Next Article in Journal
Therapeutic Strategies and New Intervention Points in Chronic Hepatitis Delta Virus Infection
Next Article in Special Issue
Membranotropic Cell Penetrating Peptides: The Outstanding Journey
Previous Article in Journal
Terpenoids from the Octocoral Sinularia gaweli
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Intracellular Delivery of Molecular Cargo Using Cell-Penetrating Peptides and the Combination Strategies

1
Department of Basic Medical Science, Huzhou University School of Medicine, Huzhou 313000, China
2
Department of General, Visceral and Thoracic Surgery, University Medical Center Hamburg-Eppendorf, Martinistr. 52, 20246 Hamburg, Germany
3
Moores Cancer Center, University of California San Diego, La Jolla, CA 92093-0820, USA
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2015, 16(8), 19518-19536; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms160819518
Submission received: 21 May 2015 / Revised: 11 July 2015 / Accepted: 30 July 2015 / Published: 18 August 2015
(This article belongs to the Special Issue Cell-Penetrating Peptides)

Abstract

:
Cell-penetrating peptides (CPPs) can cross cellular membranes in a non-toxic fashion, improving the intracellular delivery of various molecular cargos such as nanoparticles, small molecules and plasmid DNA. Because CPPs provide a safe, efficient, and non-invasive mode of transport for various cargos into cells, they have been developed as vectors for the delivery of genetic and biologic products in recent years. Most common CPPs are positively charged peptides. While delivering negatively charged molecules (e.g., nucleic acids) to target cells, the internalization efficiency of CPPs is reduced and inhibited because the cationic charges on the CPPs are neutralized through the covering of CPPs by cargos on the structure. Even under these circumstances, the CPPs can still be non-covalently complexed with the negatively charged molecules. To address this issue, combination strategies of CPPs with other typical carriers provide a promising and novel delivery system. This review summarizes the latest research work in using CPPs combined with molecular cargos including liposomes, polymers, cationic peptides, nanoparticles, adeno-associated virus (AAV) and calcium for the delivery of genetic products, especially for small interfering RNA (siRNA). This combination strategy remedies the reduced internalization efficiency caused by neutralization.

