Next Article in Journal
Hydrophobic Mismatch Controls the Mode of Membrane-Mediated Interactions of Transmembrane Peptides
Next Article in Special Issue
Design of Membrane Active Peptides Considering Multi-Objective Optimization for Biomedical Application
Previous Article in Journal
Cellulose Membranes in the Treatment of Spent Deep Eutectic Solvent Used in the Recovery of Lignin from Lignocellulosic Biomass
Previous Article in Special Issue
NMR Studies of the Ion Channel-Forming Human Amyloid-β with Zinc Ion Concentrations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Bio-Membrane Internalization Mechanisms of Arginine-Rich Cell-Penetrating Peptides in Various Species

1
Department of Laboratory Medicine and Biotechnology, Collage of Medicine, Tzu Chi University, Hualien 970374, Taiwan
2
Graduate Institute of Microbiology and Public Health, College of Veterinary Medicine, National Chung Hsing University, Taichung 402204, Taiwan
3
Department of Biological Sciences, College of Arts, Sciences, and Business, Missouri University of Science and Technology, Rolla, MO 65409, USA
4
Department of Natural Resources and Environmental Studies, College of Environmental Studies, National Dong Hwa University, Hualien 974301, Taiwan
*
Authors to whom correspondence should be addressed.
Membranes 2022, 12(1), 88; https://doi.org/10.3390/membranes12010088
Submission received: 11 December 2021 / Revised: 6 January 2022 / Accepted: 10 January 2022 / Published: 13 January 2022

Abstract

:
Recently, membrane-active peptides or proteins that include antimicrobial peptides (AMPs), cytolytic proteins, and cell-penetrating peptides (CPPs) have attracted attention due to their potential applications in the biomedical field. Among them, CPPs have been regarded as a potent drug/molecules delivery system. Various cargoes, such as DNAs, RNAs, bioactive proteins/peptides, nanoparticles and drugs, can be carried by CPPs and delivered into cells in either covalent or noncovalent manners. Here, we focused on four arginine-rich CPPs and reviewed the mechanisms that these CPPs used for intracellular uptake across cellular plasma membranes. The varying transduction efficiencies of them alone or with cargoes were discussed, and the membrane permeability was also expounded for CPP/cargoes delivery in various species. Direct membrane translocation (penetration) and endocytosis are two principal mechanisms for arginine-rich CPPs mediated cargo delivery. Furthermore, the amino acid sequence is the primary key factor that determines the cellular internalization mechanism. Importantly, the non-cytotoxic nature and the wide applicability make CPPs a trending tool for cellular delivery.

Graphical Abstract

1. Introduction

The plasma membrane is a natural barrier that separates intracellular components from the external environment and keeps the balance of osmotic pressure. However, membrane-active peptides/proteins disrupt and modulate the integrity of the cell membrane. The membrane-active peptides are a term for the peptides that are able to interact with lipid-bilayer membranes to cause leakage as well as damage on cell membranes or to transport exogenous molecules into cells [1,2,3]. Antimicrobial peptides (AMPs), cytosolic peptides and cell-penetrating peptides (CPPs) are three major types of membrane-active peptides [1,4,5]. Here, we focus on CPPs, which are highly applied in biomedical therapeutics and regarded as a potent drug delivery system [6].

2. Cell-Penetrating Peptides (CPPs)

Numerous CPPs from natural materials or artificial synthetics were explored in recent decades [6]. The first CPP, a short peptide containing 11 cationic residues, is referred to as the protein transduction domain (PTD). This PTD was initially identified from the Tat (trans-activator of transcription) protein of the human immunodeficiency virus type 1 (HIV-1) [7,8]. Since then, various CPPs derived from the PTD of the Tat were reported [9]. Up to now, 1855 CPPs have been defined in CPPsite 2.0, a manually curated database of CPPs [10,11]. Among these 1855 CPPs, 1753 CPPs belong to linear types, whereas 102 CPPs are cyclic peptides. According to the peptide nature, CPPs were categorized into three major groups: protein derived, synthetic, and chimeric peptides [11]. However, according to their chemical and physical properties, most studies have classified CPPs into cationic, amphipathic, and hydrophobic groups [6]. Both AMPs and CPPs are cationic peptides [6,12]. One of the differences between AMPs and CPPs is the lengths of amino acid sequences. The length of AMPs is between 10 and 60 residues and the average number of amino acid residues is 33.26 [12]. Most CPPs contain less than 30 amino acids, but the shortest CPP is a peptide consisting of five residues (L5a, see Appendix A, Table A1) [13,14]. These CPPs can not only enter cells by themselves but can also take cargoes with them to cross cell membranes. The cargoes include proteins/peptides, nucleic acids, peptide nucleic acids (PNA), fluorophores, nanoparticles, drugs and small molecules [6,11]. The link between CPPs and cargoes can be covalent [15,16] or noncovalent [17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34], as well as a combination of both (called covalent and noncovalent protein transductions; CNPT) [35,36]. In order to increase the transduction efficiency and stability of the cargoes and to decrease the rate of biodegradability, various strategies of modifications on CPPs were deployed. They include, but are not limited to, the use of D-form amino acid replacements on β or γ sides [37], cis- or trans-γ-amino acid modification on branches [38], and non-standard amino acid substitutions [39]. However, various modifications of the primary sequences of CPPs, types of cargoes for delivery, combination manners of cargoes, and even concentrations of CPPs were also important factors in contributing to alternative entry routes into cells [27,30,31,34,38,40].

