB55-[192,209]
TatriqpprrrqrrkkrGPeptides, small moleculesreCaco-2, ATCC, HTB-37, Calu-3, ATCC, HTB55-[192]
D-Tat 49-57rkkrrqrrrSmall molecules, nanoparticles, proteins.eJurkat T -[191,210]
D-Tat 57-49rrrqrrkkrSmall molecules, nanoparticles, proteins.reJurkat T, MCF-7 Peritoneal-TA3/St tumor-bearing mice[191,210,211]
D-dfTAT(ckrkkrrqrrrG)2
(disulfide bridge)
-eHeLa, MCH58 and HDF-[212]
D-penetratinrqikiwfqnrrmkwkkInsulineCaco-2, HepG2 and IEC-6o.a. mice[193,194,213]
D-penetraMaxkwfkiqmqirrwknkrInsuline-o.a. mice[194]
D-polyarginine(rrrr)xInsulin, peptide, nucleic acide, reCaco-2, HT-29, Fetal hepatocytes, HeLa, Mc57 fibrosarcoma and Jurkat Ti.v. male Sprague-Dawley rats, i.v. and i.p. ducklings [195,196,202,214,215,216]
DAngiopepyeetkfnnrkGrsGGyfftNanoparticlesrebEnd.3 and U87Glioma model in mice[198,217]
THRrepwvpswmpprhtSmall molecules, nanoparticlesrebEnd.3 cells-[28,199,200,218]
JNKDtdqsrpvqpflnlttprkprpprrrqrrkkrG-rePrimary cortical neuronal culturesi.p. Sprague-Dawley P7 rat pups[216,219]
D-R9F2CrrrrrrrrrffcOligonucleotideseHeLa-[220]
D-SAP(vrlppp)3-eHeLa-[221]
β-syn 36DGvlyvGsktr-eSH-SY5Y5Drosophila model, mixed with the food[222]
retro-inverso β-syn 36rtksGvylvG-reSH-SY5Y5Drosophila model, mixed with the food[222]
RI-HER-2vcsaGftyrGepnpmseftdtnytvlapchl (cyclic disulfide form)-reBT-474, SK-BR-3, MDA-468, and TS/ACombination with RI-VEGF-P4, s.c. mice[223]
RI-VEGF-P4fsmecimrikphqGqhiGcqmti (cyclic disulfide form)-reBT-474, SK-BR-3, MDA-468, and TS/ACombination with RI-HER-2, s.c. mice[223]
iMPinlkalaalakkilPeptideseU373MG-[201]
rMPLIKKALAALAKLNIPeptidesrU373MG-[201]
riMPlikkalaalaklniPeptidesreU373MG-[201]
iMitPinlkklakl(Aib)kkilPeptideseU373MG-[201]
rMitPLIKK(Aib)ALAALAKLNIPeptidesrU373MG-[201]
riMitPlikk(Aib)lkalkklniPeptidesreU373MG-[201]
riDOMqqrkrkiwsilaplGttlvklvaGic-reLipid vesicles-[224]
R.I.-p1932qpkGppppGqpknGGqpppG-rePE/CA PJ15 and hGF-[225]
NrTP5ykqchkkGGkkGsG-eHeLa and BHK21 -[226]
RI-C2arkGrsntfidcsiRNAreM17, PC12, L929, or S2103 i.v. C57BL/6 mice model[29]
D-K4kkkkPeptide nucleic acide, reFetal hepatocytesi.v. and i.p. ducklings[215]
RICKkwllrwlsrllrwlarwlGNanoparticlesreU87-[227]
D-CADY-KGlwralwrllrslwrllwkNanoparticleseU87-[227]
retro-D-HAIhrpyiah-reB cellsi.p. mice, s.c. rabbit[228]
DVSsvafpsyrhrsfwsvSmall moleculesreHUVEC and U87i.v. intracranial tumor model mice[229]
CHA-061hsfriitsitlrGrrrrrrrrr-reHK-2Streptozotocin (STZ)-induced diabetes mouse model (via i.p. injection)[230]
DCDXGreirtGraerwsekfLiposomesrebEnd.3 and U87i.p. male mice[207]
DPepH3aGilkrwProteins, antibodieseHBEC-5i-[59]
DRT-017wliymyayvaGilkrw-rebEnd.3 i.v. mouse[208]
DA7RrpplwtaPeptidesreHUVECs and U87s.c. mice[231]
OPBP-1rvysfPeptidesreHUVECs and U87s.c. mice[231]
a CPP name provided in the publication; b Uppercase letters represent L-amino acids, while lowercase letters represent D-amino acids. Abbreviations: i.v., intravenous; i.p., intraperitonial; s.c., subcutaneous; o.a., oral administration; Aib, 2-aminoisobutyric acid; e, enantio; r, retro; re, retroenantio.

4.3. ACP Topoisomers for Mitigating Side Effects in Cancer Treatments

Despite significant advances, the three main methods for cancer treatment, namely, chemotherapy [232,233,234], radiotherapy [235,236,237], and immunotherapy [235,238,239], suffer from low selectivity and serious side effects. In particular, in chemotherapy, the continued use of some antitumor drugs often gives rise to resistance [234]. Therefore, there is a clear need for new therapies that combine selective drug delivery with high toxicity against cancer cells [240].
In terms of successfully tackling the above-mentioned resistance and side effects [241], ACPs appear to be promising candidates. Currently, there are three FDA-approved ACPs, with revenues over USD 1 million [66]: goserelin (PyrHWSYs(tBu)LRP; Pyr—L-pyroglutamyl) and leuprolide (PyrHWSYlLRP), analogues of gonadotropin-releasing hormone (GnRH) [242,243]; and octreotide (fCFwLTCThre; Thre—L-threoninol), an analogue of somatostatin [244]. Since 2000, this list has expanded with entries such as ixazomib, thymalfasin, and mifamurtide [245,246].
ACPs are classified as direct- or indirect-acting based on their mechanism of action [247]. Direct-acting ACPs (DAAs) specifically target cancer cells, typically attaching to molecules that are either unique or overexpressed [248]. They are divided into five subclasses [241]: (a) CPPs acting as cytotoxic drug carriers [249], as discussed in Section 4.2.; (b) pore-forming peptides, inducing apoptosis or necrosis by interacting with phosphatidylserine anionic lipids exposed on the outer membrane of cancer cells [250]; (c) peptide inhibitors of signal transduction cascades, either inhibiting mitogenic signals or restoring the activity of tumor-suppressive proteins like p53 [251]; (d) cell-cycle-inhibitory peptides, modulating cyclin and cyclin-dependent kinase activity [252]; and (e) apoptosis-inducing peptides, inhibiting anti-apoptotic proteins from the Bcl-2 family [252].
Indirect-acting ACPs can influence the tumor environment or immune response in order to target cancer cells and are subdivided into two classes [241]: (a) immune-stimulating peptides, also referred to as peptide cancer vaccines [253], triggering immune cells such as T-cells to act as natural killers against cancer cells, and (b) analogues of hormone-releasing peptides, inhibiting the proliferation of hormone-stimulated tumor cells. These classes include the above-referred GnRH analogues goserelin and leuprolide [254,255] as well as octreotide and other somatostatin analogues [256].
Upon comparing the modes of action of pore-forming peptides, DAAs, and AMPs, it becomes clear that they exhibit similarities in terms of electrostatic interactions [257]; therefore, many AMPs tend to also be explored for their role as potential ACPs. An example, discussed in Table 1, is Ctn and its fragment Ctn [15-34]. The anti-tumor activity of both peptides towards several tumor cells has been substantiated [154], and more recently, that of their topoisomers has been, too [31]. Furthermore, the mechanism by which these peptides combat tumor cells has been elucidated. Following initial accumulation on the tumor cell surface, Ctn and Ctn [15-34] enter the tumor cell via either an endocytic pathway or an energy-independent mechanism. Ultimately, Ctn and Ctn [15-34] induce cell death through necrosis or apoptosis [258].
Table 3 presents a comprehensive collection of other topoisomeric peptides that have been explored as ACPs, regardless of their mechanisms of action.
The first, noteworthy case is that of e PMI. PMI is a peptide recognized for its interaction with MDM2 and MDMX, two oncoproteins that negatively regulate the functionality and stability of tumor-suppressing protein p53. Active MDM2 and MDMX cause p53 inactivation and ensuing tumor proliferation [259]. The binding of PMI to MDM2 and MDMX prevents their inhibitory action toward p53, thus ensuring that PMI can exert its tumor suppressing role. Li et al. identified three e peptides, DPMI-α, DPMI-β, and DPMI-γ, capable of binding MDM2 and MDMX but unable to induce p53-dependent cell death due to their non-permeability with respect to the cell membrane [260]. They overcame this hurdle by encapsulating these peptides within liposomes decorated with an integrin-targeting cyclic-RGD peptide. This strategy allowed them to curb glioblastoma activity in vivo via the activation of the p53 pathway.
Another topoisomer ACP worth mentioning is re RPL (named D(LPR) by the authors). RPL is the minimal structural part of CPQPRPLC, a peptide obtained via phage display-library screening [261] that binds both VEGFR-1 (vascular endothelial growth factor) and NRP-1 (neuropilin-1), two essential contributors to angiogenesis whose inhibition can lead to a decrease in tumor size. Giordano et al. developed re RPL as an antiangiogenic drug with promising attributes in vitro and in vivo [262]. Also, Rezazadeh et al. linked re RPL to technetium-99m and showed that the conjugate was a good radioligand for imaging and targeting tumors in vivo [263].
The well-known tripeptide RGD binds specifically to integrin αvβ3, making it an antiangiogenic candidate and a molecular imaging probe [264]. Ramezanizadeh et al. showed that the re version of RGD, i.e., dGr, in either a linear or cyclic form, was useful for tumor imaging and presented higher bioactivity than the natural version [265].
VAP is a seven-residue prostate-homing peptide that binds selectively to GRP78 (glucose-regulated protein 78) [266], which, in turn, regulates VEGF expression and is over-expressed in some tumor cells but remains absent in normal cells. Ran et al. tested e and re versions of VAP and found higher in vivo antitumoral efficacy when compared to the L-counterparts. Furthermore, tumor growth diminished with either e-VAP or re-VAP, concomitant with an increase in body weight, suggesting reduced side effects. The elevated activity of D-amino-acid-containing peptides was attributed to their resistance against proteolytic degradation [267].
As mentioned above, NRP-1 plays a pivotal role in tumorigenesis and is highly expressed within tumor cells. A library of peptides bearing the sequence motif R/K(X)nR/K (with the C-terminal R or K being particularly vital), named CendR, exhibited notable affinity for binding to NRP-1 [268] in the L-conformation. Upon binding, the peptides regulated vascular permeability (enhanced vascular permeability is indispensable for cancer metastasis [269]). Despite their promise, the susceptibility of CendR peptides to protease degradation resulted in low activity. However, Wang et al. [270] showed that topoisomeric CendRs retained functionality. Thus, using RGERPPR as an example, they showed that both topoisomers were superior, with the e version (rGerppr) displaying higher stability and stronger binding to NRP-1 than the L-peptide and the re version (rppreGr) demonstrating heightened tumor-penetrating prowess and stability. This outcome was interpreted using computational simulations revealing that the three D-Arg residues of the e version neatly aligned with the binding pockets of NRP-1, a phenomenon absent in the natural peptide [270].
A last noteworthy example of topoisomeric modulation is FP21, a 21-residue peptide (YTRDLVYGDPARPGIQGTGTF) corresponding to positions 33–53 of human follicle-stimulating hormone (FSH). The FSH receptor (FSHR) is selectively expressed in 50% to 70% of ovarian carcinomas; hence, it is a potential target in treating ovarian tumors [271,272]. Zhang et al. demonstrated that FP21, incorporated into nanoparticles, effectively bound to FSHR but suffered from a limited half-life [273]. The authors overcame this problem using the re version of FP21, which, again formulated as nanoparticles, achieved FSHR binding, improved biostability, and enabled a reduction in tumor size over the L-version, altogether positioning this topoisomer peptide as a promising candidate for treating ovarian cancer [274].
Table 3. Examples of recent ACPs topoisomers.
Table 3. Examples of recent ACPs topoisomers.
Name aSequence bTopoisomer ClassIn Vitro Cell Lines TestedReferences
DPMI-α tnwyanlekllreU87, U251, HCT116 p53+/+ and HCT116 p53−/−[259,275]
DPMI-βtawyanfekllreU87, U251, HCT116 p53+/+ and HCT116 p53−/−[259,275]
DPMI-γdwwplafeallreU87, U251, HCT116 p53+/+ and HCT116 p53−/−[275]
D(LPR)lprreHUVEC[262,263,276]
D-SP5prpspkmGvsvsreSGC7901[277]
retro-tuftsinRPKTrA549 and HL-60[278]
D(RGD)dGrreU87MG, C6and Hela[264,265,279]
RI-BKrfpsfGpprreHUVEC and C6[280]
RI-VAPpavrtnsreU87MG, HUVEC and HL7702 [267]
D-VAPsntrvapeU87MG, HUVEC and HL7702 [267]
DWSWwswGpysreU87 and HUVEC[281]
RI-3yr(Aib)rreRBL-2H3 and RBL-2H3/ETFR, osteosarcoma Saos-2[282]
D(CendR)rppreGr ceHUVEC, C6, U87 and A549 [270]
D(CendR)rGerppr creHUVEC, C6, U87 and A549 [270]
DT7hrpyiahreHepG2[283]
D-FP21ftctGqiGprapdGyvldrtyreHO8910 and HEK 293 T[274,284]
retro-inverso FSH β 33–53 peptideftctkqikprapdkyvldrtyreA2780[285]
[D]-NRC-03GrrkrkwlrriGkGvkiiGGaaldhleHMEC, HDF and HUVEC[286]
RIF7rqwllfireA549[287]
a ACP name provided in the publication; b uppercase letters represent L-amino acids, while lowercase letters represent D-amino acids; c examples of the R/K(X)nR/K motif discussed in the article. Abbreviations: Aib, 2-Aminoisobutyric acid; e, enantio; r, retro; re; retroenantio.