Graphical Abstract

1. Introduction

Cell-penetrating peptides (CPPs) are peptides composed of between four to 30 amino acids that are usually rich in arginine and lysine residues, and are characterized by their unique internalization properties. CPPs include protein transduction domain (PTD) [1,2], trojan peptides [3], arginine-rich peptides [4], and vectocell peptides, which are also known as Diatos Peptide Vectors (DPV) [5]. In general, the entry mechanisms of CPPs include (1) direct penetration (energy-independent pathway) into the cell membrane; (2) endocytosis-mediated entry (energy dependent pathway) that the plasma membrane folds inward to bring molecular cargos into cells; and (3) translocation through the formation of a transitory inverted micelles, which are aggregates of colloidal surfactants where the polar groups are concentrated in the interior and the lipophilic groups extend outward into the solvent. Although significant efforts have been made to elucidate the exact translocation mechanisms of the CPPs [6,7,8], various factors, such as the conjugation of molecular cargos and CPPs, cell types, concentration, and the types and structure of CPPs, may affect the internalization of CPPs [9,10,11,12]. The cellular uptake of therapeutic cargos becomes a major obstacle in the clinical setting and biological fields due to the lipid bilayer of the cell membrane. Since the first cell-penetrating protein was found and derived from immunodeficiency virus (HIV) with a high cellular uptake efficiency [13], numerous CPPs have been gradually discovered and utilized as carriers in the past decades. Some of the CPPs are summarized in Table 1, and other CPPs can be found in the database [14]. Among the CPPs, three main penetration mechanisms are as follow: (1) direct penetration [7], a reaction that is more favorable at high concentrations of CPPs—it occurs via an energy-independent pathway by a direct electrostatic interaction of positively charged CPPs with negatively charged phospholipids or heparin sulfate of membrane components; (2) Endocytosis [15,16], an energy-dependent process of cellular ingestion by which cells adsorb the external materials; and (3) Macropinocytosis, a form of endocytosis that accompanies cell surface ruffling [17]. Furthermore, a recent research reported that calcium influx is also involved in the process of intracellular delivery of CPPs [18].
Table 1. Representative CPPs.
Table 1. Representative CPPs.
NameSequenceReferences
Poly-R peptidesRx *[19,20,2122]
TAT (47–57)YGRKKRRQRRR[23]
TAT (49–60)RKKRRQRRRPPQ [21]
PTD4YARAAARQARA[24]
PTD5RRQRRTSKLMKRGG[25]
TP10AGYLLGKINLKALAALAKKIL[26]
M918MVTVLFRRLRIRRACGPPRVRV[27]
pAntp (43–58) [Penetratin]RQIKIWFQNRRMKWKK[21]
KNOKQINNWFINQRKRHWK[28]
Hph-1YARVRRRGPRR[29]
HIV-1 Rev (34–50)
ANP (43–58) [Antennapedia]
TRQARRNRRRRWRERQR-GC
RQIKIWFQNRRMKWKK-GC-CO
[30]
PODGGG [ARKKAAKA] 4[31]
S413-PVALWKTLLKKVLKAPKKKRKV[32]
S41CVQWSLLRGYQPC[33]
pVECLLIILRRRIRKQAHAHSK[34]
* x = 4, 8, 10, 11.
The CPPs can transfer various molecule cargoes including peptides [35], proteins [36,37,38], plasmid DNA [39,40], and nanoparticles [39,41,42]. Recently, Dr. Singh and his colleagues reported that CPPs tethered bi-ligand liposomes and short-chain CPPs are capable of crossing the blood brain barrier (BBB) [2,43,44]. The frequency and kinetics of CPPs transfection varies between different cell types. Furuhata et al. [19] showed the Arg4 was the most efficient among the (Arg)n (n = 4, 6, 8, 10) when internalized in HeLa cells, and the efficiency decreased depending on the length of oligo-Arg. In living neurons, the R11 peptide showed the highest transduction efficiency when compared with other Rx peptides such as R5, R7 and R9 [45]. Trans-Activator of Transcription (TAT), an arginine-rich CPP, was more efficient than R9 in living neurons; however, the conclusion was the contrary in rat alveolar epithelial cells [46].
Although CPPs have been notably developed to mediate the intracellular delivery of the nucleic acids with negative charge [1,47], the major drawback is the free high cationic CPPs that could bind to negatively charged nucleic acids molecules. This binding results in the aggregation of CPPs-nucleic acids and cytotoxicity. Furthermore, the ability of cell penetration will be inhibited or lost due to the neutralization of the CPPs by negatively charged molecules [48,49]. To overcome these difficulties and improve the delivery efficiency of genetic products (particularly siRNA), other carrier systems can be developed via the conjugation with CPPs (Figure 1).
Figure 1. Enhanced drug delivery system by the combination of CPPs with other carriers. Drugs (far left) can be incorporated into the typical carrier systems, which can be modified with CPPs. Under this circumstance, the CPPs-modified drug carrier system not only dramatically enhances the intracellular delivery efficiency, but also improves the endosomal escape through the alteration of its biodistribution.
Figure 1. Enhanced drug delivery system by the combination of CPPs with other carriers. Drugs (far left) can be incorporated into the typical carrier systems, which can be modified with CPPs. Under this circumstance, the CPPs-modified drug carrier system not only dramatically enhances the intracellular delivery efficiency, but also improves the endosomal escape through the alteration of its biodistribution.
Ijms 16 19518 g001