3. Mechanisms of Cellular Internalization

The understanding of the mechanisms of intracellular uptake for CPPs remains understudied. For example, early research proposed that the Tat protein engaged direct penetration to enter cells [41]. However, subsequent studies have suggested that both direct membrane translocation and endocytosis are employed by the Tat to enter cells [42,43]. To date, the proposed mechanisms for CPPs’ entry into cells are divided into two major categories: energy-dependent endocytosis and energy-independent direct membrane translocation (Figure 1) [44]. Endocytosis of CPPs may involve more than one of the following subtypes: clathrin-mediated endocytosis, caveolae-mediated endocytosis, clathrin- and caveolae-independent endocytosis, and macropinocytosis [44,45]. Macropinocytosis is a kind of F-actin driven lipid raft, which is irrelevant to receptors or any other proteins (Figure 1). It belongs to a non-classical endocytosis [44,45,46]. Both clathrin-mediated and caveolae-mediated endocytosis appertain to dynamin-dependent pathways, while GTPase is involved in all subtypes of endocytosis [45,47,48]. It is arduous to predict the cellular internalization mechanisms of CPPs simply based on their physical and chemical properties. Previous studies indicated that most amphipathic CPPs, such as VP22, KALA and GALA, enter cells via the energy-dependent endocytosis [43]. However, two CPPs, Pep-1 and MPG, which also belong to amphipathic CPP, use direct penetrations for cellular entry [43]. Endocytosis and direct membrane translocation have been proposed as the two major mechanisms used for CPP internalization. Studies were also conducted on several polyarginines with regard to their induction and shifting between different uptake mechanisms under various concentrations of CPPs, cargoes, or cell lines [44,49]. Therefore, in addition to the charge, some critical domains in CPPs were considered as the primary key to determine cellular entry routes. Cationic polyarginines containing many guanidinium groups were regarded as specific CPPs, and the term “arginine-magic” was coined [50,51]. In this communication, we focus on four types of arginine-rich CPPs. Nona-arginines were the backbone of peptide sequences as different modifications or domains were appended on the backbones. We illustrate the cargoes that four arginine-rich CPPs delivered and the cellular internalization mechanisms they used in the next subsections.

3.1. Synthetic Nona-Arginine (SR9)

Since the first CPP, the PTD of the Tat, was found, researchers have discovered that cationic amino acids are critical factors to increase protein transduction efficiency [41,52]. Moreover, arginine plays a more important role than lysine in affecting the plasma membrane permeability [52]. Among polyarginine peptides, octa-arginine and nona-arginine were the most commonly used CPPs in cargo deliveries [34,53]. The synthetic SR9 CPPs, containing only nine arginines without any modifications, were able to carry different cargoes, including plasmid DNA, proteins and nanoparticles, and to internalize cell membranes in various species (Table 1). However, properties of cargoes might affect the choice of entry routes. SR9 CPPs enter large unilamellar vesicles (LUVs), artificial plasma-membraned vesicles, via lipid raft inducing (Table 1) [54], while they carry plasmid DNAs to plant tissues through macropinocytosis [19]. The latter intracellular uptake mechanism was also observed in SR9/protein delivery into A549 cells, plant epidermal cells, and mouse skin cells, as well as in SR9/nanoparticle delivery into prokaryotes (Table 1) [17,22,26]. Interestingly, only SR9/quantum dot (QD) complexes employed multiple pathways for entry into mammalian A549 cancer cells [27,33,34]. In summary, both cargoes and entrance targets are important factors influencing the cellular internalization mechanisms of CPPs.

3.2. Histidine-Rich Nona-Arginine (HR9)

HR9 CPPs contain extra-modifications with one cysteine and five histidines at both N- and C-termini of the backbone SR9 sequence (Appendix A, Table A1) [27]. Imidazole side chains on histidines possess additional positive charges and may serve as the hydrogen donor or acceptor [57]. A higher zeta potential of HR9 allows for a higher degree of cell membrane disturbance and compactions of HR9/DNA nano-complexes [27,33,55,57,58]. Similar to SR9, HR9 CPPs are able to form complexes with plasmid DNAs, proteins and nanoparticles, and deliver them into mammalian cells, insect cells, paramecia, rotifers and mice (Table 1) [20,21,25,27,29,55,56]. Unlike SR9, different cargoes and entry species play no role in cellular internalization mechanisms of HR9. HR9 CPPs form complexes with either plasmid DNAs or fluorescent QDs and enter A549 cells by direct membrane translocation [25,27,29,33]. Zhang et al. designed a histidine- and arginine-rich CPP, NP1 (see Appendix A, Table A1) to deliver anticancer agent ellipticine and found that these complexes crossed the cell membrane via direct membrane translocation [57]. Although NP1 is composed of the lipid-like domain, HR9 and NP1 CPPs are all rich in histidine and arginine residues, suggesting that direct membrane translocation may rely on sufficient imidazole and guanidinium groups.