5. Conclusions

Peptides are making substantial inroads into therapeutic application in diverse fields. Peptide-based drug development has evolved from merely reproducing natural motifs to the rational engineering of peptide structures including modifications such as cyclization, stapling, conjugation, or the introduction of modified, non-coded amino acids. With over 80 peptides already on the market and many others in preclinical or clinical stages, the potential of peptides is becoming increasingly manifest across a wide range of indications [288]. A case in point is provided by combinatory anti-infective therapies, where a conventional antibiotic and an AMP are co-administered to foil antimicrobial resistance [289].
Despite such promise, therapeutic peptides continue to face significant challenges, mostly related to their stability in human fluids. In this review, we have shown that topoisomer formulations, particularly those featuring D-amino acids (e and re), often provide viable alternatives to canonical L-versions. This must not be misconstrued, however, as implying that topoisomers do not face challenges in their scope and applications. For instance, retro peptides can present flexibility and adaptability issues that require cautious structural scrutiny [157]. Additionally, enantio versions, particularly when used as CPPs, may also present limitations intrinsically linked to factors such as stereochemistry, mechanisms of action, or applied concentration [202]. These and other limitations notwithstanding, e.g., the higher price (at least twice) of D- compared to L-amino acids, the cost-effectiveness of topoisomer-based approaches is likely to pay off over time and thus furnish valuable tools for advanced peptide-based therapies.

Author Contributions

Conceptualization, A.C.-A., M.G., S.D., T.T. and D.A.; methodology, A.C.-A.; writing—original draft preparation, A.C.-A., M.G., S.D. and T.T.; writing—review and editing, A.C.-A., M.G., S.D., T.T. and D.A.; supervision and funding acquisition, D.A. All authors have read and agreed to the published version of the manuscript.

Funding

Work supported by funds from the La Caixa Health Foundation (project HR17_00409) and European Union (H2020-FETOPEN-2018-2019-2020-01 grant No. 828774).

Data Availability Statement

The data can be shared up on request.