2. CPPs-Modified Liposome

Once hydrophilic lipid and hydrophobic lipid chain at opposite ends, they form liposomes when exposed to water at appropriate conditions. According to the size of lamellarity and formation, the liposome is classified into multilamellar vesicles (500 to 5000 nm), small unilamellar vesicles (around 100 nm), and large unilamellar vesicles (200 to 800 nm). During the liposome formation, the water-soluble drug molecules will also be packed into the inner water space of liposome [50]; this specific geometry protects the active agent from the destructive tendencies of the external environment. In addition, liposomes are capable of crossing membranes to deliver its contents into cells and cell compartments. However, while it can transfer diverse active components to the target cells, the low transfer efficiency limits its further application in the clinical setting.
CPPs-modified liposome can improve the efficiency of cell penetration [51]. The efficiency of liposome surface modified with Antennapedia (Antp, aa 43–58) or TAT (47–57) is 15 to 25 folds higher than that of non-modified liposome [52]. Both CPPs-modified [53] and octaarginine-coated [53,54] liposomes resulted in an increased rate of liposome uptake. The cell penetration efficiency of TAT-liposome is dramatically improved by 1000 times as opposed to the TAT only [9]. Consequently, CPPs-modified polyarginine8 (R8)-liposome has been used for siRNA delivery into lung cancer cell lines with remarkably higher transfection efficiency. The siRNA degradation in the blood serum is significantly inhibited, and showed a lower non-specific toxicity [55]. Although there is no significant difference between cellular uptake of R8-liposome and the conventional cationic liposome [56], the CPPs-modified liposome can improve the capacity of the CPPs endosomal escape from the liposome surface as well as the cargo delivery efficiency of the liposome. Thus, there are two important parameters in the CPPs-modified liposome. One is the threshold amount of CPPs on the liposome surface, and the other is the type of CPPs: (1) peptides derived from proteins; (2) chimeric peptides that are formed by the fusion of two natural sequences; and (3) synthetic CPPs, which are rationally designed sequences usually based on structure–activity studies.
To circumvent the fast elimination of liposome in blood, surface-modification methods, such as the PEG-modified strategy, are used to achieve long circulation of liposomes in vivo. The CPPs TAT peptide can be attached to the liposome surface via the liposomal NGPE (TAT-LIP-NGPE) or the PEG spacer by coupling with pNP-PEG-PE (LIP-pNP-PEG-TAT) [57], as shown in Figure 2. While the long PEG (PEG 5000) chains attached to the liposome might interfere with the TAT peptide in cell surface interactions and decrease the effectiveness, the TAT peptide-modified liposomes, on the other hand, have no obstacles for this interaction. PEG 2000 [58] or PEG 3000 [57] and TAT-modified liposomes can also deliver target drugs efficiently into the cells. Although TAT, as a cellular uptake enhancer, can be hidden by long chain PEG, this problem can be avoided by degrading the latter before TAT penetrates the cell membrane, thus protecting TAT from degradation [59].
Figure 2. CPPs-modified liposome. (A) The original liposome; (B) The CPPs conjugated directly to the surface of the liposome; (C) The CPPs modified to the surface of PEG first, and then conjugated to the liposome. The blue circle in the figure panel typifies liposome, # represents the CPPs, Δ stands for other modified cargo (e.g., PEG) and * symbolizes the transferred cargo.
Figure 2. CPPs-modified liposome. (A) The original liposome; (B) The CPPs conjugated directly to the surface of the liposome; (C) The CPPs modified to the surface of PEG first, and then conjugated to the liposome. The blue circle in the figure panel typifies liposome, # represents the CPPs, Δ stands for other modified cargo (e.g., PEG) and * symbolizes the transferred cargo.
Ijms 16 19518 g002
The key barrier for the treatment of central nervous system (CNS) diseases is the BBB penetration. CPPs exhibited the promising advantage of freely trafficking across the BBB. With the aim of increasing the amount of drug that could reach the CNS system, CPPs-conjugated liposomes were produced and showed higher brain delivery efficiency [60,61]. Similarly to CPPs, TAT-conjugated PEGylated magnetic polymeric liposomes (MPLs), TAT-PEG-MPLs are consumed more in the injured spinal cord and inside the neuron when compared to non-TAT-modified MPLs [62].