3.3. Pas Nona-Arginine (PR9)

Endocytosis may lead to cargo trapping and degradations in lysosome. Modifications of the SR9 CPPs with lysosomal escape signals become an option when cargoes are delivered by SR9 into cells via either macropinocytosis or multiple endocytic pathways (Table 1). An inverted and cleft sequence (FFLIPKG), adopted from a lysosomal aspartyl protease, cathepsin D, was chosen for the purpose of lysosomal escape, and it was named as a penetration accelerating sequence (Pas) [59]. PR9 CPPs composed of the SR9 backbone, and a Pas at the N-terminus are able to deliver plasmid DNAs and nanoparticles into A549 cells, Sf9 cells, and paramecia via classical endocytosis (Table 1) [20,21,25,27,28,31,32,33]. Although nanoparticles delivered by PR9 were co-localized with early endosomes and lysosomes at the early stage of intracellular trafficking, lysosomal escape and nuclear accumulation of these fluorescent nanoparticles were observed at the late period of delivery [31,32,33]. Dual Pas and more arginine addition formed a new CPP, known as the Pas2r12(for primary sequence see Appendix A, Table A1) [60,61]. This CPP can deliver large molecules, such as enhanced green fluorescent proteins and immunoglobulin G, through caveolae-mediated endocytosis, and enhance the release of cargoes to cytosol [60,61]. The ways in which Pas-CPPs and cargoes interact do not change the entry routes. Glucagon-like peptide-2 fused with Pas-octa-arginine by covalent bonds can enhance cellular uptake in A549 and MDCK cells via macropinocytosis, as well as translocate to mouse brains via intranasal administration [62]. Lysosomal escape and nuclear localization have also been observed in this Pas-containing arginine-rich CPP [62]. These studies inferred that the critical domain—Pas—plays the key role in endocytic entry route and lysosomal escape. The study by Takayama et al. has further supported this hypothesis. In their study, the Pas domain was replaced by four phenylalanines, and this newly designed F4R8 (primary sequence in Appendix A, Table A1) increased hydrophobicity, thus promoting membrane translocations [63].

3.4. INF7 Fusion Nona-Arginine (IR9)

Another endosomolytic peptide—INF7—derived from the influenza hemagglutinin peptide HA2, was fused with SR9 to enhance endosomal escape and cytosolic release [30]. In another study, various CPPs, such as the Tat and Penetratin, were also fused with either HA2 or INF7 to compare the membrane lytic properties [64]. INF7 fused with SR9 (called IR9) enters A549 cells by macropinocytosis, but switches to classical endocytosis when IR9 carries plasmid DNAs or nanoparticles [30]. IR9 can also enter alone or deliver DNAs or QDs as cargoes in rotifers [29]. While the INF7 domain in IR9 triggers membrane insertion using highly conserved hydrophobic residues, the polyarginine domain in IR9 executes cellular internalization [65]. Various amino acids in INF7-Tat were replaced for pH sensitivity studies, and the replacement of glutamic acid at the 6th position of INF7 (A→E) has enhanced lytic potency at lysosomal pH with marvelous pH selectivity [64]. This suggested that the glutamic acid replacement in INF7-CPPs only alters acidic lysosomal membrane without perturbing the cell membrane and, consequently, releases bioactive molecules into cytosol without any cellular injuries. Interestingly, similar to PR9, IR9 is able to enter cells via endocytosis and escape from lysosome [31,32,64]. INF7-CPPs are able to alter secondary structures in pH 5.5 and pH 7.0, while PR9 can maintain the same secondary structures in both pH conditions [31,64]. Many factors, including INF7 positions at either the N- or C-terminus of CPPs, ratios between INF7-CPPs and cargoes, and the types of cargoes, can affect the cellular uptake efficiencies [30,66,67]. Collectively, each variable entails unique properties of INF7-infused CPPs, and more extensive studies are required to identify key variables that are associated with uptake mechanisms and transduction efficiencies.

4. Evidence of Cellular Internalization

In this communication, we reveal various cargo-delivery capacities and entry mechanisms of four arginine-rich CPPs in various species (Table 1). Although types of cargoes and penetrating targets influence entry mechanisms, primary sequences play fundamental roles in cellular internalization. Among these four arginine-rich CPPs, SR9 is the backbone of the other CPPs (Table 2). The histidine-rich domains added at both ends of SR9 provide more positive charges, while reverted cathepsin D and INF7 domains give rise to membrane disturbance and lysosome escapes. Some peptide properties, such as net charges, hydrophilicities, hydrophobicities, and theoretical pI, were altered due to these modifications.
Correlations among the structure, properties, mechanism and transduction efficiency of CPPs are a sophisticated issue due to plenty of variables, such as the types and characteristics of cargoes, targets and primary sequences, as well as the chemical structure of CPPs. For example, arginine-rich CPPs usually formed complexes with cargoes in noncovalent manners because of positive/negative charged electrostatic interactions [68]. However, Kamei et al. illustrated that only three cargoes with different pI can be delivered by octa-arginine (R8), while the other 13 peptide drugs cannot [69]. The pI of cargoes and the pH value in the environment are irrelevant to cellular internalization efficiencies. In contrast, the chemical structures, such as L- and D-form, affect cellular internalization efficiencies [70,71]. D-form CPPs are resistant to enzymatic digestions and possess higher stability, which creases cellular internalization [70,71]. Primary sequences of CPPs alter not only cellular internalization efficiencies but also entry mechanisms. While different modified CPPs complexed with or without QDs as the variants, significant differences in cellular internalization efficiencies were observed [25,27,29,33]. The highest efficiency was observed in HR9/QD complexes, while the poor transduction efficiency was displayed in IR9/QD at the same combination ratio of 60 [33]. The INF7 domain in IR9 contributes the highest scale of hydrophobicity and negative charges (Table 2). Therefore, added domains on the SR9 backbone play an important role in determining CPP primary sequences and their transduction efficiencies. Peptide lengths also define the protein transduction efficiencies [70]. The highest insulin absorption is enhanced by R8, instead of 6 or 10 arginine residues [70].