Acknowledgments

The authors thank the Department of Medicine and Life Sciences at UPF, which is a member of the “María de Maeztu” Program for Units of Excellence in R&D (CEX2018-000792-M) developed by the Spanish Ministry of Economy and Competitiveness (MINECO). They are also grateful to Diana Leiva for editing the images used.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Escribá, P.V.; González-Ros, J.M.; Goñi, F.M.; Kinnunen, P.K.J.; Vigh, L.; Sánchez-Magraner, L.; Fernández, A.M.; Busquets, X.; Horváth, I.; Barceló-Coblijn, G. Membranes: A Meeting Point for Lipids, Proteins and Therapies. J. Cell. Mol. Med. 2008, 12, 829–875. [Google Scholar] [CrossRef] [PubMed]
  2. Singer, S.J. Some Early History of Membrane Molecular Biology. Annu. Rev. Physiol. 2004, 66, 1–27. [Google Scholar] [CrossRef]
  3. Lucio, M.; Lima, J.L.F.C.; Reis, S. Drug-Membrane Interactions: Significance for Medicinal Chemistry. Curr. Med. Chem. 2010, 17, 1795–1809. [Google Scholar] [CrossRef]
  4. Seddon, A.M.; Casey, D.; Law, R.V.; Gee, A.; Templer, R.H.; Ces, O. Drug Interactions with Lipid Membranes. Chem. Soc. Rev. 2009, 38, 2509–2519. [Google Scholar] [CrossRef] [PubMed]
  5. Lee, A.C.L.; Harris, J.L.; Khanna, K.K.; Hong, J.H. A Comprehensive Review on Current Advances in Peptide Drug Development and Design. Int. J. Mol. Sci. 2019, 20, 2383. [Google Scholar] [CrossRef]
  6. Erak, M.; Bellmann-Sickert, K.; Els-Heindl, S.; Beck-Sickinger, A.G. Peptide Chemistry Toolbox—Transforming Natural Peptides into Peptide Therapeutics. Bioorg. Med. Chem. 2018, 26, 2759–2765. [Google Scholar] [CrossRef]
  7. Agamennone, M.; Fantacuzzi, M.; Vivenzio, G.; Scala, M.C.; Campiglia, P.; Superti, F.; Sala, M. Antiviral Peptides as Anti-Influenza Agents. Int. J. Mol. Sci. 2022, 23, 11433. [Google Scholar] [CrossRef]
  8. Mazurkiewicz-Pisarek, A.; Baran, J.; Ciach, T. Antimicrobial Peptides: Challenging Journey to the Pharmaceutical, Biomedical, and Cosmeceutical Use. Int. J. Mol. Sci. 2023, 24, 9031. [Google Scholar] [CrossRef]
  9. Kalmouni, M.; Al-Hosani, S.; Magzoub, M. Cancer Targeting Peptides. Cell. Mol. Life Sci. 2019, 76, 2171–2183. [Google Scholar] [CrossRef]
  10. Kurrikoff, K.; Aphkhazava, D.; Langel, Ü. The Future of Peptides in Cancer Treatment. Curr. Opin. Pharmacol. 2019, 47, 27–32. [Google Scholar] [CrossRef]
  11. Lv, Y.; Chen, X.; Chen, Z.; Shang, Z.; Li, Y.; Xu, W.; Mo, Y.; Wang, X.; Xu, D.; Li, S.; et al. Melittin Tryptophan Substitution with a Fluorescent Amino Acid Reveals the Structural Basis of Selective Antitumor Effect and Subcellular Localization in Tumor Cells. Toxins 2022, 14, 428. [Google Scholar] [CrossRef] [PubMed]
  12. Kuwahara, K. The Natriuretic Peptide System in Heart Failure: Diagnostic and Therapeutic Implications. Pharmacol. Ther. 2021, 227, 107863. [Google Scholar] [CrossRef]
  13. Liao, W.; Fan, H.; Davidge, S.T.; Wu, J. Egg White–Derived Antihypertensive Peptide IRW (Ile-Arg-Trp) Reduces Blood Pressure in Spontaneously Hypertensive Rats via the ACE2/Ang (1-7)/Mas Receptor Axis. Mol. Nutr. Food Res. 2019, 63, 1900063. [Google Scholar] [CrossRef] [PubMed]
  14. Ribarič, S. Peptides as Potential Therapeutics for Alzheimer’s Disease. Molecules 2018, 23, 283. [Google Scholar] [CrossRef] [PubMed]
  15. Jeremic, D.; Jiménez-Díaz, L.; Navarro-López, J.D. Past, Present and Future of Therapeutic Strategies against Amyloid-β Peptides in Alzheimer’s Disease: A Systematic Review. Ageing Res. Rev. 2021, 72, 101496. [Google Scholar] [CrossRef] [PubMed]
  16. Sasabe, J.; Suzuki, M. Distinctive Roles of D-Amino Acids in the Homochiral World: Chirality of Amino Acids Modulates Mammalian Physiology and Pathology. Keio J. Med. 2018, 68, 1–16. [Google Scholar] [CrossRef] [PubMed]
  17. Kreil, G. D-Amino Acids in Animal Peptides. Annu. Rev. Biochem. 1997, 66, 337–345. [Google Scholar] [CrossRef]
  18. Amiche, M.; Sagan, S.; Mor, A.; Delfour, A.; Nicolas, P. Dermenkephalin (Tyr-D-Met-Phe-His-Leu-Met-Asp-NH2): A Potent and Fully Specific Agonist for the Delta Opioid Receptor. Mol. Pharmacol. 1989, 35, 774–779. [Google Scholar] [PubMed]
  19. Broccardo, M.; Erspamer, V.; Falconieri, G.; Improta, G.; Linari, G.; Melchiorri, P.; Montecucchi, P. Pharmacological Data on Dermorphins, a New Class of Potent Opioid Peptides from Amphibian Skin. Br. J. Pharmacol. 1981, 73, 625–631. [Google Scholar] [CrossRef] [PubMed]
  20. Montecucchi, P.C.; De Castiglione, R.; Piani, S.; Gozzini, L.; Erspamer, V. Amino Acid Composition and Sequence of Dermorphin, a Novel Opiate-like Peptide from the Skin of Phyllomedusa Sauvagei. Int. J. Pept. Protein Res. 1981, 17, 275–283. [Google Scholar] [CrossRef]
  21. Mor, A.; Delfour, A.; Sagan, S.; Amiche, M.; Pradelles, P.; Rossier, J.; Nicolas, P. Isolation of Dermenkephalin from Amphibian Skin, a High-Affinity (δ-Selective Opioid Heptapeptide Containing a D-Amino Acid Residue. FEBS Lett. 1989, 255, 269–274. [Google Scholar] [CrossRef] [PubMed]
  22. Teramoto, T.; Kuwada, M.; Niidome, T.; Sawada, K.; Nishizawa, Y.; Katayama, K. A Novel Peptide from Funnel Web Spider Venom, ω-Aga-TK, Selectively Blocks P-Type Calcium Channels. Biochem. Biophys. Res. Commun. 1993, 196, 134–140. [Google Scholar] [CrossRef]
  23. Kamatani, Y.; Minakata, H.; Kenny, P.T.M.; Iwashita, T.; Watanabe, K.; Funase, K.; Sun, X.P.; Yongsiri, A.; Kim, K.H.; Novales-Li, P.; et al. Achatin-I, an Endogenous Neuroexcitatory Tetrapeptide from Achatina Fulica Férussac Containing A d-Amino Acid Residue. Biochem. Biophys. Res. Commun. 1989, 160, 1015–1020. [Google Scholar] [CrossRef]
  24. Fujimoto, K.; Kubota, I.; Yasuda-Kamatani, Y.; Minakata, H.; Nomoto, K.; Yoshida, M.; Harada, A.; Muneoka, Y.; Kobayashi, M. Purification of Achatin-I from the Atria of the African Giant Snail, Achatina Fulica, and Its Possible Function. Biochem. Biophys. Res. Commun. 1991, 177, 847–853. [Google Scholar] [CrossRef] [PubMed]
  25. Doti, N.; Mardirossian, M.; Sandomenico, A.; Ruvo, M.; Caporale, A. Recent Applications of Retro-Inverso Peptides. Int. J. Mol. Sci. 2021, 22, 8677. [Google Scholar] [CrossRef] [PubMed]
  26. Hong, S.Y.; Oh, J.E.; Lee, K.H. Effect of D-Amino Acid Substitution on the Stability, the Secondary Structure, and the Activity of Membrane-Active Peptide. Biochem. Pharmacol. 1999, 58, 1775–1780. [Google Scholar] [CrossRef] [PubMed]
  27. Lucana, M.C.; Arruga, Y.; Petrachi, E.; Roig, A.; Lucchi, R.; Oller-Salvia, B. Protease-Resistant Peptides for Targeting and Intracellular Delivery of Therapeutics. Pharmaceutics 2021, 13, 2065. [Google Scholar] [CrossRef] [PubMed]
  28. Prades, R.; Oller-Salvia, B.; Schwarzmaier, S.M.; Selva, J.; Moros, M.; Balbi, M.; Grazú, V.; De La Fuente, J.M.; Egea, G.; Plesnila, N.; et al. Applying the Retro-Enantio Approach To Obtain a Peptide Capable of Overcoming the Blood–Brain Barrier. Angew. Chem. Int. Ed. 2015, 54, 3967–3972. [Google Scholar] [CrossRef]
  29. Javed, H.; Menon, S.A.; Al-Mansoori, K.M.; Al-Wandi, A.; Majbour, N.K.; Ardah, M.T.; Varghese, S.; Vaikath, N.N.; Haque, M.E.; Azzouz, M.; et al. Development of Nonviral Vectors Targeting the Brain as a Therapeutic Approach For Parkinson’s Disease and Other Brain Disorders. Mol. Ther. 2016, 24, 746–758. [Google Scholar] [CrossRef]
  30. Zheng, Y.; Mao, K.; Chen, S.; Zhu, H. Chirality Effects in Peptide Assembly Structures. Front. Bioeng. Biotechnol. 2021, 9, 703004. [Google Scholar] [CrossRef] [PubMed]
  31. Carrera-Aubesart, A.; Defaus, S.; Pérez-Peinado, C.; Sandín, D.; Torrent, M.; Jiménez, M.Á.; Andreu, D. Examining Topoisomers of a Snake-Venom-Derived Peptide for Improved Antimicrobial and Antitumoral Properties. Biomedicines 2022, 10, 2110. [Google Scholar] [CrossRef]
  32. Xi, W.; Hansmann, U.H.E. The Effect of Retro-Inverse D-Amino Acid Aβ-Peptides on Aβ-Fibril Formation. J. Chem. Phys. 2019, 150, 095101. [Google Scholar] [CrossRef]
  33. Neves, V.; Aires-Da-Silva, F.; Morais, M.; Gano, L.; Ribeiro, E.; Pinto, A.; Aguiar, S.; Gaspar, D.; Fernandes, C.; Correia, J.D.G.; et al. Novel Peptides Derived from Dengue Virus Capsid Protein Translocate Reversibly the Blood-Brain Barrier through a Receptor-Free Mechanism. ACS Chem. Biol. 2017, 12, 1257–1268. [Google Scholar] [CrossRef] [PubMed]
  34. Schrödinger, L. The PyMOL Molecular Graphics System, Version 1.2r3pre; Schrodinger LLC: New York, NY, USA, 2008; 3p. [Google Scholar]
  35. Stewart, J.M.; Woolley, D.W. All-D-Bradykinin and the Problem of Peptide Antimetabolites. Nature 1965, 206, 619–620. [Google Scholar] [CrossRef]
  36. Vogler, K.; Studer, R.O.; Lergier, W.; Lanz, P. Synthese von All-D-Val5-Angiotensin II-Asp1-β-Amid. Helv. Chim. Acta 1965, 48, 1407–1414. [Google Scholar] [CrossRef] [PubMed]
  37. Shemyakin, M.M.; Ovchinnikov, Y.A.; Ivanov, V.T.; Evstratov, A.V. Topochemical Approach in Studies of the Structure-Activity Relation: Enantio-Enniatin B. Nature 1967, 213, 412–413. [Google Scholar] [CrossRef]
  38. Merrifield, R.B.; Juvvadi, P.; Andreu, D.; Ubach, J.; Boman, A.; Boman, H.G. Retro and Retroenantio Analogs of Cecropin-Melittin Hybrids. Proc. Natl. Acad. Sci. USA 1995, 92, 3449–3453. [Google Scholar] [CrossRef]
  39. Shemyakin, M.M.; Ovchinnikov, Y.A.; Ivanov, V.T. Topochemical Investigations on Peptide Systems. Angew. Chem. Int. Ed. Engl. 1969, 8, 492–499. [Google Scholar] [CrossRef]
  40. Merrifield, E.L.; Mitchell, S.A.; Ubach, J.; Boman, H.G.; Andreu, D.; Merrifield, R.B. D-Enantiomers of 15-Residue Cecropin A-Melittin Hybrids. Int. J. Pept. Protein Res. 1995, 46, 214–220. [Google Scholar] [CrossRef] [PubMed]
  41. Wade, D.; Andreu, D.; Mitchell, S.A.; Silveira, A.M.V.; Boman, A.; Boman, H.G.; Merrifield, R.B. Antibacterial Peptides Designed as Analogs or Hybrids of Cecropins and Melittin. Int. J. Pept. Protein Res. 1992, 40, 429–436. [Google Scholar] [CrossRef]
  42. Wade, D.; Boman, A.; Wåhlint, B.; Drain, C.M.; Andreu, D.; Boman, H.G.; Merrifield, R.B. All-D Amino Acid-Containing Channel-Forming Antibiotic Peptides. Proc. Natl. Acad. Sci. USA 1990, 87, 4761–4765. [Google Scholar] [CrossRef]
  43. Vunnam, S.; Juvvadi, P.; Rotondi, K.S.; Merrifield, R.B. Synthesis and Study of Normal, Enantio, Retro, and Retroenantio Isomers of Cecropin A-Melittin Hybrids, Their End Group Effects and Selective Enzyme Inactivation. J. Pept. Res. 1998, 51, 38–44. [Google Scholar] [CrossRef] [PubMed]
  44. Staubitz, P.; Peschel, A.; Nieuwenhuizen, W.F.; Otto, M.; Götz, F.; Jung, G.; Jack, R.W. Structure–Function Relationships in the Tryptophan-Rich, Antimicrobial Peptide Indolicidin. J. Pept. Sci. 2001, 7, 552–564. [Google Scholar] [CrossRef]
  45. Ruvo, M.; Fassina, G. End-Group Modified Retro-Inverso Isomers of Tripeptide Oxytocin Analogues: Binding to Neurophysin II and Enhancement of Its Self-Association Properties. Int. J. Pept. Protein Res. 1995, 45, 356–365. [Google Scholar] [CrossRef] [PubMed]
  46. Ho, S.; Pothoulakis, C.; Wai Koon, H. Antimicrobial Peptides and Colitis. Curr. Pharm. Des. 2013, 19, 40–47. [Google Scholar]
  47. Savini, F.; Loffredo, M.R.; Troiano, C.; Bobone, S.; Malanovic, N.; Eichmann, T.O.; Caprio, L.; Canale, V.C.; Park, Y.; Mangoni, M.L.; et al. Binding of an Antimicrobial Peptide to Bacterial Cells: Interaction with Different Species, Strains and Cellular Components. Biochim. Biophys. Acta-Biomembr. 2020, 1862, 183291. [Google Scholar] [CrossRef] [PubMed]
  48. 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]
  49. Epand, R.M.; Walker, C.; Epand, R.F.; Magarvey, N.A. Molecular Mechanisms of Membrane Targeting Antibiotics. Biochim. Biophys. Acta-Biomembr. 2016, 1858, 980–987. [Google Scholar] [CrossRef]
  50. Kang, S.J.; Park, S.J.; Mishig-Ochir, T.; Lee, B.J. Antimicrobial Peptides: Therapeutic Potentials. Expert Rev. Anti. Infect. Ther. 2014, 12, 1477–1486. [Google Scholar] [CrossRef]
  51. Thapa, R.K.; Diep, D.B.; Tønnesen, H.H. Topical Antimicrobial Peptide Formulations for Wound Healing: Current Developments and Future Prospects. Acta Biomater. 2020, 103, 52–67. [Google Scholar] [CrossRef]