3. Combination of CPPs with Polymer

Polymer is a popular drug-carrying material that makes nanoparticles and micelles, which can deliver siRNA, DNA, protein, as well as hydrophilic and hydrophobic cargos, and the cationic polymer is a new type of siRNA delivery vector. Currently, multiple substructures have been established in one carrier, and each of them has a special function in the siRNA delivery process. For example, Poly-l-Lysine (PLL), a synthetic chiral polymer, acts as the positive-charged backbone that binds to nucleic acids. To increase the solubility and reduce aggregation, Polyethylene glycol (PEG) can be added in. Additionally, to enable the endosomal escape of this carrier, the DMMAnMel-shielded melittin (CPPs) can also be added in [63].
CPPs conjugated polymer enhances the cellular uptake of the molecular cargos, including the chemotherapy drug doxorubicin and the model molecule coumarin, when compared to the polymer without CPPs conjugation [64,65,66,67]. Cellular uptake efficiency of 5(6)-carboxy fluoresce transferred with the R9-polymer is increased 25 times when compared with that of the polymer [68]. In vitro data demonstrated that TAT modified cystamine bisacrylamide-diaminohexane (CBA-DAH) polymer enhanced cellular uptake [67]. In addition, TAT-poly-lysine [69] and the oligoarginine-linked polymer also enhance the transfection efficacy [70]. Similarly, TAT-conjugated methoxy poly (ethylene glycol) (MPEG)/poly (epsilon-caprolactone) (PCL) copolymers with disulfide linkage [71], showed a high-efficiency transfection and low cytotoxicity when the nanoparticles range from 100 to 200 nm encapsulated plasmid DNA or siRNA.
Many polymeric nanoparticles are designed for efficient delivery, and the CPPs-modified particles showed a trend of increased particle size and a looser complex [72]. However, TAT-modified nanoparticles exhibited the reversed result, as the polymer size decreased after modifying with TAT [73]. In conclusion, the crosslink between CPPs and the polymer is the key parameter for efficient CPPs modification; the valid linker MPEG-PCL disulfide couplings between TAT and polymer can significantly improve the delivery efficiency [74].