5. Conclusions

CPPs have been recognized as a promising tool for intracellular deliveries of bioactive molecules due to their high cellular internalization efficiency. Arginine-rich CPPs are the most common membrane active peptides in biomedical applications. We described unique properties of four arginine-rich CPPs and their capability to internalize a variety of cargoes using different uptake mechanisms. By manipulating their structures, lysosomal escape can be achieved and higher efficiency can be accomplished. A greater understanding of correlations among structures, properties, mechanisms and transduction efficiencies can certainly facilitate the design of more effective CPPs for broader applications in the future.

Author Contributions

Conceptualization, B.R.L.; writing—original draft preparation, B.R.L.; writing—review and editing, S.-H.C., Y.-W.H. and H.-J.L.; visualization, B.R.L.; supervision, H.-J.L.; funding acquisition, H.-J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data available upon request from the corresponding authors.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Table A1. Primary Sequences of cell-penetrating peptides.
Table A1. Primary Sequences of cell-penetrating peptides.
CPPPrimary Sequence
F4R8FFFFGRRRRRRRRGC
HR9CHHHHHRRRRRRRRRHHHHHC
IR9GLFEAIEGFIENGWEGMIDGWYGRRRRRRRRR
L5aRRWQW
NP1stearyl-HHHHHHHHHHHHHHHHRRRRRRRR-NH2
Pas2r12FFLIGFFLIGRRRRRRRRRRRR
PR9FFLIPKGRRRRRRRRR
SR9RRRRRRRRR