  52. Lai, Z.; Yuan, X.; Chen, H.; Zhu, Y.; Dong, N.; Shan, A. Strategies Employed in the Design of Antimicrobial Peptides with Enhanced Proteolytic Stability. Biotechnol. Adv. 2022, 59, 107962. [Google Scholar] [CrossRef]
  53. Derossi, D.; Joliot, A.H.; Chassaing, G.; Prochiantz, A. The Third Helix of the Antennapedia Homeodomain Translocates through Biological Membranes. J. Biol. Chem. 1994, 269, 10444–10450. [Google Scholar] [CrossRef]
  54. Derossi, D.; Calvet, S.; Trembleau, A.; Brunissen, A.; Chassaing, G.; Prochiantz, A. Cell Internalization of the Third Helix of the Antennapedia Homeodomain Is Receptor-Independent. J. Biol. Chem. 1996, 271, 18188–18193. [Google Scholar] [CrossRef]
  55. De Figueiredo, I.R.; Freire, J.M.; Flores, L.; Veiga, A.S.; Castanho, M.A.R.B. Cell-Penetrating Peptides: A Tool for Effective Delivery in Gene-Targeted Therapies. IUBMB Life 2014, 66, 182–194. [Google Scholar] [CrossRef]
  56. Kardani, K.; Milani, A.; Shabani, S.H.; Bolhassani, A. Cell Penetrating Peptides: The Potent Multi-Cargo Intracellular Carriers. Expert Opin. Drug Deliv. 2019, 16, 1227–1258. [Google Scholar] [CrossRef]
  57. Porosk, L.; Gaidutšik, I.; Langel, Ü. Approaches for the Discovery of New Cell-Penetrating Peptides. Expert Opin. Drug Discov. 2020, 16, 553–565. [Google Scholar] [CrossRef]
  58. Sajid, M.I.; Moazzam, M.; Stueber, R.; Park, S.E.; Cho, Y.; Malik, N.u.A.; Tiwari, R.K. Applications of Amphipathic and Cationic Cyclic Cell-Penetrating Peptides: Significant Therapeutic Delivery Tool. Peptides 2021, 141, 170542. [Google Scholar] [CrossRef]
  59. Cavaco, M.; Valle, J.; da Silva, R.; Correia, J.D.G.; Castanho, M.A.R.B.; Andreu, D.; Neves, V. D PepH3, an Improved Peptide Shuttle for Receptor-Independent Transport Across the Blood-Brain Barrier. Curr. Pharm. Des. 2020, 26, 1495–1506. [Google Scholar] [CrossRef]
  60. Guichard, G.; Benkirane, N.; Zeder-Lutz, G.; Van Regenmortel, M.H.; Briand, J.P.; Muller, S. Antigenic Mimicry of Natural L-Peptides with Retro-Inverso-Peptidomimetics. Proc. Natl. Acad. Sci. USA 1994, 91, 9765–9769. [Google Scholar] [CrossRef]
  61. Röckendorf, N.; Nehls, C.; Gutsmann, T. Design of Membrane Active Peptides Considering Multi-Objective Optimization for Biomedical Application. Membranes 2022, 12, 180. [Google Scholar] [CrossRef]
  62. Drexelius, M.; Reinhardt, A.; Grabeck, J.; Cronenberg, T.; Nitsche, F.; Huesgen, P.F.; Maier, B.; Neundorf, I. Multistep Optimization of a Cell-Penetrating Peptide towards Its Antimicrobial Activity. Biochem. J. 2021, 478, 63–78. [Google Scholar] [CrossRef]
  63. Hollmann, A.; Martinez, M.; Maturana, P.; Semorile, L.C.; Maffia, P.C. Antimicrobial Peptides: Interaction with Model and Biological Membranes and Synergism with Chemical Antibiotics. Front. Chem. 2018, 6, 204. [Google Scholar] [CrossRef]
  64. Nyström, L.; Malmsten, M. Membrane Interactions and Cell Selectivity of Amphiphilic Anticancer Peptides. Curr. Opin. Colloid Interface Sci. 2018, 38, 1–17. [Google Scholar] [CrossRef]
  65. Gawde, U.; Chakraborty, S.; Waghu, F.H.; Barai, R.S.; Khanderkar, A.; Indraguru, R.; Shirsat, T.; Idicula-Thomas, S. CAMPR4: A Database of Natural and Synthetic Antimicrobial Peptides. Nucleic Acids Res. 2023, 51, D377–D383. [Google Scholar] [CrossRef]
  66. Tripathi, A.K.; Vishwanatha, J.K. Role of Anti-Cancer Peptides as Immunomodulatory Agents: Potential and Design Strategy. Pharmaceutics 2022, 14, 2686. [Google Scholar] [CrossRef]
  67. Matsuzaki, K. Why and How Are Peptide–Lipid Interactions Utilized for Self-Defense? Magainins and Tachyplesins as Archetypes. Biochim. Biophys. Acta-Biomembr. 1999, 1462, 1–10. [Google Scholar] [CrossRef]
  68. 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]
  69. Gallo, M.; Defaus, S.; Andreu, D. 1988–2018: Thirty Years of Drug Smuggling at the Nano Scale. Challenges and Opportunities of Cell-Penetrating Peptides in Biomedical Research. Arch. Biochem. Biophys. 2019, 661, 74–86. [Google Scholar] [CrossRef]
  70. Sánchez-Navarro, M.; Giralt, E. Peptide Shuttles for Blood-Brain Barrier Drug Delivery. Pharmaceutics 2022, 14, 1874. [Google Scholar] [CrossRef]
  71. Todorovski, T.; Kalafatovic, D.; Andreu, D. Antiviral Peptide-Based Conjugates: State of the Art and Future Perspectives. Pharmaceutics 2023, 15, 357. [Google Scholar] [CrossRef]
  72. Ramaker, K.; Henkel, M.; Krause, T.; Röckendorf, N.; Frey, A. Cell Penetrating Peptides: A Comparative Transport Analysis for 474 Sequence Motifs. Drug Deliv. 2018, 25, 928–937. [Google Scholar] [CrossRef] [PubMed]
  73. He, L.; Sayers, E.J.; Watson, P.; Jones, A.T. Contrasting Roles for Actin in the Cellular Uptake of Cell Penetrating Peptide Conjugates. Sci. Rep. 2018, 8, 7318. [Google Scholar] [CrossRef] [PubMed]
  74. Gestin, M.; Dowaidar, M.; Langel, Ü. Uptake Mechanism of Cell-Penetrating Peptides. In Peptides and Peptide-Based Biomaterials and Their Biomedical Applications; Sunna, A., Care, A., Bergquist, P., Eds.; Springer: New York, NY, USA, 2017; Volume 1030, pp. 255–264. [Google Scholar]
  75. Mueller, J.; Kretzschmar, I.; Volkmer, R.; Boisguerin, P. Comparison of Cellular Uptake Using 22 CPPs in 4 Different Cell Lines. Bioconjug. Chem. 2008, 19, 2363–2374. [Google Scholar] [CrossRef]
  76. Walrant, A.; Cardon, S.; Burlina, F.; Sagan, S. Membrane Crossing and Membranotropic Activity of Cell-Penetrating Peptides: Dangerous Liaisons? Acc. Chem. Res. 2017, 50, 2968–2975. [Google Scholar] [CrossRef] [PubMed]
  77. Kulbacka, J.; Choromańska, A.; Rossowska, J.; Weżgowiec, J.; Saczko, J.; Rols, M.P. Cell Membrane Transport Mechanisms: Ion Channels and Electrical Properties of Cell Membranes. In Transport across Natural and Modified Biological Membranes and Its Implications in Physiology and Therapy; Kulbacka, J., Satkauskas, S., Eds.; Springer: New York, NY, USA, 2017; Volume 227, pp. 39–58. [Google Scholar]
  78. Sugano, K.; Kansy, M.; Artursson, P.; Avdeef, A.; Bendels, S.; Di, L.; Ecker, G.F.; Faller, B.; Fischer, H.; Gerebtzoff, G.; et al. Coexistence of Passive and Carrier-Mediated Processes in Drug Transport. Nat. Rev. Drug Discov. 2010, 9, 597–614. [Google Scholar] [CrossRef] [PubMed]
  79. Lin, L.; Yee, S.W.; Kim, R.B.; Giacomini, K.M. SLC Transporters as Therapeutic Targets: Emerging Opportunities. Nat. Rev. Drug Discov. 2015, 14, 543–560. [Google Scholar] [CrossRef] [PubMed]
  80. Perland, E.; Fredriksson, R. Classification Systems of Secondary Active Transporters. Trends Pharmacol. Sci. 2017, 38, 305–315. [Google Scholar] [CrossRef]
  81. Alleva, K.; Chara, O.; Amodeo, G. Aquaporins: Another Piece in the Osmotic Puzzle. FEBS Lett. 2012, 586, 2991–2999. [Google Scholar] [CrossRef]
  82. Yukutake, Y.; Hirano, Y.; Suematsu, M.; Yasui, M. Rapid and Reversible Inhibition of Aquaporin-4 by Zinc. Biochemistry 2009, 48, 12059–12061. [Google Scholar] [CrossRef]
  83. Gerbeau, P.; Amodeo, G.; Henzler, T.; Santoni, V.; Ripoche, P.; Maurel, C. The Water Permeability of Arabidopsis Plasma Membrane Is Regulated by Divalent Cations and PH. Plant J. 2002, 30, 71–81. [Google Scholar] [CrossRef] [PubMed]
  84. Cleal, K.; He, L.; Watson, P.D.; Jones, A.T. Endocytosis, Intracellular Traffic and Fate of Cell Penetrating Peptide Based Conjugates and Nanoparticles. Curr. Pharm. Des. 2013, 19, 2878–2894. [Google Scholar] [CrossRef] [PubMed]
  85. Yang, N.J.; Hinner, M.J. Getting across the Cell Membrane: An Overview for Small Molecules, Peptides, and Proteins. In Site-Specific Protein Labeling; Gautier, A., Hinner, M., Eds.; Humana Press: New York, NY, USA, 2015; Volume 1266, pp. 29–53. [Google Scholar]
  86. Mayor, S.; Pagano, R.E. Pathways of Clathrin-Independent Endocytosis. Nat. Rev. Mol. Cell Biol. 2007, 8, 603–612. [Google Scholar] [CrossRef] [PubMed]
  87. Desale, K.; Kuche, K.; Jain, S. Cell-Penetrating Peptides (CPPs): An Overview of Applications for Improving the Potential of Nanotherapeutics. Biomater. Sci. 2021, 9, 1153–1188. [Google Scholar] [CrossRef] [PubMed]
  88. Madani, F.; Lindberg, S.; Langel, U.; Futaki, S.; Gräslund, A.; Langel, Ü.; Futaki, S.; Gräslund, A. Mechanisms of Cellular Uptake of Cell-Penetrating Peptides. J. Biophys. 2011, 2011, 414729. [Google Scholar] [CrossRef] [PubMed]
  89. Stewart, M.P.; Langer, R.; Jensen, K.F. Intracellular Delivery by Membrane Disruption: Mechanisms, Strategies, and Concepts. Chem. Rev. 2018, 118, 7409–7531. [Google Scholar] [CrossRef] [PubMed]
  90. Last, N.B.; Schlamadinger, D.E.; Miranker, A.D. A Common Landscape for Membrane-Active Peptides. Protein Sci. 2013, 22, 870–882. [Google Scholar] [CrossRef]
  91. Sun, D.; Forsman, J.; Woodward, C.E. Current Understanding of the Mechanisms by Which Membrane-Active Peptides Permeate and Disrupt Model Lipid Membranes. Curr. Top. Med. Chem. 2016, 16, 170–186. [Google Scholar] [CrossRef] [PubMed]
  92. 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]
  93. Zasloff, M. Antimicrobial Peptides of Multicellular Organisms. Nature 2002, 415, 389–395. [Google Scholar] [CrossRef] [PubMed]
  94. Yeaman, M.R.; Yount, N.Y. Mechanisms of Antimicrobial Peptide Action and Resistance. Pharmacol. Rev. 2003, 55, 27–55. [Google Scholar] [CrossRef]
  95. Brogden, K.A. Antimicrobial Peptides: Pore Formers or Metabolic Inhibitors in Bacteria? Nat. Rev. Microbiol. 2005, 3, 238–250. [Google Scholar] [CrossRef]
  96. Lee, T.-H.; Hall, K.N.; Aguilar, M.-I. Antimicrobial Peptide Structure and Mechanism of Action: A Focus on the Role of Membrane Structure. Curr. Top. Med. Chem. 2016, 16, 25–39. [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]
  98. Hale, J.D.F.; Hancock, R.E.W. Alternative Mechanisms of Action of Cationic Antimicrobial Peptides on Bacteria. Expert Rev. Anti. Infect. Ther. 2007, 5, 951–959. [Google Scholar] [CrossRef]
  99. Lu, J.; Xu, H.; Xia, J.; Ma, J.; Xu, J.; Li, Y.; Feng, J. D- and Unnatural Amino Acid Substituted Antimicrobial Peptides with Improved Proteolytic Resistance and Their Proteolytic Degradation Characteristics. Front. Microbiol. 2020, 11, 2869. [Google Scholar] [CrossRef] [PubMed]
  100. Cavaco, M.; Andreu, D.; Castanho, M.A.R.B. The Challenge of Peptide Proteolytic Stability Studies: Scarce Data, Difficult Readability, and the Need for Harmonization. Angew. Chem. 2021, 133, 1710–1712. [Google Scholar] [CrossRef]
  101. Klompas, M. Overuse of Broad-Spectrum Antibiotics for Pneumonia. JAMA Intern. Med. 2020, 180, 485–486. [Google Scholar] [CrossRef] [PubMed]
  102. Lee, R.A.; Centor, R.M.; Humphrey, L.L.; Jokela, J.A.; Andrews, R.; Qaseem, A. Appropriate Use of Short-Course Antibiotics in Common Infections: Best Practice Advice from the American College of Physicians. Ann. Intern. Med. 2021, 174, 822–827. [Google Scholar] [CrossRef]
  103. Interagency Coordination Group on Antimicrobial Resistance. No Time to Wait: Securing the Future from Drug-Resistant Infections. Report to the Secretary-General of the United Nations; WHO: Geneva, Switzerland, 2019. [Google Scholar]
  104. Yu, T.; Jiang, G.; Gao, R.; Chen, G.; Ren, Y.; Liu, J.; van der Mei, H.C.; Busscher, H.J. Circumventing Antimicrobial-Resistance and Preventing Its Development in Novel, Bacterial Infection-Control Strategies. Expert Opin. Drug Deliv. 2020, 17, 1151–1164. [Google Scholar] [CrossRef]
  105. Ciotti, M.; Ciccozzi, M.; Pieri, M.; Bernardini, S. The COVID-19 Pandemic: Viral Variants and Vaccine Efficacy. Crit. Rev. Clin. Lab. Sci. 2021, 59, 66–75. [Google Scholar] [CrossRef]
  106. Hadj Hassine, I. COVID-19 Vaccines and Variants of Concern: A Review. Rev. Med. Virol. 2022, 32, e2313. [Google Scholar] [CrossRef] [PubMed]
  107. Chams, N.; Chams, S.; Badran, R.; Shams, A.; Araji, A.; Raad, M.; Mukhopadhyay, S.; Stroberg, E.; Duval, E.J.; Barton, L.M.; et al. COVID-19: A Multidisciplinary Review. Front. Public Health 2020, 8, 383. [Google Scholar] [CrossRef] [PubMed]