4. Combination of CPPs with Cationic Peptide

The cationic peptide could intensively bind to the cargos with negative charge such as nucleic acids. When CPPs are conjugated to cationic peptides in a covalent binding manner, they can deliver the cationic peptide with desired cargos to certain cells or tissues. The CPPs are also short stranded cationic peptides. TAT, a branch of CPPs, has been developed as a vector to deliver DNA and resulted in a four- to eight-fold higher gene expression when compared to Polyethylenimine (PEI), another popular transfection reagent [75].
Results other than those mentioned above could be attributed to the structure of the CPPs-peptide complex. Positive charge on the CPPs binds to the negative charge of nucleic acids, a behavior that is an important factor for nucleic acid delivery. A siRNA delivery experiment showed that CPPs with cationic peptides for siRNA delivery was significantly dependent on the types of CPPs and the length of the cationic peptides [76]. The experiment reported that the siRNA was bound to the MAP (model amphipathic peptide) based 21-mer oligolysine (K21-PDP), and the R6 or MAP CPPs was conjugated to the complex of siRNA-K21-PDP. The results showed that MAP based siRNA delivery efficiency was 170 and 600 folds greater at 1 h and 6 h post-transfection, respectively, when compared to the CPPs-based R6-K21-PDP vector [77]. As a result, the MAP-attached complex induced the comparable GFP silencing effects with the GFP siRNA transferred by the Lipofectamine 2000, which is a commercial siRNA delivery vector [77]. The consequence of this binding between nucleic acids and CPPs-modified cationic peptide led to the neutralization of positively charged CPPs, and the reduced or abrogated internalization efficiency of CPPs.
To solve the issue above, a new delivery method of nucleic acids emerged with the CPPs-modified dsRNA (double strand RNA) binding domain fusion protein (PTD-DRBD), which binds to the siRNA with a higher avidity [37,78,79], and therefore resulting in the PTD-mediated cellular uptake. The PTD-DRBD could deliver siRNA and induce the rapid RNA interference with high efficiency in various cells (e.g., the primary T cell, human umbilical vein endothelial cells, and human embryonic stem cells, etc.) with no cytotoxicity both in vitro and in vivo. The CPPs of TAT conjugated with U1A dsRBD as siRNA carrier (TAT-U1A dsRBD) [80] can also transfer siRNA into cells, but all the molecular complex of TAT-U1A dsRBD with siRNA are localized mainly in the endosome instead of cytoplasm.
Nanoparticles with a diameter under 100 nm have been employed in various drug deliveries and have emerged as another carrier. Like other delivery carriers, they also have some restrictions that need to be overcome for further utilization. These issues are summarized as side effects including low efficiency and crossing the BBB. To improve the overall performance and effectiveness of nanoparticles in the field of drug delivery, several approaches have been developed, including the surface-modification of nanoparticles by aggregating with CPPs. Meanwhile, the positive charges of the CPPs can bind to the negative charges of drug, siRNA, DNA, and shRNA; and also can be coated with virus to form nanoparticles (Figure 3) [81,82,83].
Figure 3. CPPs-based drug carriers. (A) Nanoparticles modified with CPPs; (B) poly CPPs as the vector; (C) The CPPs-loaded nanoparticles; (D) Nanoparticle formed by the bounding of CPPs with positive charge to drugs and cargoes with the negative charge; (E) Nanoparticles formed when CPP coated the carriers. b represents CPPs in the figure. b, colored in purple in above figure panel stands for CPPs; blue color represents nanoparticle in panel A, chemical links in panel B, Drugs in panel C, negative charge drugs (e.g., siRNA, etc.) in panel D, and other carriers (e.g., viral vectors, etc.) in panel E.
Figure 3. CPPs-based drug carriers. (A) Nanoparticles modified with CPPs; (B) poly CPPs as the vector; (C) The CPPs-loaded nanoparticles; (D) Nanoparticle formed by the bounding of CPPs with positive charge to drugs and cargoes with the negative charge; (E) Nanoparticles formed when CPP coated the carriers. b represents CPPs in the figure. b, colored in purple in above figure panel stands for CPPs; blue color represents nanoparticle in panel A, chemical links in panel B, Drugs in panel C, negative charge drugs (e.g., siRNA, etc.) in panel D, and other carriers (e.g., viral vectors, etc.) in panel E.
Ijms 16 19518 g003
Nanoparticles attached with CPPs improve the efficiency of drug delivery when compared to the non-modified CPPs-based nanoparticles [84,85]. For example, the LMWP (CPPs of low molecular weight protamine)-nanoparticles exhibited significant increased cellular accumulation when compared to the unmodified nanoparticles [86]. Gold nanoparticles modified with CPPs also enhance the cellular uptake [87]. In addition, the magnetic nanoparticles (MNPs) coated with either polyArg, polylysine (pLys) and PEI were evaluated for siRNA delivery; consequently, gene silencing was obtained from a brain cancer cell line C6, breast cancer cell line MCF7, and prostate cancer cell line TC2. However, it was reported that nanoparticle-pArg-siRNA can deliver a remarkably higher quantity of siRNA than that of nanoparticle-pLys-siRNA, furthermore, nanoparticle-PEI-siRNA comes with a higher cell viability when doing transfection [88]. Similarly, another study showed that the target gene expression could be increased 100 fold when using CPPs-based nanoparticles for DNA delivery as compared to the mutant TAT2-M1 that has no activity; and it was still four to eight times higher when compared to PEI, a popular cationic polymer for transfection [75].
In addition to the enhancement of cellular uptake, the CPPs modification can also improve the endosomes escape [89]. Since CPPs have been reported in a large scope to improve endosomal escape [47], nanoparticles in cooperation with octaarginine provide an ideal carrier system for delivering siRNA to the cells, and consequently induce a specific knock-down of targeted genes [90,91]. Thus, RNA interference is a promising strategy for gene therapy.
The number of CPPs on the nanoparticle surface is a pivotal property for efficient delivery in surface-modified nanoparticles. The uptake of nanoparticles is nonlinearly in accordance with the increase of CPPs amounts on the nanoparticle surface, and the delivery efficiency is 100-fold higher at the condition of 15 CPP (TAT) per cross-linked iron oxide (CLIO) particle [92]. In addition, the release rate from cells is also dependent on the amount of CPPs loading [93]. Meanwhile, the preparation methods of CPPs-based nanoparticles also affect the overall efficiency of drug delivery. It is evident that the nanoparticles FA-PC/R8-PC/pDNA-M and FA-PC/R8-PC/pDNA-L undergoing different preparation processes with an addition of CPPs formed different structures of nanoparticles, and thus showing different delivery efficiencies [94].
Another interesting application of nanoparticles conjugated with CPPs is that it can cross BBB and deliver drug components, while nanoparticles without TAT are unable to cross the BBB [84]. Nanoparticle size also plays an important role in its degree of permeability. For example, PEGylated gold nanoparticles can be located in the nucleus when the particle size is 2.4 nm as opposed to delivered the cytoplasm and the membrane when the particle size is 5.5 and 8.2 nm, respectively. Meanwhile, the nanoparticles cannot enter into the cells and are located at the cellular periphery when the particle size is 16 nm or larger [95]. In conclusion, the smaller the nanoparticle size, the closer to the nucleus it travels.