References

  1. Almeida, P.F. Membrane-active peptides: Binding, translocation, and flux in lipid vesicles. Biochim. Biophys. Acta 2014, 1838, 2216–2227. [Google Scholar] [CrossRef] [Green Version]
  2. Avci, F.G.; Akbulut, B.S.; Ozkirimli, E. Membrane active peptides and their biophysical characterization. Biomolecules 2018, 8, 77. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Plank, C.; Zauner, W.; Wagner, E. Application of membrane-active peptides for drug and gene delivery across cellular membranes. Adv. Drug Deliv. Rev. 1998, 34, 21–35. [Google Scholar] [CrossRef]
  4. Bernheimer, A.W.; Rudy, B. Interactions between membranes and cytolytic peptides. Biochim. Biophys. Acta 1986, 864, 123–141. [Google Scholar] [CrossRef]
  5. Brand, G.D.; Ramada, M.H.S.; Genaro-Mattos, T.C.; Bloch, C., Jr. Towards an experimental classification system for membrane active peptides. Sci. Rep. 2018, 8, 1194. [Google Scholar] [CrossRef] [Green Version]
  6. Guidotti, G.; Brambilla, L.; Rossi, D. Cell-penetrating peptides: From basic research to clinics. Trends Pharmacol. Sci. 2017, 38, 406–424. [Google Scholar] [CrossRef]
  7. Frankel, A.D.; Pabo, C.O. Cellular uptake of the tat protein from human immunodeficiency virus. Cell 1988, 55, 1189–1193. [Google Scholar] [CrossRef]
  8. Green, M.; Loewenstein, P.M. Autonomous functional domains of chemically synthesized human immunodeficiency virus tat trans-activator protein. Cell 1988, 55, 1179–1188. [Google Scholar] [CrossRef]
  9. Zhang, X.; Zhang, X.; Wang, F. Intracellular transduction and potential of Tat PTD and its analogs: From basic drug delivery mechanism to application. Expert Opin. Drug Deliv. 2012, 9, 457–472. [Google Scholar] [CrossRef]
  10. Agrawal, P.; Bhalla, S.; Usmani, S.S.; Singh, S.; Chaudhary, K.; Raghava, G.P.; Gautam, A. CPPsite 2.0: A repository of experimentally validated cell-penetrating peptides. Nucleic Acids Res. 2016, 44, D1098–D1103. [Google Scholar] [CrossRef]
  11. Kardani, K.; Bolhassani, A. Cppsite 2.0: An available database of experimentally validated cell-penetrating peptides predicting their secondary and tertiary structures. J. Mol. Biol. 2021, 433, 166703. [Google Scholar] [CrossRef]
  12. Huan, Y.; Kong, Q.; Mou, H.; Yi, H. Antimicrobial peptides: Classification, design, application and research progress in multiple fields. Front. Microbiol. 2020, 11, 582779. [Google Scholar] [CrossRef]
  13. Liu, B.R.; Huang, Y.W.; Aronstam, R.S.; Lee, H.J. Identification of a short cell-penetrating peptide from bovine lactoferricin for intracellular delivery of DNA in human A549 cells. PLoS ONE 2016, 11, e0150439. [Google Scholar] [CrossRef] [PubMed]
  14. Stiltner, J.; McCandless, K.; Zahid, M. Cell-penetrating peptides: Applications in tumor diagnosis and therapeutics. Pharmaceutics 2021, 13, 890. [Google Scholar] [CrossRef]
  15. Yang, W.C.; Patel, K.G.; Lee, J.; Ghebremariam, Y.T.; Wong, H.E.; Cooke, J.P.; Swartz, J.R. Cell-free production of transducible transcription factors for nuclear reprogramming. Biotechnol. Bioeng. 2009, 104, 1047–1058. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Chang, M.; Chou, J.C.; Lee, H.J. Cellular internalization of fluorescent proteins via arginine-rich intracellular delivery peptide in plant cells. Plant Cell Physiol. 2005, 46, 482–488. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Chang, M.; Chou, J.C.; Chen, C.P.; Liu, B.R.; Lee, H.J. Noncovalent protein transduction in plant cells by macropinocytosis. New Phytol. 2007, 174, 46–56. [Google Scholar] [CrossRef]
  18. Chang, M.; Huang, Y.W.; Aronstam, R.S.; Lee, H.J. Cellular delivery of noncovalently-associated macromolecules by cell-penetrating peptides. Curr. Pharm. Biotechnol. 2014, 15, 267–275. [Google Scholar] [CrossRef] [PubMed]
  19. Chen, C.P.; Chou, J.C.; Liu, B.R.; Chang, M.; Lee, H.J. Transfection and expression of plasmid DNA in plant cells by an arginine-rich intracellular delivery peptide without protoplast preparation. FEBS Lett. 2007, 581, 1891–1897. [Google Scholar] [CrossRef] [Green Version]
  20. Chen, Y.J.; Liu, B.R.; Dai, Y.H.; Lee, C.Y.; Chan, M.H.; Chen, H.H.; Chiang, H.J.; Lee, H.J. A gene delivery system for insect cells mediated by arginine-rich cell-penetrating peptides. Gene 2012, 493, 201–210. [Google Scholar] [CrossRef]
  21. Dai, Y.H.; Liu, B.R.; Chiang, H.J.; Lee, H.J. Gene transport and expression by arginine-rich cell-penetrating peptides in Paramecium. Gene 2011, 489, 89–97. [Google Scholar] [CrossRef]
  22. Hou, Y.W.; Chan, M.H.; Hsu, H.R.; Liu, B.R.; Chen, C.P.; Chen, H.H.; Lee, H.J. Transdermal delivery of proteins mediated by non-covalently associated arginine-rich intracellular delivery peptides. Exp. Dermatol. 2007, 16, 999–1006. [Google Scholar] [CrossRef]
  23. Lee, C.Y.; Li, J.F.; Liou, J.S.; Charng, Y.C.; Huang, Y.W.; Lee, H.J. A gene delivery system for human cells mediated by both a cell-penetrating peptide and a piggyBac transposase. Biomaterials 2011, 32, 6264–6276. [Google Scholar] [CrossRef]