  108. Parasher, A. COVID-19: Current Understanding of Its Pathophysiology, Clinical Presentation and Treatment. Postgrad. Med. J. 2021, 97, 312–320. [Google Scholar] [CrossRef]
  109. Department of Health and Human Services. COVID-19 U.S. Impact on Antimicrobial Resistance, Special Report 2022; Department of Health and Human Services, CDC: Atlanta, GA, USA, 2022. [Google Scholar]
  110. Chedid, M.; Waked, R.; Haddad, E.; Chetata, N.; Saliba, G.; Choucair, J. Antibiotics in Treatment of COVID-19 Complications: A Review of Frequency, Indications, and Efficacy. J. Infect. Public Health 2021, 14, 570–576. [Google Scholar] [CrossRef] [PubMed]
  111. Rizvi, S.G.; Ahammad, S.Z. COVID-19 and Antimicrobial Resistance: A Cross-Study. Sci. Total Environ. 2022, 807, 150873. [Google Scholar] [CrossRef]
  112. 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] [PubMed]
  113. Aloke, C.; Achilonu, I. Coping with the ESKAPE Pathogens: Evolving Strategies, Challenges and Future Prospects. Microb. Pathog. 2023, 175, 105963. [Google Scholar] [CrossRef] [PubMed]
  114. Mancuso, G.; Midiri, A.; Gerace, E.; Biondo, C. Bacterial Antibiotic Resistance: The Most Critical Pathogens. Pathogens 2021, 10, 1310. [Google Scholar] [CrossRef] [PubMed]
  115. Manniello, M.D.; Moretta, A.; Salvia, R.; Scieuzo, C.; Lucchetti, D.; Vogel, H.; Sgambato, A.; Falabella, P. Insect Antimicrobial Peptides: Potential Weapons to Counteract the Antibiotic Resistance. Cell. Mol. Life Sci. 2021, 78, 4259–4282. [Google Scholar] [CrossRef]
  116. Moravej, H.; Moravej, Z.; Yazdanparast, M.; Heiat, M.; Mirhosseini, A.; Moosazadeh Moghaddam, M.; Mirnejad, R. Antimicrobial Peptides: Features, Action, and Their Resistance Mechanisms in Bacteria. Microb. Drug Resist. 2018, 24, 747–767. [Google Scholar] [CrossRef] [PubMed]
  117. 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] [PubMed]
  118. Lima, P.G.; Oliveira, J.T.A.; Amaral, J.L.; Freitas, C.D.T.; Souza, P.F.N. Synthetic Antimicrobial Peptides: Characteristics, Design, and Potential as Alternative Molecules to Overcome Microbial Resistance. Life Sci. 2021, 278, 119647. [Google Scholar] [CrossRef]
  119. Valenti, G.E.; Alfei, S.; Caviglia, D.; Domenicotti, C.; Marengo, B. Antimicrobial Peptides and Cationic Nanoparticles: A Broad-Spectrum Weapon to Fight Multi-Drug Resistance Not Only in Bacteria. Int. J. Mol. Sci. 2022, 23, 6108. [Google Scholar] [CrossRef] [PubMed]
  120. Hotchkiss, R.D.; Dubos, R.J. Fractionation of the Bactericidal Agent from Cultures of a Soil Bacillus. J. Biol. Chem. 1940, 132, 791–792. [Google Scholar] [CrossRef]
  121. Dubos, R.J. Studies on a Bactericidal Agent Extracted from a Soil Bacillus: I. Preparation of the Agent. Its Activity In Vitro. J. Exp. Med. 1939, 70, 1–10. [Google Scholar] [CrossRef] [PubMed]
  122. Dubos, R.J. Studies on a Bactericidal Agent Extracted from a Soil Bacillus: II. Protective Effect of the Bactericidal Agent against Experimental Pneumococcus Infections in Mice. J. Exp. Med. 1939, 70, 11–17. [Google Scholar] [CrossRef] [PubMed]
  123. Phoenix, D.A.; Dennison, S.R.; Harris, F. (Eds.) Antimicrobial Peptides: Their History, Evolution, and Functional Promiscuity. In Antimicrobial Peptides; John Wiley & Sons, Ltd.: London, UK, 2013; pp. 1–37. [Google Scholar]
  124. Fernandez De Caleya, R.; Gonzalez-Pascual, B.; García-Olmedo, F.; Carbonero, P. Susceptibility of Phytopathogenic Bacteria to Wheat Purothionins In Vitro. Appl. Microbiol. 1972, 23, 998–1000. [Google Scholar] [CrossRef]
  125. Kiss, G.; Michl, H. Uber Das Giftsekret Der Gelbbauchunke, Bombina variegata L. Toxicon 1962, 1, 33–34. [Google Scholar] [CrossRef]
  126. Simmaco, M.; Kreil, G.; Barra, D. Bombinins, Antimicrobial Peptides from Bombina Species. Biochim. Biophys. Acta-Biomembr. 2009, 1788, 1551–1555. [Google Scholar] [CrossRef]
  127. Hultmark, D.; Steiner, H.; Rasmuson, T.; Boman, H.G. Insect Immunity. Purification and Properties of Three Inducible Bactericidal Proteins from Hemolymph of Immunized Pupae of Hyalophora Cecropia. Eur. J. Biochem. 1980, 106, 7–16. [Google Scholar] [CrossRef] [PubMed]
  128. Steiner, H.; Hultmark, D.; Engström, Å.; Bennich, H.; Boman, H.G. Sequence and Specificity of Two Antibacterial Proteins Involved in Insect Immunity. Nature 1981, 292, 246–248. [Google Scholar] [CrossRef] [PubMed]
  129. Selsted, M.E.; Brown, D.M.; DeLange, R.J.; Lehrer, R.I. Primary Structures of MCP-1 and MCP-2, Natural Peptide Antibiotics of Rabbit Lung Macrophages. J. Biol. Chem. 1983, 258, 14485–14489. [Google Scholar] [CrossRef] [PubMed]
  130. Selsted, M.E.; Szklarek, D.; Lehrer, R.I. Purification and Antibacterial Activity of Antimicrobial Peptides of Rabbit Granulocytes. Infect. Immun. 1984, 45, 150–154. [Google Scholar] [CrossRef]
  131. Selsted, M.E.; Brown, D.M.; DeLange, R.J.; Harwig, S.S.; Lehrer, R.I. Primary Structures of Six Antimicrobial Peptides of Rabbit Peritoneal Neutrophils. J. Biol. Chem. 1985, 260, 4579–4584. [Google Scholar] [CrossRef]
  132. Ganz, T.; Selsted, M.E.; Szklarek, D.; Harwig, S.S.; Daher, K.; Bainton, D.F.; Lehrer, R.I. Defensins. Natural Peptide Antibiotics of Human Neutrophils. J. Clin. Investig. 1985, 76, 1427–1435. [Google Scholar] [CrossRef] [PubMed]
  133. Zasloff, M. Magainins, a Class of Antimicrobial Peptides from Xenopus Skin: Isolation, Characterization of Two Active Forms, and Partial CDNA Sequence of a Precursor. Proc. Natl. Acad. Sci. USA 1987, 84, 5449–5453. [Google Scholar] [CrossRef] [PubMed]
  134. Selsted, M.E.; Tang, Y.Q.; Morris, W.L.; McGuire, P.A.; Novotny, M.J.; Smith, W.; Henschen, A.H.; Cullor, J.S. Purification, Primary Structures, and Antibacterial Activities of Beta-Defensins, a New Family of Antimicrobial Peptides from Bovine Neutrophils. J. Biol. Chem. 1993, 268, 6641–6648. [Google Scholar] [CrossRef] [PubMed]
  135. Tang, Y.Q.; Yuan, J.; Ösapay, G.; Ösapay, K.; Tran, D.; Miller, C.J.; Ouellette, A.J.; Selsted, M.E. A Cyclic Antimicrobial Peptide Produced in Primate Leukocytes by the Ligation of Two Truncated α-Defensins. Science 1999, 286, 498–502. [Google Scholar] [CrossRef] [PubMed]
  136. Brogden, K.A.; Ackermann, M.; Huttner, K.M. Small, Anionic, and Charge-Neutralizing Propeptide Fragments of Zymogens Are Antimicrobial. Antimicrob. Agents Chemother. 1997, 41, 1615–1617. [Google Scholar] [CrossRef]
  137. Tam, J.P.; Wang, S.; Wong, K.H.; Tan, W.L. Antimicrobial Peptides from Plants. Pharmaceuticals 2015, 8, 711–757. [Google Scholar] [CrossRef]
  138. Masso-Silva, J.A.; Diamond, G. Antimicrobial Peptides from Fish. Pharmaceuticals 2014, 7, 265–310. [Google Scholar] [CrossRef] [PubMed]
  139. Pérez-Peinado, C.; Defaus, S.; Andreu, D. Hitchhiking with Nature: Snake Venom Peptides to Fight Cancer and Superbugs. Toxins 2020, 12, 255. [Google Scholar] [CrossRef]
  140. Santana, F.L.; Estrada, K.; Ortiz, E.; Corzo, G. Reptilian β-Defensins: Expanding the Repertoire of Known Crocodylian Peptides. Peptides 2021, 136, 170473. [Google Scholar] [CrossRef] [PubMed]
  141. Andreu, D.; Rivas, L. Animal Antimicrobial Peptides: An Overview. Pept. Sci. 1998, 47, 415–433. [Google Scholar] [CrossRef]
  142. Mangoni, M.L.; Shai, Y. Temporins and Their Synergism against Gram-Negative Bacteria and in Lipopolysaccharide Detoxification. Biochim. Biophys. Acta-Biomembr. 2009, 1788, 1610–1619. [Google Scholar] [CrossRef]
  143. Simons, A.; Alhanout, K.; Duval, R.E. Bacteriocins, Antimicrobial Peptides from Bacterial Origin: Overview of Their Biology and Their Impact against Multidrug-Resistant Bacteria. Microorganisms 2020, 8, 639. [Google Scholar] [CrossRef]
  144. Mookherjee, N.; Anderson, M.A.; Haagsman, H.P.; Davidson, D.J. Antimicrobial Host Defence Peptides: Functions and Clinical Potential. Nat. Rev. Drug Discov. 2020, 19, 311–332. [Google Scholar] [CrossRef] [PubMed]
  145. Harris, F.; Dennison, S.; Phoenix, D. Anionic Antimicrobial Peptides from Eukaryotic Organisms. Curr. Protein Pept. Sci. 2009, 10, 585–606. [Google Scholar] [CrossRef]
  146. Boparai, J.K.; Sharma, P.K. Mini Review on Antimicrobial Peptides, Sources, Mechanism and Recent Applications. Protein Pept. Lett. 2020, 27, 4–16. [Google Scholar] [CrossRef]
  147. Savini, F.; Bobone, S.; Roversi, D.; Mangoni, M.L.; Stella, L. From Liposomes to Cells: Filling the Gap between Physicochemical and Microbiological Studies of the Activity and Selectivity of Host-Defense Peptides. Pept. Sci. 2018, 110, e24041. [Google Scholar] [CrossRef]
  148. Pfalzgraff, A.; Brandenburg, K.; Weindl, G. Antimicrobial Peptides and Their Therapeutic Potential for Bacterial Skin Infections and Wounds. Front. Pharmacol. 2018, 9, 281. [Google Scholar] [CrossRef]
  149. Kumar, P.; Takayesu, A.; Abbasi, U.; Kalathottukaren, M.T.; Abbina, S.; Kizhakkedathu, J.N.; Straus, S.K. Antimicrobial Peptide-Polymer Conjugates with High Activity: Influence of Polymer Molecular Weight and Peptide Sequence on Antimicrobial Activity, Proteolysis, and Biocompatibility. ACS Appl. Mater. Interfaces 2017, 9, 37575–37586. [Google Scholar] [CrossRef] [PubMed]
  150. Kumar, P.; Pletzer, D.; Haney, E.F.; Rahanjam, N.; Cheng, J.T.J.; Yue, M.; Aljehani, W.; Hancock, R.E.W.; Kizhakkedathu, J.N.; Straus, S.K. Aurein-Derived Antimicrobial Peptides Formulated with Pegylated Phospholipid Micelles to Target Methicillin-Resistant Staphylococcus Aureus Skin Infections. ACS Infect. Dis. 2019, 5, 443–453. [Google Scholar] [CrossRef]
  151. Lynn, M.A.; Kindrachuk, J.; Marr, A.K.; Jenssen, H.; Panté, N.; Elliott, M.R.; Napper, S.; Hancock, R.E.; McMaster, W.R. Effect of BMAP-28 Antimicrobial Peptides on Leishmania Major Promastigote and Amastigote Growth: Role of Leishmanolysin in Parasite Survival. PLoS Negl. Trop. Dis. 2011, 5, e1141. [Google Scholar] [CrossRef]
  152. Kelly, B.L.; Nelson, T.N.; McMaster, W.R. Stage-Specific Expression in Leishmania Conferred by 3′ Untranslated Regions of L. Major Leishmanolysin Genes (GP63). Mol. Biochem. Parasitol. 2001, 116, 101–104. [Google Scholar] [CrossRef] [PubMed]
  153. Selsted, M.E.; Novotny, M.J.; Morris, W.L.; Tang, Y.Q.; Smith, W.; Cullor, J.S. Indolicidin, a Novel Bactericidal Tridecapeptide Amide from Neutrophils. J. Biol. Chem. 1992, 267, 4292–4295. [Google Scholar] [CrossRef]
  154. Falcao, C.B.; Pérez-Peinado, C.; De La Torre, B.G.; Mayol, X.; Zamora-Carreras, H.; Jiménez, M.Á.; Rádis-Baptista, G.; Andreu, D. Structural Dissection of Crotalicidin, a Rattlesnake Venom Cathelicidin, Retrieves a Fragment with Antimicrobial and Antitumor Activity. J. Med. Chem. 2015, 58, 8553–8563. [Google Scholar] [CrossRef]
  155. Pérez-Peinado, C.; Dias, S.A.; Mendonça, D.A.; Castanho, M.A.R.B.; Veiga, A.S.; Andreu, D. Structural Determinants Conferring Unusual Long Life in Human Serum to Rattlesnake-Derived Antimicrobial Peptide Ctn[15-34]. J. Pept. Sci. 2019, 25, e3195. [Google Scholar] [CrossRef]
  156. Pérez-Peinado, C.; Defaus, S.; Sans-Comerma, L.; Valle, J.; Andreu, D. Decoding the Human Serum Interactome of Snake-Derived Antimicrobial Peptide Ctn[15–34]: Toward an Explanation for Unusually Long Half-Life. J. Proteom. 2019, 204, 103372. [Google Scholar] [CrossRef] [PubMed]
  157. Neubauer, D.; Jaśkiewicz, M.; Migoń, D.; Bauer, M.; Sikora, K.; Sikorska, E.; Kamysz, E.; Kamysz, W. Retro Analog Concept: Comparative Study on Physico-Chemical and Biological Properties of Selected Antimicrobial Peptides. Amino Acids 2017, 49, 1755–1771. [Google Scholar] [CrossRef]
  158. Chen, Y.; Vasil, A.I.; Rehaume, L.; Mant, C.T.; Burns, J.L.; Vasil, M.L.; Hancock, R.E.W.; Hodges, R.S. Comparison of Biophysical and Biologic Properties of α-Helical Enantiomeric Antimicrobial Peptides. Chem. Biol. Drug Des. 2006, 67, 162–173. [Google Scholar] [CrossRef] [PubMed]
  159. Won, A.; Khan, M.; Gustin, S.; Akpawu, A.; Seebun, D.; Avis, T.J.; Leung, B.O.; Hitchcock, A.P.; Ianoul, A. Investigating the Effects of L- to D-Amino Acid Substitution and Deamidation on the Activity and Membrane Interactions of Antimicrobial Peptide Anoplin. Biochim. Biophys. Acta-Biomembr. 2011, 1808, 1592–1600. [Google Scholar] [CrossRef]