5. Binding of CPPs with the Viral Carrier

Adenovirus (Ad) is one of the viral vectors most widely used for gene delivery; they have been widely used in gene therapy studies due to their high transduction efficiency [96,97]. However, their serious shortcomings including immunogenicity, promiscuous tropism, and the inability to efficiently infect certain types of cells limited their translational utility. Briefly, Ad vectors based delivery is dependent on the cell surface receptor such as the coxsackievirus and adenovirus receptor (CAR), and has difficult delivering genes to the cells that lack CAR expression [98,99].
To improve the ability of an Ad-based vector to efficiently transform cells that lack the CAR, the peptide TAT-G-C [100] was chemically conjugated to the Ad vector with the cross linker 6-maleimidohexanoic acid N-hydroxysuccinimide ester. The conjugation of Ad vector demonstrated that it has a significantly enhanced efficiency of gene delivery in B16BL6 cells when compared to the wild type adenovirus vector. A similar experiment was also applied to the TAT related peptide. When G-R5-Q-R3-P2-Q-G-C was bound to the adenovirus vector, it showed an enhanced efficiency of the gene delivery in the B16BL6 cells, 50-fold greater than that of the wild type adenovirus vector [101]. Recently, the Ramsey group produced the vector by using PEGylating Ad, which packages a lacZ reporter gene, and then conjugating CPPs to form CPP-PEG-Ad particles. Meanwhile, these studies compared the effectiveness of four different CPPs (Pen, Tat, Pep1, and pArg). The results showed that CPP-PEG-Ad particles transduced the lack of CAR cells significantly better than unmodified Ad; Pen, the most effective CPP, produced an 80-fold improvement in transduction compared to the unmodified virus [102].
The CPPs can also be gene recombined into the adenovirus and produce the CPPs-modified Ad vectors. The gene recombined Ad vectors AVV- TAT (HI)-L2 and AVV-TAT(C)-L2 including a TAT improved the gene delivery efficiency at approximately 50 to 500 folds higher gene expression when compared to the wild type vector [103]. Finally, the recombined Ad vectors with CPPs dramatically increased gene transduction efficiency in the lack of CAR cells when compared to the wild type vector [104]. Addition of CPPs to virus vector can deliver genes to the cells that lack CAR expression.

6. Calcium and CPPs

Calcium is an important factor for cell signal transport and can enhance cellular uptake of CPPs as the cargo carrier [105,106,107]. For instance, the HeLa cellular uptake of TAT increased 44 folds when 6 mM Ca2+ was added into the culture medium [106], the CPPs dimer form with siRNA can be condensed, and consequently improve the cellular uptake of siRNA delivery [108,109,110].
Although the mechanism of Ca2+ as a delivery enhancer has not yet been clearly proven [111,112], Ca2+ is mainly involved in endocytosis [111], and the disruption of endosomes [113,114] improves the transfection and formation of nanoparticles with calcium [108,115]. Notably, the CPPs-based drug transferring induced the influx calcium [18,116]. Upon the addition of calcium, the “loose” complexes of siRNA/CPP can be condensed into small nanoparticles and thus, offers high efficiency and low toxicity gene silencing.

7. Conclusions

The application of CPPs is one of the most attractive approaches for drug delivery, as summarized in Figure 4. The cellular uptake of the carrier system, including liposome, polymer, cationic peptide, nanoparticles, and AVV are dramatically enhanced without significant cytotoxicity after modifying with CPPs. Moreover, the calcium can be further applied to enhance the cellular uptake of CPPs.
Figure 4. The potential use of CPPs. CPPs has been widely used as modulators or vehicles for the delivery of various biological products into cells. In order to overcome the neutralization and aggregation of CPPs that result in the inability to cross the cell membrane, calcium can be used to improve the nucleic acid delivery. The nucleic acids could be assembled with the typical drug carrier (e.g., liposome, cationic peptide, polymer, viral carrier and nanoparticle), and then be further modified with CPPs, consequently dramatically enhancing the nucleic acid transferring efficiency.
Figure 4. The potential use of CPPs. CPPs has been widely used as modulators or vehicles for the delivery of various biological products into cells. In order to overcome the neutralization and aggregation of CPPs that result in the inability to cross the cell membrane, calcium can be used to improve the nucleic acid delivery. The nucleic acids could be assembled with the typical drug carrier (e.g., liposome, cationic peptide, polymer, viral carrier and nanoparticle), and then be further modified with CPPs, consequently dramatically enhancing the nucleic acid transferring efficiency.
Ijms 16 19518 g004

Acknowledgments

This work was supported from the funding of the Scientific Research Foundation for the Returned Overseas Chinese Scholars, Ministry of Education, Huzhou Municipal Bureau of Science & Technology (2014GY08), and Zhejiang Provincial Natural Science Foundation (LY14H160015), China.