  24. Li, J.F.; Huang, Y.; Chen, R.L.; Lee, H.J. Induction of apoptosis by gene transfer of human TRAIL mediated by arginine-rich intracellular delivery peptides. Anticancer Res. 2010, 30, 2193–2202. [Google Scholar] [PubMed]
  25. Liu, B.R.; Chen, H.H.; Chan, M.H.; Huang, Y.W.; Aronstam, R.S.; Lee, H.J. Three arginine-rich cell-penetrating peptides facilitate cellular internalization of red-emitting quantum dots. J. Nanosci. Nanotechnol. 2015, 15, 2067–2078. [Google Scholar] [CrossRef]
  26. Liu, B.R.; Huang, Y.W.; Aronstam, R.S.; Lee, H.J. Comparative mechanisms of protein transduction mediated by cell-penetrating peptides in prokaryotes. J. Membr. Biol. 2015, 248, 355–368. [Google Scholar] [CrossRef]
  27. Liu, B.R.; Huang, Y.W.; Winiarz, J.G.; Chiang, H.J.; Lee, H.J. Intracellular delivery of quantum dots mediated by a histidine- and arginine-rich HR9 cell-penetrating peptide through the direct membrane translocation mechanism. Biomaterials 2011, 32, 3520–3537. [Google Scholar] [CrossRef] [PubMed]
  28. Liu, B.R.; Lin, M.D.; Chiang, H.J.; Lee, H.J. Arginine-rich cell-penetrating peptides deliver gene into living human cells. Gene 2012, 505, 37–45. [Google Scholar] [CrossRef]
  29. Liu, B.R.; Liou, J.S.; Chen, Y.J.; Huang, Y.W.; Lee, H.J. Delivery of nucleic acids, proteins, and nanoparticles by arginine-rich cell-penetrating peptides in rotifers. Mar. Biotechnol. 2013, 15, 584–595. [Google Scholar] [CrossRef] [PubMed]
  30. Liu, B.R.; Liou, J.S.; Huang, Y.W.; Aronstam, R.S.; Lee, H.J. Intracellular delivery of nanoparticles and DNAs by IR9 cell-penetrating peptides. PLoS ONE 2013, 8, e64205. [Google Scholar] [CrossRef] [Green Version]
  31. Liu, B.R.; Lo, S.Y.; Liu, C.C.; Chyan, C.L.; Huang, Y.W.; Aronstam, R.S.; Lee, H.J. Endocytic trafficking of nanoparticles delivered by cell-penetrating peptides comprised of nona-arginine and a penetration accelerating sequence. PLoS ONE 2013, 8, e67100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Liu, B.R.; Winiarz, J.G.; Moon, J.S.; Lo, S.Y.; Huang, Y.W.; Aronstam, R.S.; Lee, H.J. Synthesis, characterization and applications of carboxylated and polyethylene-glycolated bifunctionalized InP/ZnS quantum dots in cellular internalization mediated by cell-penetrating peptides. Colloids Surf. B Biointerfaces 2013, 111, 162–170. [Google Scholar] [CrossRef] [PubMed]
  33. Liu, B.R.; Chiang, H.J.; Huang, Y.W.; Chan, M.H.; Chen, H.H.; Lee, H.J. Cellular internalization of quantum dots mediated by cell-penetrating peptides. Pharm. Nanotechnol. 2013, 1, 151–161. [Google Scholar]
  34. Xu, Y.; Liu, B.R.; Lee, H.J.; Shannon, K.B.; Winiarz, J.G.; Wang, T.C.; Chiang, H.J.; Huang, Y.W. Nona-arginine facilitates delivery of quantum dots into cells via multiple pathways. J. Biomed. Biotechnol. 2010, 2010, 948543. [Google Scholar] [CrossRef]
  35. Lu, S.W.; Hu, J.W.; Liu, B.R.; Lee, C.Y.; Li, J.F.; Chou, J.C.; Lee, H.J. Arginine-rich intracellular delivery peptides synchronously deliver covalently and noncovalently linked proteins into plant cells. J. Agric. Food Chem. 2010, 58, 2288–2294. [Google Scholar] [CrossRef]
  36. Hu, J.W.; Liu, B.R.; Wu, C.Y.; Lu, S.W.; Lee, H.J. Protein transport in human cells mediated by covalently and noncovalently conjugated arginine-rich intracellular delivery peptides. Peptides 2009, 30, 1669–1678. [Google Scholar] [CrossRef]
  37. Farrera-Sinfreu, J.; Giralt, E.; Castel, S.; Albericio, F.; Royo, M. Cell-penetrating cis-gamma-amino-l-proline-derived peptides. J. Am. Chem. Soc. 2005, 127, 9459–9468. [Google Scholar] [CrossRef]
  38. Illa, O.; Ospina, J.; Sánchez-Aparicio, J.E.; Pulido, X.; Abengozar, M.; Gaztelumendi, N.; Carbajo, D.; Nogués, C.; Rivas, L.; Maréchal, J.D.; et al. Hybrid cyclobutane/proline-containing peptidomimetics: The conformational constraint influences their cell-penetration ability. Int. J. Mol. Sci. 2021, 22, 5092. [Google Scholar] [CrossRef]
  39. Ezzat, K.; Andaloussi, S.E.; Zaghloul, E.M.; Lehto, T.; Lindberg, S.; Moreno, P.M.; Viola, J.R.; Magdy, T.; Abdo, R.; Guterstam, P.; et al. PepFect 14, a novel cell-penetrating peptide for oligonucleotide delivery in solution and as solid formulation. Nucleic Acids Res. 2011, 39, 5284–5298. [Google Scholar] [CrossRef]
  40. Pae, J.; Säälik, P.; Liivamägi, L.; Lubenets, D.; Arukuusk, P.; Langel, Ü.; Pooga, M. Translocation of cell-penetrating peptides across the plasma membrane is controlled by cholesterol and microenvironment created by membranous proteins. J. Control. Release 2014, 192, 103–113. [Google Scholar] [CrossRef] [PubMed]
  41. 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] [CrossRef] [PubMed] [Green Version]
  42. Koren, E.; Torchilin, V.P. Cell-penetrating peptides: Breaking through to the other side. Trends Mol. Med. 2012, 18, 385–393. [Google Scholar] [CrossRef]
  43. Layek, B.; Lipp, L.; Singh, J. Cell penetrating peptide conjugated chitosan for enhanced delivery of nucleic acid. Int. J. Mol. Sci. 2015, 16, 28912–28930. [Google Scholar] [CrossRef] [Green Version]
  44. Ruseska, I.; Zimmer, A. Internalization mechanisms of cell-penetrating peptides. Beilstein J. Nanotechnol. 2020, 11, 101–123. [Google Scholar] [CrossRef]