  160. Hirt, H.; Hall, J.W.; Larson, E.; Gorr, S.U. A D-Enantiomer of the Antimicrobial Peptide GL13K Evades Antimicrobial Resistance in the Gram Positive Bacteria Enterococcus Faecalis and Streptococcus Gordonii. PLoS ONE 2018, 13, e0194900. [Google Scholar] [CrossRef] [PubMed]
  161. Hirt, H.; Gorr, S.U. Antimicrobial Peptide GL13K Is Effective in Reducing Biofilms of Pseudomonas Aeruginosa. Antimicrob. Agents Chemother. 2013, 57, 4903–4910. [Google Scholar] [CrossRef] [PubMed]
  162. Cleophas, R.T.C.; Riool, M.; Quarles Van Ufford, H.C.; Zaat, S.A.J.; Kruijtzer, J.A.W.; Liskamp, R.M.J. Convenient Preparation of Bactericidal Hydrogels by Covalent Attachment of Stabilized Antimicrobial Peptides Using Thiol-Ene Click Chemistry. ACS Macro Lett. 2014, 3, 477–480. [Google Scholar] [CrossRef]
  163. Buckholtz, G.A.; Reger, N.A.; Anderton, W.D.; Schimoler, P.J.; Roudebush, S.L.; Meng, W.S.; Miller, M.C.; Gawalt, E.S. Reducing Escherichia Coli Growth on a Composite Biomaterial by a Surface Immobilized Antimicrobial Peptide. Mater. Sci. Eng. C 2016, 65, 126–134. [Google Scholar] [CrossRef]
  164. Cleophas, R.T.C.; Sjollema, J.; Busscher, H.J.; Kruijtzer, J.A.W.; Liskamp, R.M.J. Characterization and Activity of an Immobilized Antimicrobial Peptide Containing Bactericidal PEG-Hydrogel. Biomacromolecules 2014, 15, 3390–3395. [Google Scholar] [CrossRef]
  165. Ong, Z.Y.; Cheng, J.; Huang, Y.; Xu, K.; Ji, Z.; Fan, W.; Yang, Y.Y. Effect of Stereochemistry, Chain Length and Sequence Pattern on Antimicrobial Properties of Short Synthetic β-Sheet Forming Peptide Amphiphiles. Biomaterials 2014, 35, 1315–1325. [Google Scholar] [CrossRef]
  166. 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]
  167. Czechowicz, P.; Jaśkiewicz, M.; Neubauer, D.; Gościniak, G.; Kamysz, W. Anticandidal Activity of Omiganan and Its Retro Analog Alone and in Combination with Fluconazole. Probiotics Antimicrob. Proteins 2021, 13, 1173–1182. [Google Scholar] [CrossRef]
  168. Mohamed, M.F.; Brezden, A.; Mohammad, H.; Chmielewski, J.; Seleem, M.N. A Short D-Enantiomeric Antimicrobial Peptide with Potent Immunomodulatory and Antibiofilm Activity against Multidrug-Resistant Pseudomonas Aeruginosa and Acinetobacter Baumannii. Sci. Rep. 2017, 7, 6953. [Google Scholar] [CrossRef]
  169. Greco, I.; Hansen, J.E.; Jana, B.; Molchanova, N.; Oddo, A.; Thulstrup, P.W.; Damborg, P.; Guardabassi, L.; Hansen, P.R. Structure–Activity Study, Characterization, and Mechanism of Action of an Antimicrobial Peptoid D2 and Its D- and L-Peptide Analogues. Molecules 2019, 24, 1121. [Google Scholar] [CrossRef]
  170. Cesaro, A.; Torres, M.D.T.; Gaglione, R.; Dell’Olmo, E.; Di Girolamo, R.; Bosso, A.; Pizzo, E.; Haagsman, H.P.; Veldhuizen, E.J.A.; de la Fuente-Nunez, C.; et al. Synthetic Antibiotic Derived from Sequences Encrypted in a Protein from Human Plasma. ACS Nano 2022, 16, 1880–1895. [Google Scholar] [CrossRef] [PubMed]
  171. López-Sanmartín, M.; Rengel, R.; López-López, M.; Lebrón, J.A.; Molina-Márquez, A.; de la Rosa, I.; López-Cornejo, P.; Cuesta, A.; Vigara, J.; León, R. D-Amino Acid Peptides as Antimicrobial Agents against Vibrio-Associated Diseases in Aquaculture. Aquaculture 2023, 569, 739362. [Google Scholar]
  172. Schenk, M.; Mueller, C. The Mucosal Immune System at the Gastrointestinal Barrier. Best Pract. Res. Clin. Gastroenterol. 2008, 22, 391–409. [Google Scholar] [CrossRef]
  173. Tetro, N.; Moushaev, S.; Rubinchik-Stern, M.; Eyal, S. The Placental Barrier: The Gate and the Fate in Drug Distribution. Pharm. Res. 2018, 35, 71. [Google Scholar]
  174. Pardridge, W.M. The Blood-Brain Barrier: Bottleneck in Brain Drug Development. NeuroRx 2005, 2, 3–14. [Google Scholar] [CrossRef]
  175. Futaki, S.; Suzuki, T.; Ohashi, W.; Yagami, T.; Tanaka, S.; Ueda, K.; Sugiura, Y. Arginine-Rich Peptides: An Abundant Source of Membrane-Permeable Peptides Having Potential as Carriers for Intracellular Protein Delivery. J. Biol. Chem. 2001, 276, 5836–5840. [Google Scholar] [CrossRef]
  176. Liu, Y.; Mei, L.; Xu, C.; Yu, Q.; Shi, K.; Zhang, L.; Wang, Y.; Zhang, Q.; Gao, H.; Zhang, Z.; et al. Dual Receptor Recognizing Cell Penetrating Peptide for Selective Targeting, Efficient Intratumoral Diffusion and Synthesized Anti-Glioma Therapy. Theranostics 2016, 6, 177–191. [Google Scholar] [CrossRef] [PubMed]
  177. Futaki, S.; Ohashi, W.; Suzuki, T.; Niwa, M.; Tanaka, S.; Ueda, K.; Harashima, H.; Sugiura, Y. Stearylated Arginine-Rich Peptides: A New Class of Transfection Systems. Bioconjug. Chem. 2001, 12, 1005–1011. [Google Scholar] [CrossRef] [PubMed]
  178. Iwasaki, T. Cationic Peptide. In Cell-Penetrating Peptides: Design, Development and Applications; Oba, M., Demizu, Y., Eds.; John Wiley & Sons, Ltd.: London, UK, 2023; pp. 45–55. [Google Scholar]
  179. Pujals, S.; Sabidó, E.; Tarragó, T.; Giralt, E. All-D Proline-Rich Cell-Penetrating Peptides: A Preliminary In Vivo Internalization Study. Biochem. Soc. Trans. 2007, 35, 794–796. [Google Scholar] [PubMed]
  180. Fernández-Carneado, J.; Kogan, M.J.; Pujals, S.; Giralt, E. Amphipathic Peptides and Drug Delivery. Pept. Sci. 2004, 76, 196–203. [Google Scholar]
  181. Oba, M. Amphipathic Peptide. In Cell-Penetrating Peptides: Design, Development and Applications; Oba, M., Demizu, Y., Eds.; John Wiley & Sons, Ltd.: London, UK, 2023; pp. 57–67. [Google Scholar]
  182. Demizu, Y. Hydrophobic CPPs. In Cell-Penetrating Peptides: Design, Development and Applications; Oba, M., Demizu, Y., Eds.; John Wiley & Sons, Ltd.: London, UK, 2023; pp. 69–77. [Google Scholar]
  183. Schmidt, S.; Adjobo-Hermans, M.J.W.; Kohze, R.; Enderle, T.; Brock, R.; Milletti, F. Identification of Short Hydrophobic Cell-Penetrating Peptides for Cytosolic Peptide Delivery by Rational Design. Bioconjug. Chem. 2017, 28, 382–389. [Google Scholar] [PubMed]
  184. Jirka, S.M.G.; Heemskerk, H.; Tanganyika-De Winter, C.L.; Muilwijk, D.; Pang, K.H.; De Visser, P.C.; Janson, A.; Karnaoukh, T.G.; Vermue, R.; ’T Hoen, P.A.C.; et al. Peptide Conjugation of 2′-O-Methyl Phosphorothioate Antisense Oligonucleotides Enhances Cardiac Uptake and Exon Skipping in Mdx Mice. Nucleic Acid Ther. 2014, 24, 25–36. [Google Scholar]
  185. Martín, I.; Teixidó, M.; Giralt, E. Design, Synthesis and Characterization of a New Anionic Cell-Penetrating Peptide: SAP(E). ChemBioChem 2011, 12, 896–903. [Google Scholar]
  186. Kam, A.; Loo, S.; Fan, J.S.; Sze, S.K.; Yang, D.; Tam, J.P. Roseltide RT7 Is a Disulfide-Rich, Anionic, and Cell-Penetrating Peptide That Inhibits Proteasomal Degradation. J. Biol. Chem. 2019, 294, 19604–19615. [Google Scholar]
  187. Szabó, I.; Yousef, M.; Soltész, D.; Bató, C.; Mező, G.; Bánóczi, Z. Redesigning of Cell-Penetrating Peptides to Improve Their Efficacy as a Drug Delivery System. Pharmaceutics 2022, 14, 907. [Google Scholar]
  188. Palm, C.; Jayamanne, M.; Kjellander, M.; Hällbrink, M. Peptide Degradation Is a Critical Determinant for Cell-Penetrating Peptide Uptake. Biochim. Biophys. Acta-Biomembr. 2007, 1768, 1769–1776. [Google Scholar]
  189. Kalafatovic, D.; Giralt, E. Cell-Penetrating Peptides: Design Strategies beyond Primary Structure and Amphipathicity. Molecules 2017, 22, 1929. [Google Scholar]
  190. Vivès, E.; Brodin, P.; Lebleu, B. A Truncated HIV-1 Tat Protein Basic Domain Rapidly Translocates through the Plasma Membrane and Accumulates in the Cell Nucleus. J. Biol. Chem. 1997, 272, 16010–16017. [Google Scholar]
  191. Wender, P.A.; Mitchell, D.J.; Pattabiraman, K.; Pelkey, E.T.; Steinman, L.; Rothbard, J.B. The Design, Synthesis, and Evaluation of Molecules That Enable or Enhance Cellular Uptake: Peptoid Molecular Transporters. Proc. Natl. Acad. Sci. USA 2000, 97, 13003–13008. [Google Scholar] [PubMed]
  192. Seisel, Q.; Lakumpa, I.; Josse, E.; Vivès, E.; Varilh, J.; Taulan-Cadars, M.; Boisguérin, P. Highway to Cell: Selection of the Best Cell-Penetrating Peptide to Internalize the CFTR-Stabilizing ICAL36 Peptide. Pharmaceutics 2022, 14, 808. [Google Scholar] [PubMed]
  193. Nielsen, E.J.B.; Yoshida, S.; Kamei, N.; Iwamae, R.; Khafagy, E.S.; Olsen, J.; Rahbek, U.L.; Pedersen, B.L.; Takayama, K.; Takeda-Morishita, M. In Vivo Proof of Concept of Oral Insulin Delivery Based on a Co-Administration Strategy with the Cell-Penetrating Peptide Penetratin. J. Control. Release 2014, 189, 19–24. [Google Scholar] [PubMed]
  194. Kamei, N.; Shigei, C.; Hasegawa, R.; Takeda-Morishita, M. Exploration of the Key Factors for Optimizing the In Vivo Oral Delivery of Insulin by Using a Noncovalent Strategy with Cell-Penetrating Peptides. Biol. Pharm. Bull. 2018, 41, 239–246. [Google Scholar]
  195. Garcia, J.; Fernández-Blanco, Á.; Teixidó, M.; Sánchez-Navarro, M.; Giralt, E. D-Polyarginine Lipopeptides as Intestinal Permeation Enhancers. ChemMedChem 2018, 13, 2045–2052. [Google Scholar] [PubMed]
  196. Kamei, N.; Yamaoka, A.; Fukuyama, Y.; Itokazu, R.; Takeda-Morishita, M. Noncovalent Strategy with Cell-Penetrating Peptides to Facilitate the Brain Delivery of Insulin through the Blood–Brain Barrier. Biol. Pharm. Bull. 2018, 41, 546–554. [Google Scholar] [PubMed]
  197. Wang, C.; Lin, G.; Shen, Z.; Wang, R. Angiopep-2 as an Exogenous Chemical Exchange Saturation Transfer Contrast Agent in Diagnosis of Alzheimer’s Disease. J. Healthc. Eng. 2022, 2022, 7480519. [Google Scholar] [PubMed]
  198. Xie, R.; Wu, Z.; Zeng, F.; Cai, H.; Wang, D.; Gu, L.; Zhu, H.; Lui, S.; Guo, G.; Song, B.; et al. Retro-Enantio Isomer of Angiopep-2 Assists Nanoprobes across the Blood-Brain Barrier for Targeted Magnetic Resonance/Fluorescence Imaging of Glioblastoma. Signal Transduct. Target. Ther. 2021, 6, 309. [Google Scholar]
  199. Díaz-Perlas, C.; Oller-Salvia, B.; Sánchez-Navarro, M.; Teixidó, M.; Giralt, E. Branched BBB-Shuttle Peptides: Chemoselective Modification of Proteins to Enhance Blood–Brain Barrier Transport. Chem. Sci. 2018, 9, 8409–8415. [Google Scholar]
  200. Bukchin, A.; Sanchez-Navarro, M.; Carrera, A.; Teixidó, M.; Carcaboso, A.M.; Giralt, E.; Sosnik, A. Amphiphilic Polymeric Nanoparticles Modified with a Retro-Enantio Peptide Shuttle Target the Brain of Mice. Chem. Mater. 2020, 32, 7679–7693. [Google Scholar]
  201. Jones, S.; Howl, J. Enantiomer-Specific Bioactivities of Peptidomimetic Analogues of Mastoparan and Mitoparan: Characterization of Inverso Mastoparan as a Highly Efficient Cell Penetrating Peptide. Bioconjug. Chem. 2012, 23, 47–56. [Google Scholar] [PubMed]
  202. Verdurmen, W.P.R.; Bovee-Geurts, P.H.; Wadhwani, P.; Ulrich, A.S.; Hällbrink, M.; Van Kuppevelt, T.H.; Brock, R. Preferential Uptake of L- versus D-Amino Acid Cell-Penetrating Peptides in a Cell Type-Dependent Manner. Chem. Biol. 2011, 18, 1000–1010. [Google Scholar]
  203. Li, J.P.; Kusche-Gullberg, M. Heparan Sulfate: Biosynthesis, Structure, and Function. Int. Rev. Cell Mol. Biol. 2016, 325, 215–273. [Google Scholar] [PubMed]
  204. Nakase, I.; Niwa, M.; Takeuchi, T.; Sonomura, K.; Kawabata, N.; Koike, Y.; Takehashi, M.; Tanaka, S.; Ueda, K.; Simpson, J.C.; et al. Cellular Uptake of Arginine-Rich Peptides: Roles for Macropinocytosis and Actin Rearrangement. Mol. Ther. 2004, 10, 1011–1022. [Google Scholar]
  205. Zhan, C.; Li, B.; Hu, L.; Wei, X.; Feng, L.; Fu, W.; Lu, W. Micelle-Based Brain-Targeted Drug Delivery Enabled by a Nicotine Acetylcholine Receptor Ligand. Angew. Chem. Int. Ed. 2011, 50, 5482–5485. [Google Scholar]
  206. Wei, X.; Zhan, C.; Shen, Q.; Fu, W.; Xie, C.; Gao, J.; Peng, C.; Zheng, P.; Lu, W. A D-Peptide Ligand of Nicotine Acetylcholine Receptors for Brain-Targeted Drug Delivery. Angew. Chem. Int. Ed. 2015, 54, 3023–3027. [Google Scholar]
  207. Han, B.; Xie, W.; Zhang, Y.; Zhou, S.; Yang, J.; Wang, R.; Sun, Y.; Wang, X.; Xu, J.; Chen, D.; et al. The Influx/Efflux Mechanisms of d-Peptide Ligand of NAChRs across the Blood–Brain Barrier and Its Therapeutic Value in Treating Glioma. J. Control. Release 2020, 327, 384–396. [Google Scholar] [PubMed]