Author Contributions

Hua Li, Tung Yu Tsui wrote the manuscript, Wenxue Ma wrote and edited it.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AbbreviationFull Name
ArgArginine
AspAspartic acid
AVVAdeno-associated virus
BBBBlood brain barrier
CCysteine
CARCoxsackievirus and adenovirus receptor
CBA-DAHCystamine bisacrylamide-diaminohexane
CNSCentral nervous system diseases
CPPCell penetrating peptide
dsRBD/DRBDDoubled RNA binding domain
GGlycine
GluGlutamic acid
HHistidine
HA2hemagglutinin peptide
HIVHuman immunodeficiency virus
LLeucine
LMWPCell penetrating peptide of LMWP
Lys/KLysine
MAPModel amphipathic peptide
MNPsMagnetic nanoparticles
MPEGMethoxy poly (ethylene glycol)
PProline
PCLPoly(epsilon-caprolactone)
PEGPolyethylenglycol
PLLPoly-L-Lysine
pLysPolylysine
PTDProtein transduction domain
PTD-DRBDCPP of PTD modified dsRNA binding domain
QGlutamine
RArginine
R8Polyarginine8
RNAiRNA interference
shRNAShort hairpin RNA
siRNASmall interfering RNA
TATCell penetrating peptide of TAT
TAT-PEG-MPLsTAT-conjugated PEGlated Magnetic polymeric liposome
TAT-U1A dsRBDTAT conjugated with U1A dsRBD
TP 10Transportan10
U1AU1 small nuclear ribonucleoprotein A
VValine