  45. Conner, S.D.; Schmid, S.L. Regulated portals of entry into the cell. Nature 2003, 422, 37–44. [Google Scholar] [CrossRef]
  46. Wadia, J.S.; Stan, R.V.; Dowdy, S.F. Transducible TAT-HA fusogenic peptide enhances escape of TAT-fusion proteins after lipid raft macropinocytosis. Nat. Med. 2004, 10, 310–315. [Google Scholar] [CrossRef] [PubMed]
  47. Milosevic, I. Revisiting the role of clathrin-mediated endoytosis in synaptic vesicle recycling. Front. Cell. Neurosci. 2018, 12, 27. [Google Scholar] [CrossRef] [Green Version]
  48. Rennick, J.J.; Johnston, A.P.R.; Parton, R.G. Key principles and methods for studying the endocytosis of biological and nanoparticle therapeutics. Nat. Nanotechnol. 2021, 16, 266–276. [Google Scholar] [CrossRef]
  49. Crosio, M.A.; Via, M.A.; Cámara, C.I.; Mangiarotti, A.; Del Pópolo, M.G.; Wilke, N. Interaction of a polyarginine peptide with membranes of different mechanical properties. Biomolecules 2019, 9, 625. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Vazdar, M.; Heyda., J.; Mason, P.E.; Tesei, G.; Allolio, C.; Lund, M.; Jungwirth, P. Arginine “Magic”: Guanidinium like-charge ion pairing from aqueous salts to cell penetrating peptides. Acc. Chem. Res. 2018, 51, 1455–1464. [Google Scholar] [CrossRef] [Green Version]
  51. Verbeek, S.F.; Awasthi, N.; Teiwes, N.K.; Mey, I.; Hub, J.S.; Janshoff, A. How arginine derivatives alter the stability of lipid membranes: Dissecting the roles of side chains, backbone and termini. Eur. Biophys. J. 2021, 50, 127–142. [Google Scholar] [CrossRef] [PubMed]
  52. Futaki, S. Arginine-rich peptides: Potential for intracellular delivery of macromolecules and the mystery of the translocation mechanisms. Int. J. Pharm. 2002, 245, 1–7. [Google Scholar] [CrossRef]
  53. Melikov, K.; Hara, A.; Yamoah, K.; Zaitseva, E.; Zaitsev, E.; Chernomordik, L.V. Efficient entry of cell-penetrating peptide nona-arginine into adherent cells involves a transient increase in intracellular calcium. Biochem. J. 2015, 471, 221–230. [Google Scholar] [CrossRef] [Green Version]
  54. Almeida, C.; Maniti, O.; Di Pisa, M.; Swiecicki, J.M.; Ayala-Sanmartin, J. Cholesterol re-organisation and lipid de-packing by arginine-rich cell penetrating peptides: Role in membrane translocation. PLoS ONE 2019, 14, e0210985. [Google Scholar] [CrossRef] [PubMed]
  55. Alizadeh, S.; Irani, S.; Bolhassani, A.; Sadat, S.M. HR9: An important cell penetrating peptide for delivery of HCV NS3 DNA into HEK-293T cells. Avicenna J. Med. Biotechnol. 2020, 12, 44–51. [Google Scholar] [PubMed]
  56. Rostami, B.; Irani, S.; Bolhassani, A.; Cohan, R.A. Gene and protein delivery using four cell penetrating peptides for HIV-1 vaccine development. IUBMB Life 2019, 71, 1619–1633. [Google Scholar] [CrossRef]
  57. Zhang, L.; Xu, J.; Wang, F.; Ding, Y.; Wang, T.; Jin, G.; Martz, M.; Gui, Z.; Ouyang, P.; Chen, P. Histidine-rich cell-penetrating peptide for cancer drug delivery and its uptake mechanism. Langmuir 2019, 35, 3513–3523. [Google Scholar] [CrossRef]
  58. Liu, B.R.; Huang, Y.W.; Korivi, M.; Lo, S.Y.; Aronstam, R.S.; Lee, H.J. The primary mechanism of cellular internalization for a short cell-penetrating peptide as a nano-scale delivery system. Curr. Pharm. Biotechnol. 2017, 18, 569–584. [Google Scholar] [CrossRef] [PubMed]
  59. Takayama, K.; Nakase, I.; Michiue, H.; Takeuchi, T.; Tomizawa, K.; Matsui, H.; Futaki, S. Enhanced intracellular delivery using arginine-rich peptides by the addition of penetration accelerating sequences (Pas). J. Control. Release 2009, 138, 128–133. [Google Scholar] [CrossRef] [Green Version]
  60. Okuda, A.; Futaki, S. Protein delivery to cytosol by cell-penetrating peptide bearing tandem repeat penetration-accelerating sequence. Methods Mol. Biol. 2022, 2383, 265–273. [Google Scholar]
  61. Okuda, A.; Tahara, S.; Hirose, H.; Takeuchi, T.; Nakase, I.; Ono, A.; Takehashi, M.; Tanaka, S.; Futaki, S. Oligoarginine-bearing tandem repeat penetration-accelerating sequence delivers protein to cytosol via caveolae-mediated endocytosis. Biomacromolecules 2019, 20, 1849–1859. [Google Scholar] [CrossRef] [PubMed]
  62. Akita, T.; Kimura, R.; Akaguma, S.; Nagai, M.; Nakao, Y.; Tsugane, M.; Suzuki, H.; Oka, J.I.; Yamashita, C. Usefulness of cell-penetrating peptides and penetration accelerating sequence for nose-to-brain delivery of glucagon-like peptide-2. J. Control. Release 2021, 335, 575–583. [Google Scholar] [CrossRef] [PubMed]
  63. Takayama, K.; Hirose, H.; Tanaka, G.; Pujals, S.; Katayama, S.; Nakase, I.; Futaki, S. Effect of the attachment of a penetration accelerating sequence and the influence of hydrophobicity on octaarginine-mediated intracellular delivery. Mol. Pharm. 2012, 9, 1222–1230. [Google Scholar] [CrossRef]
  64. Algayer, B.; O’Brien, A.; Momose, A.; Murphy, D.J.; Procopio, W.; Tellers, D.M.; Tucker, T.J. Novel pH selective, highly lytic peptides based on a chimeric influenza hemagglutinin peptide/cell penetrating peptide motif. Molecules 2019, 24, 2079. [Google Scholar] [CrossRef] [Green Version]