  208. Gallo, M.; Moreno, E.; Defaus, S.; Ortega-Alvaro, A.; Gonzalez, A.; Robledo, P.; Cavaco, M.; Neves, V.; Castanho, M.A.R.B.; Casadó, V.; et al. Orally Active Peptide Vector Allows Using Cannabis to Fight Pain While Avoiding Side Effects. J. Med. Chem. 2021, 64, 6937–6948. [Google Scholar] [PubMed]
  209. Meloni, B.P.; Craig, A.J.; Milech, N.; Hopkins, R.M.; Watt, P.M.; Knuckey, N.W. The Neuroprotective Efficacy of Cell-Penetrating Peptides TAT, Penetratin, Arg-9, and Pep-1 in Glutamic Acid, Kainic Acid, and In Vitro Ischemia Injury Models Using Primary Cortical Neuronal Cultures. Cell. Mol. Neurobiol. 2014, 34, 173–181. [Google Scholar]
  210. Snyder, E.L.; Meade, B.R.; Saenz, C.C.; Dowdy, S.F. Treatment of Terminal Peritoneal Carcinomatosis by a Transducible P53-Activating Peptide. PLoS Biol. 2004, 2, 0186–0193. [Google Scholar]
  211. Liu, L.; Xie, H.J.; Mu, L.M.; Liu, R.; Su, Z.B.; Cui, Y.N.; Xie, Y.; Lu, W.L. Functional Chlorin Gold Nanorods Enable to Treat Breast Cancer by Photothermal/Photodynamic Therapy. Int. J. Nanomed. 2018, 13, 8119–8135. [Google Scholar]
  212. Najjar, K.; Erazo-Oliveras, A.; Brock, D.J.; Wang, T.Y.; Pellois, J.P. An L- to D-Amino Acid Conversion in an Endosomolytic Analog of the Cell-Penetrating Peptide TAT Influences Proteolytic Stability, Endocytic Uptake, and Endosomal Escape. J. Biol. Chem. 2017, 292, 847–861. [Google Scholar] [PubMed]
  213. Birch, D.; Christensen, M.V.; Staerk, D.; Franzyk, H.; Nielsen, H.M. Stereochemistry as a Determining Factor for the Effect of a Cell-Penetrating Peptide on Cellular Viability and Epithelial Integrity. Biochem. J. 2018, 475, 1773–1788. [Google Scholar]
  214. Liu, X.; Liu, C.; Zhang, W.; Xie, C.; Wei, G.; Lu, W. Oligoarginine-Modified Biodegradable Nanoparticles Improve the Intestinal Absorption of Insulin. Int. J. Pharm. 2013, 448, 159–167. [Google Scholar] [PubMed]
  215. Ndeboko, B.; Ramamurthy, N.; Lemamy, G.J.; Jamard, C.; Nielsen, P.E.; Cova, L. Role of Cell-Penetrating Peptides in Intracellular Delivery of Peptide Nucleic Acids Targeting Hepadnaviral Replication. Mol. Ther.-Nucleic Acids 2017, 9, 162–169. [Google Scholar]
  216. Edwards, A.B.; Cross, J.L.; Anderton, R.S.; Knuckey, N.W.; Meloni, B.P. Poly-Arginine R18 and R18D (D-Enantiomer) Peptides Reduce Infarct Volume and Improves Behavioural Outcomes Following Perinatal Hypoxic-Ischaemic Encephalopathy in the P7 Rat. Mol. Brain 2018, 11, 8. [Google Scholar]
  217. Wei, X.; Zhan, C.; Chen, X.; Hou, J.; Xie, C.; Lu, W. Retro-Inverso Isomer of Angiopep-2: A Stable d -Peptide Ligand Inspires Brain-Targeted Drug Delivery. Mol. Pharm. 2014, 11, 3261–3268. [Google Scholar]
  218. Bukchin, A.; Sanchez-Navarro, M.; Carrera, A.; Resa-Pares, C.; Castillo-Ecija, H.; Balaguer-Lluna, L.; Teixidó, M.; Olaciregui, N.G.; Giralt, E.; Carcaboso, A.M.; et al. Amphiphilic Polymeric Nanoparticles Modified with a Protease-Resistant Peptide Shuttle for the Delivery of SN-38 in Diffuse Intrinsic Pontine Glioma. ACS Appl. Nano Mater. 2021, 4, 1314–1329. [Google Scholar]
  219. Bonny, C.; Oberson, A.; Negri, S.; Sauser, C.; Schorderet, D.F. Cell-Permeable Peptide Inhibitors of JNKNovel Blockers of β-Cell Death. Diabetes 2001, 50, 77–82. [Google Scholar] [PubMed]
  220. Youngblood, D.S.; Hatlevig, S.A.; Hassinger, J.N.; Iversen, P.L.; Moulton, H.M. Stability of Cell-Penetrating Peptide-Morpholino Oligomer Conjugates in Human Serum and in Cells. Bioconjug. Chem. 2007, 18, 50–60. [Google Scholar]
  221. Pujals, S.; Fernández-Carneado, J.; Ludevid, M.D.; Giralt, E. D-SAP: A New, Noncytotoxic, and Fully Protease Resistant Cell-Penetrating Peptide. ChemMedChem 2008, 3, 296–301. [Google Scholar]
  222. Shaltiel-Karyo, R.; Frenkel-Pinter, M.; Egoz-Matia, N.; Frydman-Marom, A.; Shalev, D.E.; Segal, D.; Gazit, E. Inhibiting α-Synuclein Oligomerization by Stable Cell-Penetrating β-Synuclein Fragments Recovers Phenotype of Parkinson’s Disease Model Flies. PLoS ONE 2010, 5, e13863. [Google Scholar]
  223. Foy, K.C.; Liu, Z.; Phillips, G.; Miller, M.; Kaumaya, P.T.P. Combination Treatment with HER-2 and VEGF Peptide Mimics Induces Potent Anti-Tumor and Anti-Angiogenic Responses In Vitro and In Vivo. J. Biol. Chem. 2011, 286, 13626–13637. [Google Scholar]
  224. Québatte, G.; Kitas, E.; Seelig, J. RiDOM, a Cell Penetrating Peptide. Interaction with Phospholipid Bilayers. Biochim. Biophys. Acta-Biomembr. 2014, 1838, 968–977. [Google Scholar]
  225. Radicioni, G.; Stringaro, A.; Molinari, A.; Nocca, G.; Longhi, R.; Pirolli, D.; Scarano, E.; Iavarone, F.; Manconi, B.; Cabras, T.; et al. Characterization of the Cell Penetrating Properties of a Human Salivary Proline-Rich Peptide. Biochim. Biophys. Acta-Biomembr. 2015, 1848, 2868–2877. [Google Scholar]
  226. Rodrigues, M.; Andreu, D.; Santos, N.C. Uptake and Cellular Distribution of Nucleolar Targeting Peptides (NrTPs) in Different Cell Types. Pept. Sci. 2015, 104, 101–109. [Google Scholar]
  227. Vaissière, A.; Aldrian, G.; Konate, K.; Lindberg, M.F.; Jourdan, C.; Telmar, A.; Seisel, Q.; Fernandez, F.; Viguier, V.; Genevois, C.; et al. A Retro-Inverso Cell-Penetrating Peptide for SiRNA Delivery. J. Nanobiotechnol. 2017, 15, 34. [Google Scholar]
  228. Arranz-Gibert, P.; Ciudad, S.; Seco, J.; García, J.; Giralt, E.; Teixidó, M. Immunosilencing Peptides by Stereochemical Inversion and Sequence Reversal: Retro-D-Peptides. Sci. Rep. 2018, 8, 6446. [Google Scholar]
  229. Ren, Y.; Zhan, C.; Gao, J.; Zhang, M.; Wei, X.; Ying, M.; Liu, Z.; Lu, W. A D-Peptide Ligand of Integrins for Simultaneously Targeting Angiogenic Blood Vasculature and Glioma Cells. Mol. Pharm. 2018, 15, 592–601. [Google Scholar] [PubMed]
  230. Chai, Z.; Wu, T.; Dai, A.; Huynh, P.; Koentgen, F.; Krippner, G.; Ren, S.; Cooper, M.E. Targeting the CDA1/CDA1BP1 Axis Retards Renal Fibrosis in Experimental Diabetic Nephropathy. Diabetes 2019, 68, 395–408. [Google Scholar] [PubMed]
  231. Jiao, L.; Dong, Q.; Zhai, W.; Zhao, W.; Shi, P.; Wu, Y.; Zhou, X.; Gao, Y. A PD-L1 and VEGFR2 Dual Targeted Peptide and Its Combination with Irradiation for Cancer Immunotherapy. Pharmacol. Res. 2022, 182, 106343. [Google Scholar] [CrossRef]
  232. Knezevic, C.E.; Clarke, W. Cancer Chemotherapy: The Case for Therapeutic Drug Monitoring. Ther. Drug Monit. 2020, 42, 6–19. [Google Scholar]
  233. Springfeld, C.; Jäger, D.; Büchler, M.W.; Strobel, O.; Hackert, T.; Palmer, D.H.; Neoptolemos, J.P. Chemotherapy for Pancreatic Cancer. Presse Med. 2019, 48, e159–e174. [Google Scholar]
  234. Bukowski, K.; Kciuk, M.; Kontek, R. Mechanisms of Multidrug Resistance in Cancer Chemotherapy. Int. J. Mol. Sci. 2020, 21, 3233. [Google Scholar] [CrossRef]
  235. Arina, A.; Gutiontov, S.I.; Weichselbaum, R.R. Radiotherapy and Immunotherapy for Cancer: From “Systemic” to “Multisite". Clin. Cancer Res. 2020, 26, 2777–2782. [Google Scholar] [CrossRef]
  236. Vinod, S.K.; Hau, E. Radiotherapy Treatment for Lung Cancer: Current Status and Future Directions. Respirology 2020, 25, 61–71. [Google Scholar] [CrossRef] [PubMed]
  237. Podder, T.K.; Fredman, E.T.; Ellis, R.J. Advances in Radiotherapy for Prostate Cancer Treatment. In Mechanisms of Multidrug Resistance in Cancer Chemotherapy; Schatten, H., Ed.; Springer: New York, NY, USA, 2018; Volume 1126, pp. 31–47. [Google Scholar]
  238. Dijksteel, G.S.; Ulrich, M.M.W.; Middelkoop, E.; Boekema, B.K.H.L. Review: Lessons Learned from Clinical Trials Using Antimicrobial Peptides (AMPs). Front. Microbiol. 2021, 12, 616979. [Google Scholar]
  239. Riley, R.S.; June, C.H.; Langer, R.; Mitchell, M.J. Delivery Technologies for Cancer Immunotherapy. Nat. Rev. Drug Discov. 2019, 18, 175–196. [Google Scholar]
  240. Hoda, M. Potential Alternatives to Conventional Cancer Therapeutic Approaches: The Way Forward. Curr. Pharm. Biotechnol. 2021, 22, 1141–1148. [Google Scholar] [PubMed]
  241. Chiangjong, W.; Chutipongtanate, S.; Hongeng, S. Anticancer Peptide: Physicochemical Property, Functional Aspect and Trend in Clinical Application. Int. J. Oncol. 2020, 57, 678–696. [Google Scholar] [CrossRef] [PubMed]
  242. Johansson, A.; Dar, H.; van ’t Veer, L.J.; Tobin, N.P.; Perez-Tenorio, G.; Nordenskjöld, A.; Johansson, U.; Hartman, J.; Skoog, L.; Yau, C.; et al. Twenty-Year Benefit from Adjuvant Goserelin and Tamoxifen in Premenopausal Patients with Breast Cancer in a Controlled Randomized Clinical Trial. J. Clin. Oncol. 2022, 40, 4071–4082. [Google Scholar] [CrossRef] [PubMed]
  243. Wilson, A.C.; Meethal, S.V.; Bowen, R.L.; Atwood, C.S. Leuprolide Acetate: A Drug of Diverse Clinical Applications. Expert Opin. Investig. Drugs 2007, 16, 1851–1863. [Google Scholar] [CrossRef]
  244. Lamberts, S.W.J.; Hofland, L.J. Octreotide, 40 Years Later. Eur. J. Endocrinol. 2019, 181, R173–R183. [Google Scholar] [CrossRef]
  245. de la Torre, B.G.; Albericio, F. Peptide Therapeutics 2.0. Molecules 2020, 25, 2293. [Google Scholar] [CrossRef]
  246. D’Hondt, M.; Bracke, N.; Taevernier, L.; Gevaert, B.; Verbeke, F.; Wynendaele, E.; De Spiegeleer, B. Related Impurities in Peptide Medicines. J. Pharm. Biomed. Anal. 2014, 101, 2–30. [Google Scholar] [CrossRef]
  247. Hilchie, A.L.; Hoskin, D.W.; Power Coombs, M.R. Anticancer Activities of Natural and Synthetic Peptides. Adv. Exp. Med. Biol. 2019, 1117, 131–147. [Google Scholar]
  248. Hwang, J.S.; Kim, S.G.; Shin, T.H.; Jang, Y.E.; Kwon, D.H.; Lee, G. Development of Anticancer Peptides Using Artificial Intelligence and Combinational Therapy for Cancer Therapeutics. Pharmaceutics 2022, 14, 997. [Google Scholar] [CrossRef]
  249. Matijass, M.; Neundorf, I. Cell-Penetrating Peptides as Part of Therapeutics Used in Cancer Research. Med. Drug Discov. 2021, 10, 100092. [Google Scholar] [CrossRef]
  250. Norouzi, P.; Mirmohammadi, M.; Houshdar Tehrani, M.H. Anticancer Peptides Mechanisms, Simple and Complex. Chem. Biol. Interact. 2022, 368, 110194. [Google Scholar] [CrossRef]
  251. Bidwell, G.L.; Raucher, D. Therapeutic Peptides for Cancer Therapy. Part I—Peptide Inhibitors of Signal Transduction Cascades. Expert Opin. Drug Deliv. 2009, 6, 1033–1047. [Google Scholar] [CrossRef]
  252. Raucher, D.; Moktan, S.; Massodi, I.; Bidwell, G.L. Therapeutic Peptides for Cancer Therapy. Part II—Cell Cycle Inhibitory Peptides and Apoptosis-Inducing Peptides. Expert Opin. Drug Deliv. 2009, 6, 1049–1064. [Google Scholar] [CrossRef]
  253. Dutoit, V.; Migliorini, D.; Ranzanici, G.; Marinari, E.; Widmer, V.; Lobrinus, J.A.; Momjian, S.; Costello, J.; Walker, P.R.; Okada, H.; et al. Antigenic Expression and Spontaneous Immune Responses Support the Use of a Selected Peptide Set from the IMA950 Glioblastoma Vaccine for Immunotherapy of Grade II and III Glioma. Oncoimmunology 2018, 7, e1391972. [Google Scholar] [CrossRef] [PubMed]
  254. Thundimadathil, J. Cancer Treatment Using Peptides: Current Therapies and Future Prospects. J. Amino Acids 2012, 2012, 1–13. [Google Scholar] [CrossRef]
  255. Miller, W.R.; Scott, W.N.; Morris, R.; Fraser, H.M.; Sharpe, R.M. Growth of Human Breast Cancer Cells Inhibited by a Luteinizing Hormone-Releasing Hormone Agonist. Nature 1985, 313, 231–233. [Google Scholar] [CrossRef] [PubMed]
  256. Schally, A.V.; Zhang, X.; Cai, R.; Hare, J.M.; Granata, R.; Bartoli, M. Actions and Potential Therapeutic Applications of Growth Hormone–Releasing Hormone Agonists. Endocrinology 2019, 160, 1600–1612. [Google Scholar] [CrossRef]
  257. Tornesello, A.L.; Borrelli, A.; Buonaguro, L.; Buonaguro, F.M.; Tornesello, M.L. Antimicrobial Peptides as Anticancer Agents: Functional Properties and Biological Activities. Molecules 2020, 25, 2850. [Google Scholar] [CrossRef] [PubMed]
  258. Pérez-Peinado, C.; Valle, J.; Freire, J.M.; Andreu, D. Tumor Cell Attack by Crotalicidin (Ctn) and Its Fragment Ctn[15-34]: Insights into Their Dual Membranolytic and Intracellular Targeting Mechanism. ACS Chem. Biol. 2020, 15, 2945–2957. [Google Scholar] [CrossRef] [PubMed]
  259. Pazgier, M.; Liu, M.; Zou, G.; Yuan, W.; Li, C.; Li, C.; Li, J.; Monbo, J.; Zella, D.; Tarasov, S.G.; et al. Structural Basis for High-Affinity Peptide Inhibition of P53 Interactions with MDM2 and MDMX. Proc. Natl. Acad. Sci. USA 2009, 106, 4665–4670. [Google Scholar] [CrossRef]
  260. Liu, M.; Pazgier, M.; Li, C.; Yuan, W.; Li, C.; Lu, W. A Left-Handed Solution to Peptide Inhibition of the P53–MDM2 Interaction. Angew. Chem. Int. Ed. 2010, 49, 3649–3652. [Google Scholar] [CrossRef]
  261. Giordano, R.J.; Cardó-Vila, M.; Lahdenranta, J.; Pasqualini, R.; Arap, W. Biopanning and Rapid Analysis of Selective Interactive Ligands. Nat. Med. 2001, 7, 1249–1253. [Google Scholar] [CrossRef]
  262. Giordano, R.J.; Cardó-Vila, M.; Salameh, A.; Anobom, C.D.; Zeitlin, B.D.; Hawke, D.H.; Valente, A.P.; Almeida, F.C.L.; Nör, J.E.; Sidman, R.L.; et al. From Combinatorial Peptide Selection to Drug Prototype (I): Targeting the Vascular Endothelial Growth Factor Receptor Pathway. Proc. Natl. Acad. Sci. USA 2010, 107, 5112–5117. [Google Scholar] [CrossRef]
  263. Rezazadeh, F.; Sadeghzadeh, N.; Abedi, S.M.; Abediankenari, S. 99mTc Labeled D(LPR): A Novel Retro-Inverso Peptide for VEGF Receptor-1 Targeted Tumor Imaging. Nucl. Med. Biol. 2018, 62–63, 54–62. [Google Scholar] [CrossRef] [PubMed]
  264. Liu, Y.; Mei, L.; Yu, Q.; Zhang, Q.; Gao, H.; Zhang, Z.; He, Q. Integrin Avβ3 Targeting Activity Study of Different Retro-Inverso Sequences of RGD and Their Potentiality in the Designing of Tumor Targeting Peptides. Amino Acids 2015, 47, 2533–2539. [Google Scholar] [CrossRef]
  265. Ramezanizadeh, M.; Masterifarahani, A.; Sadeghzadeh, N.; Abediankenari, S.; Mardanshahi, A.; Maleki, F. 99mTc-D(RGD): Molecular Imaging Probe for Diagnosis of α β -Positive Tumors. Nucl. Med. Commun. 2020, 41, 104–109. [Google Scholar] [CrossRef] [PubMed]
  266. Mandelin, J.; Cardó-Vila, M.; Driessen, W.H.P.; Mathew, P.; Navonea, N.M.; Lin, S.H.; Logothetis, C.J.; Rietz, A.C.; Dobroff, A.S.; Proneth, B.; et al. Selection and Identification of Ligand Peptides Targeting a Model of Castrate-Resistant Osteogenic Prostate Cancer and Their Receptors. Proc. Natl. Acad. Sci. USA 2015, 112, 3776–3781. [Google Scholar] [CrossRef] [PubMed]
  267. Ran, D.; Mao, J.; Shen, Q.; Xie, C.; Zhan, C.; Wang, R.; Lu, W. GRP78 Enabled Micelle-Based Glioma Targeted Drug Delivery. J. Control. Release 2017, 255, 120–131. [Google Scholar] [CrossRef]
  268. Teesalu, T.; Sugahara, K.N.; Kotamraju, V.R.; Ruoslahti, E. C-End Rule Peptides Mediate Neuropilin-1-Dependent Cell, Vascular, and Tissue Penetration. Proc. Natl. Acad. Sci. USA 2009, 106, 16157–16162. [Google Scholar] [CrossRef]
  269. Tomita, T.; Kato, M.; Hiratsuka, S. Regulation of Vascular Permeability in Cancer Metastasis. Cancer Sci. 2021, 112, 2966–2974. [Google Scholar] [CrossRef]
  270. Wang, R.; Shen, Q.; Li, X.; Xie, C.; Lu, W.; Wang, S.; Wang, J.; Wang, D.; Liu, M. Efficacy of Inverso Isomer of CendR Peptide on Tumor Tissue Penetration. Acta Pharm. Sin. B 2018, 8, 825–832. [Google Scholar] [CrossRef]
  271. Papadimitriou, K.; Kountourakis, P.; Kottorou, A.E.; Antonacopoulou, A.G.; Rolfo, C.; Peeters, M.; Kalofonos, H.P. Follicle-Stimulating Hormone Receptor (FSHR): A Promising Tool in Oncology? Mol. Diagn. Ther. 2016, 20, 523–530. [Google Scholar] [CrossRef]
  272. Perales-Puchalt, A.; Svoronos, N.; Rutkowski, M.R.; Allegrezza, M.J.; Tesone, A.J.; Payne, K.K.; Wickramasinghe, J.; Nguyen, J.M.; O’Brien, S.W.; Gumireddy, K.; et al. Follicle-Stimulating Hormone Receptor Is Expressed by Most Ovarian Cancer Subtypes and Is a Safe and Effective Immunotherapeutic Target. Clin. Cancer Res. 2017, 23, 441–453. [Google Scholar] [CrossRef] [PubMed]
  273. Zhang, X.Y.; Chen, J.; Zheng, Y.F.; Gao, X.L.; Kang, Y.; Liu, J.C.; Cheng, M.J.; Sun, H.; Xu, C.J. Follicle-Stimulating Hormone Peptide Can Facilitate Paclitaxel Nanoparticles to Target Ovarian Carcinoma In Vivo. Cancer Res. 2009, 69, 6506–6514. [Google Scholar] [CrossRef] [PubMed]
  274. Zhang, M.M.; Zhang, M.M.; Wang, J.; Cai, Q.; Zhao, R.; Yu, Y.; Tai, H.; Zhang, X.; Xu, C. Retro-Inverso Follicle-Stimulating Hormone Peptide-Mediated Polyethylenimine Complexes for Targeted Ovarian Cancer Gene Therapy. Drug Deliv. 2018, 25, 995–1003. [Google Scholar] [CrossRef]
  275. Liu, M.; Li, C.; Pazgier, M.; Li, C.; Mao, Y.; Lv, Y.; Gu, B.; Wei, G.; Yuan, W.; Zhan, C.; et al. D-Peptide Inhibitors of the P53-MDM2 Interaction for Targeted Molecular Therapy of Malignant Neoplasms. Proc. Natl. Acad. Sci. USA 2010, 107, 14321–14326. [Google Scholar] [CrossRef]
  276. Calvanese, L.; Caporale, A.; Focà, G.; Iaccarino, E.; Sandomenico, A.; Doti, N.; Apicella, I.; Incisivo, G.M.; De Falco, S.; Falcigno, L.; et al. Targeting VEGF Receptors with Non-Neutralizing Cyclopeptides for Imaging Applications. Amino Acids 2018, 50, 321–329. [Google Scholar] [CrossRef] [PubMed]
  277. Li, X.; Xie, Z.; Xie, C.; Lu, W.; Gao, C.; Ren, H.; Ying, M.; Wei, X.; Gao, J.; Su, B.; et al. D-SP5 Peptide-Modified Highly Branched Polyethylenimine for Gene Therapy of Gastric Adenocarcinoma. Bioconjug. Chem. 2015, 26, 1494–1503. [Google Scholar] [CrossRef]
  278. Januchta, W.; Serocki, M.; Dzierzbicka, K.; Cholewiński, G.; Skladanowski, A. Synthesis of Functionalized New Conjugates of Batracylin with Tuftsin/Retro-Tuftsin Derivatives and Their Biological Evaluation. Eur. J. Med. Chem. 2015, 106, 85–94. [Google Scholar] [CrossRef] [PubMed]
  279. Torabizadeh, S.A.; Abedi, S.M.; Noaparast, Z.; Hosseinimehr, S.J. Comparative Assessment of a 99mTc Labeled H1299.2-HYNIC Peptide Bearing Two Different Co-Ligands for Tumor-Targeted Imaging. Bioorg. Med. Chem. 2017, 25, 2583–2592. [Google Scholar] [CrossRef]
  280. Xie, Z.; Shen, Q.; Xie, C.; Lu, W.; Peng, C.; Wei, X.; Li, X.; Su, B.; Gao, C.; Liu, M. Retro-Inverso Bradykinin Opens the Door of Blood–Brain Tumor Barrier for Nanocarriers in Glioma Treatment. Cancer Lett. 2015, 369, 144–151. [Google Scholar] [CrossRef]
  281. Ran, D.; Mao, J.; Zhan, C.; Xie, C.; Ruan, H.; Ying, M.; Zhou, J.; Lu, W.-L.; Lu, W. D-Retroenantiomer of Quorum-Sensing Peptide-Modified Polymeric Micelles for Brain Tumor-Targeted Drug Delivery. ACS Appl. Mater. Interfaces 2017, 9, 25672–25682. [Google Scholar] [CrossRef]
  282. Carriero, M.V.; Bifulco, K.; Ingangi, V.; Costantini, S.; Botti, G.; Ragone, C.; Minopoli, M.; Motti, M.L.; Rea, D.; Scognamiglio, G.; et al. Retro-Inverso Urokinase Receptor Antagonists for the Treatment of Metastatic Sarcomas. Sci. Rep. 2017, 7, 1312. [Google Scholar] [CrossRef]
  283. Tang, J.; Wang, Q.; Yu, Q.; Qiu, Y.; Mei, L.; Wan, D.; Wang, X.; Li, M.; He, Q. A Stabilized Retro-Inverso Peptide Ligand of Transferrin Receptor for Enhanced Liposome-Based Hepatocellular Carcinoma-Targeted Drug Delivery. Acta Biomater. 2019, 83, 379–389. [Google Scholar] [CrossRef] [PubMed]
  284. Zhang, M.; Du, M.; Qi, X.; Wang, Y.; Li, G.; Xu, C.; Zhang, X. Retro-Inversion Follicle-Stimulating Hormone Peptide-Modified Nanoparticles for Delivery of PDK2 ShRNA against Chemoresistant Ovarian Cancer by Switching Glycolysis to Oxidative Phosphorylation. Cancer Nanotechnol. 2022, 13, 23. [Google Scholar] [CrossRef]
  285. Zhang, M.; Liu, Q.; Zhang, M.; Cao, C.; Liu, X.; Zhang, M.; Li, G.; Xu, C.; Zhang, X. Enhanced Antitumor Effects of Follicle-Stimulating Hormone Receptor-Mediated Hexokinase-2 Depletion on Ovarian Cancer Mediated by a Shift in Glucose Metabolism. J. Nanobiotechnol. 2020, 18, 161. [Google Scholar] [CrossRef] [PubMed]
  286. Hilchie, A.L.; Haney, E.F.; Pinto, D.M.; Hancock, R.E.W.; Hoskin, D.W. Enhanced Killing of Breast Cancer Cells by a D-Amino Acid Analog of the Winter Flounder-Derived Pleurocidin NRC-03. Exp. Mol. Pathol. 2015, 99, 426–434. [Google Scholar] [CrossRef] [PubMed]
  287. Chen, X.; Fan, Z.; Chen, Y.; Fang, X.; Sha, X. Retro-Inverso Carbohydrate Mimetic Peptides with Annexin1-Binding Selectivity, Are Stable In Vivo, and Target Tumor Vasculature. PLoS ONE 2013, 8, e80390. [Google Scholar] [CrossRef] [PubMed]
  288. Wang, L.; Wang, N.; Zhang, W.; Cheng, X.; Yan, Z.; Shao, G.; Wang, X.; Wang, R.; Fu, C. Therapeutic Peptides: Current Applications and Future Directions. Signal Transduct. Target. Ther. 2022, 7, 48. [Google Scholar] [CrossRef]
  289. Pizzolato-Cezar, L.R.; Okuda-Shinagawa, N.M.; Teresa Machini, M. Combinatory Therapy Antimicrobial Peptide-Antibiotic to Minimize the Ongoing Rise of Resistance. Front. Microbiol. 2019, 10, 1703. [Google Scholar] [CrossRef]
Figure 1. (A) Retro (WRKLIGA), enantio (aGilkrw), and retroenantio (wrkliGa) modifications of PepH3 (AGILKRW) [33] obtained using Pymol [34]. (B) Overlaying the canonical (yellow backbone, conventional left-to-right orientation, and N- to C-terminus layout) with the re (purple backbone, right-to-left orientation, N- to C-terminus layout) version shows that the side chains adopt similar orientations while the amide bonds are reversed.
Figure 1. (A) Retro (WRKLIGA), enantio (aGilkrw), and retroenantio (wrkliGa) modifications of PepH3 (AGILKRW) [33] obtained using Pymol [34]. (B) Overlaying the canonical (yellow backbone, conventional left-to-right orientation, and N- to C-terminus layout) with the re (purple backbone, right-to-left orientation, N- to C-terminus layout) version shows that the side chains adopt similar orientations while the amide bonds are reversed.
Pharmaceutics 15 02451 g001
Figure 2. Enniatin B and its enantio form. (A) The configuration of enniatin B chiral centers. (B) Enantio-enniatin B, with the same chemical formula as enniatin B, is a mirror-image isomer with all its chiral centers reversed. Rotating enniatin B by 60 degrees in the plane illustrates the similarity.
Figure 2. Enniatin B and its enantio form. (A) The configuration of enniatin B chiral centers. (B) Enantio-enniatin B, with the same chemical formula as enniatin B, is a mirror-image isomer with all its chiral centers reversed. Rotating enniatin B by 60 degrees in the plane illustrates the similarity.
Pharmaceutics 15 02451 g002
Figure 3. Multiple models showcase the diverse mechanisms by which AMPs disrupt bacterial membranes. In the classical view of the (not mutually exclusive) mechanisms, after an initial electrostatic interaction, peptides approach the membrane and, once they reach a critical concentration threshold, can create peptide-lined pores in the barrel-stave model, or dissolve the membrane into micellar structures in the carpet model, or establish peptide-and-lipid-lined pores in the toroidal pore model. (adapted from [68]).
Figure 3. Multiple models showcase the diverse mechanisms by which AMPs disrupt bacterial membranes. In the classical view of the (not mutually exclusive) mechanisms, after an initial electrostatic interaction, peptides approach the membrane and, once they reach a critical concentration threshold, can create peptide-lined pores in the barrel-stave model, or dissolve the membrane into micellar structures in the carpet model, or establish peptide-and-lipid-lined pores in the toroidal pore model. (adapted from [68]).
Pharmaceutics 15 02451 g003
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

Carrera-Aubesart, A.; Gallo, M.; Defaus, S.; Todorovski, T.; Andreu, D. Topoisomeric Membrane-Active Peptides: A Review of the Last Two Decades. Pharmaceutics 2023, 15, 2451. https://doi.org/10.3390/pharmaceutics15102451

AMA Style

Carrera-Aubesart A, Gallo M, Defaus S, Todorovski T, Andreu D. Topoisomeric Membrane-Active Peptides: A Review of the Last Two Decades. Pharmaceutics. 2023; 15(10):2451. https://doi.org/10.3390/pharmaceutics15102451

Chicago/Turabian Style

Carrera-Aubesart, Adam, Maria Gallo, Sira Defaus, Toni Todorovski, and David Andreu. 2023. "Topoisomeric Membrane-Active Peptides: A Review of the Last Two Decades" Pharmaceutics 15, no. 10: 2451. https://doi.org/10.3390/pharmaceutics15102451

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