  65. Cross, K.J.; Wharton, S.A.; Skehel, J.J.; Wiley, D.C.; Steinhauer, D.A. Studies on influenza haemagglutinin fusion peptide mutants generated by reverse genetics. EMBO J. 2001, 20, 4432–4442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Liou, J.S.; Liu, B.R.; Martin, A.L.; Huang, Y.W.; Chiang, H.J.; Lee, H.J. Protein transduction in human cells is enhanced by cell-penetrating peptides fused with an endosomolytic HA2 sequence. Peptides 2012, 37, 273–284. [Google Scholar] [CrossRef]
  67. Neundorf, I.; Rennert, R.; Hoyer, J.; Schramm, F.; Löbner, K.; Kitanovic, I.; Wölfl, S. Fusion of a short HA2-derived peptide sequence to cell-penetrating peptides improves cytosolic uptake, but enhances cytotoxic activity. Pharmaceuticals 2009, 2, 49–65. [Google Scholar] [CrossRef] [Green Version]
  68. Habault, J.; Poyet, J.L. Recent advances in cell penetrating peptide-based anticancer therapies. Molecules 2019, 24, 927. [Google Scholar] [CrossRef] [Green Version]
  69. Kamei, N.; Morishita, M.; Takayama, K. Importance of intermolecular interaction on the improvement of intestinal therapeutic peptide/protein absorption using cell-penetrating peptides. J. Control. Release 2009, 136, 179–186. [Google Scholar] [CrossRef]
  70. Morishita, M.; Kamei, N.; Ehara, J.; Isowa, K.; Takayama, K. A novel approach using functional peptides for efficient intestinal absorption of insulin. J. Control. Release 2007, 118, 177–184. [Google Scholar] [CrossRef]
  71. 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] [CrossRef] [PubMed]
Figure 1. Different cellular entry routes for either cell-penetrating peptides (CPPs) alone or CPP/cargo complexes. Direct membrane translocation and endocytosis are two major routes which have been proposed. Endocytic pathways are further divided into four categories: macropinocytosis, clathrin-mediated endocytosis, caveolae-mediated endocytosis, and clathrin- and caveolae-independent endocytosis.
Figure 1. Different cellular entry routes for either cell-penetrating peptides (CPPs) alone or CPP/cargo complexes. Direct membrane translocation and endocytosis are two major routes which have been proposed. Endocytic pathways are further divided into four categories: macropinocytosis, clathrin-mediated endocytosis, caveolae-mediated endocytosis, and clathrin- and caveolae-independent endocytosis.
Membranes 12 00088 g001
Table 1. Primary sequences of four arginine-rich CPPs and their cellular internalization mechanisms.
Table 1. Primary sequences of four arginine-rich CPPs and their cellular internalization mechanisms.
CPPsPeptide SequencesCargoesEntrance TargetsMechanismsReferences
SR9RRRRRRRRRartificial large unilamellar vesicles (LUVs) in 100 nm diameterlipid raft[54]
plasmid DNAsplant tissuesmacropinocytosis[19]
A549 cells, Sf9 cells, parameciaunknown[20,21,23,24,28]
proteinsA549 cells, plant epidermal cells, mouse skin cellsmacropinocytosis[17,22]
nanoparticlesA549 cellsmultiple pathways[33,34]
prokaryotesmacropinocytosis[26]
HR9CHHHHHRRRRRRRRRHHHHHCA549 cells, rotifers, parameciaunknown[21,28,29]
plasmid DNAsA549 cellsdirect membrane translocation[29]
HEK293T cells, Sf9 cells, rotifers, paramecia, miceunknown[20,21,29,55,56]
proteinsrotifersunknown[29]
nanoparticlesA549 cellsdirect membrane translocation[25,27,33]
rotifersunknown[29]
PR9FFLIPKGRRRRRRRRRplasmid DNAsA549 cells, Sf9 cells, parameciaunknown[20,21,28,31]
nanoparticlesA549 cellsclassical endocytosis[25,27,31,32,33]
IR9GLFEAIEGFIENGWEGMIDGWYGRRRRRRRRRA549 cellsmacropinocytosis[30]
rotifersunknown[29]
plasmid DNAsA549 cellsclassical endocytosis[30]
rotifersunknown[29]
proteinsrotifersunknown[29]
nanoparticlesA549 cellsclassical endocytosis[30]
rotifersunknown[29]
Table 2. The properties of four arginine-rich CPPs and their additional modified domains.
Table 2. The properties of four arginine-rich CPPs and their additional modified domains.
CPPAdditional Modified DomainNet Charge at pH 7.0 1Hydrophilicity 1Hydrophobicity 2pI 2
Full SequenceDomainFull SequenceDomainFull SequenceDomainFull SequenceDomain
SR9+9.03.00.7713.4
HR9polyhistidine+9.82+0.820.95−0.58−25.32−6.3412.87.5
PR9reverted cathepsin D+9.95+1.01.2−0.819.7431.5213.010.1
IR9INF7 domain+4.0−5.00.6−0.3452.2565.7311.92.8
1 Bachem Peptide Calculator. 2 Thermo Fisher Scientific Peptide Analyzing Tool.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, B.R.; Chiou, S.-H.; Huang, Y.-W.; Lee, H.-J. Bio-Membrane Internalization Mechanisms of Arginine-Rich Cell-Penetrating Peptides in Various Species. Membranes 2022, 12, 88. https://doi.org/10.3390/membranes12010088

AMA Style

Liu BR, Chiou S-H, Huang Y-W, Lee H-J. Bio-Membrane Internalization Mechanisms of Arginine-Rich Cell-Penetrating Peptides in Various Species. Membranes. 2022; 12(1):88. https://doi.org/10.3390/membranes12010088

Chicago/Turabian Style

Liu, Betty Revon, Shiow-Her Chiou, Yue-Wern Huang, and Han-Jung Lee. 2022. "Bio-Membrane Internalization Mechanisms of Arginine-Rich Cell-Penetrating Peptides in Various Species" Membranes 12, no. 1: 88. https://doi.org/10.3390/membranes12010088

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