References

  1. Eguchi, A.; Akuta, T.; Okuyama, H.; Senda, T.; Yokoi, H.; Inokuchi, H.; Fujita, S.; Hayakawa, T.; Takeda, K.; Hasegawa, M.; et al. Protein transduction domain of HIV-1 Tat protein promotes efficient delivery of DNA into mammalian cells. J. Biol. Chem. 2001, 276, 26204–26210. [Google Scholar] [CrossRef] [PubMed]
  2. Lindgren, M.; Hallbrink, M.; Prochiantz, A.; Langel, U. Cell-penetrating peptides. Trends Pharmacol. Sci. 2000, 21, 99–103. [Google Scholar] [CrossRef]
  3. Derossi, D.; Chassaing, G.; Prochiantz, A. Trojan peptides: The penetratin system for intracellular delivery. Trends Cell Biol. 1998, 8, 84–87. [Google Scholar] [CrossRef]
  4. 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]
  5. De Coupade, C.; Fittipaldi, A.; Chagnas, V.; Michel, M.; Carlier, S.; Tasciotti, E.; Darmon, A.; Ravel, D.; Kearsey, J.; Giacca, M.; et al. Novel human-derived cell-penetrating peptides for specific subcellular delivery of therapeutic biomolecules. Biochem. J. 2005, 390, 407–418. [Google Scholar] [CrossRef] [PubMed]
  6. Gump, J.M.; Dowdy, S.F. TAT transduction: The molecular mechanism and therapeutic prospects. Trends Mol. Med. 2007, 13, 443–448. [Google Scholar] [CrossRef] [PubMed]
  7. Ter-Avetisyan, G.; Tunnemann, G.; Nowak, D.; Nitschke, M.; Herrmann, A.; Drab, M.; Cardoso, M.C. Cell entry of arginine-rich peptides is independent of endocytosis. J. Biol. Chem. 2009, 284, 3370–3378. [Google Scholar] [CrossRef] [PubMed]
  8. Mishra, A.; Lai, G.H.; Schmidt, N.W.; Sun, V.Z.; Rodriguez, A.R.; Tong, R.; Tang, L.; Cheng, J.; Deming, T.J.; Kamei, D.T.; et al. Translocation of HIV TAT peptide and analogues induced by multiplexed membrane and cytoskeletal interactions. Proc. Natl. Acad. Sci. USA 2011, 108, 16883–16888. [Google Scholar] [CrossRef] [PubMed]
  9. Li, G.H.; Li, W.; Mumper, R.J.; Nath, A. Molecular mechanisms in the dramatic enhancement of HIV-1 Tat transduction by cationic liposomes. FASEB J. 2012, 26, 2824–2834. [Google Scholar] [CrossRef] [PubMed]
  10. Brasseur, R.; Divita, G. Happy birthday cell penetrating peptides: Already 20 years. Biochim. Biophys. Acta 2010, 1798, 2177–2181. [Google Scholar] [CrossRef] [PubMed]
  11. Morris, M.C.; Deshayes, S.; Heitz, F.; Divita, G. Cell-penetrating peptides: From molecular mechanisms to therapeutics. Biol. Cell 2008, 100, 201–217. [Google Scholar] [CrossRef] [PubMed]
  12. Heitz, F.; Morris, M.C.; Divita, G. Twenty years of cell-penetrating peptides: From molecular mechanisms to therapeutics. Br. J. Pharmacol. 2009, 157, 195–206. [Google Scholar] [CrossRef] [PubMed]
  13. Frankel, A.D.; Pabo, C.O. Cellular uptake of the tat protein from human immunodeficiency virus. Cell 1988, 55, 1189–1193. [Google Scholar] [CrossRef]
  14. Gautam, A.; Singh, H.; Tyagi, A.; Chaudhary, K.; Kumar, R.; Kapoor, P.; Raghava, G.P. CPPsite: A curated database of cell penetrating peptides. Database 2012, 2012, bas015. [Google Scholar] [CrossRef] [PubMed]
  15. Richard, J.P.; Melikov, K.; Vives, E.; Ramos, C.; Verbeure, B.; Gait, M.J.; Chernomordik, L.V.; Lebleu, B. Cell-penetrating peptides. A reevaluation of the mechanism of cellular uptake. J. Biol. Chem. 2003, 278, 585–590. [Google Scholar] [CrossRef] [PubMed]
  16. Fotin-Mleczek, M.; Fischer, R.; Brock, R. Endocytosis and cationic cell-penetrating peptides—A merger of concepts and methods. Curr. Pharm. Des. 2005, 11, 3613–3628. [Google Scholar] [CrossRef] [PubMed]
  17. Jones, A.T.; Sayers, E.J. Cell entry of cell penetrating peptides: Tales of tails wagging dogs. J. Control. Release 2012, 161, 582–591. [Google Scholar] [CrossRef] [PubMed]
  18. Lorents, A.; Kodavali, P.K.; Oskolkov, N.; Langel, U.; Hallbrink, M.; Pooga, M. Cell-penetrating peptides split into two groups based on modulation of intracellular calcium concentration. J. Biol. Chem. 2012, 287, 16880–16889. [Google Scholar] [CrossRef] [PubMed]
  19. Furuhata, M.; Kawakami, H.; Toma, K.; Hattori, Y.; Maitani, Y. Intracellular delivery of proteins in complexes with oligoarginine-modified liposomes and the effect of oligoarginine length. Bioconj. Chem. 2006, 17, 935–942. [Google Scholar] [CrossRef] [PubMed]
  20. Takayama, K.; Tadokoro, A.; Pujals, S.; Nakase, I.; Giralt, E.; Futaki, S. Novel system to achieve one-pot modification of cargo molecules with oligoarginine vectors for intracellular delivery. Bioconj. Chem. 2009, 20, 249–257. [Google Scholar] [CrossRef] [PubMed]
  21. Zaidi, N.; Burster, T.; Sommandas, V.; Herrmann, T.; Boehm, B.O.; Driessen, C.; Voelter, W.; Kalbacher, H. A novel cell penetrating aspartic protease inhibitor blocks processing and presentation of tetanus toxoid more efficiently than pepstatin A. Biochem. Biophys. Res. Commun. 2007, 364, 243–249. [Google Scholar] [CrossRef] [PubMed]
  22. Ogawa, T.; Ono, S.; Ichikawa, T.; Arimitsu, S.; Onoda, K.; Tokunaga, K.; Sugiu, K.; Tomizawa, K.; Matsui, H.; Date, I. Novel protein transduction method by using 11R: An effective new drug delivery system for the treatment of cerebrovascular diseases. Stroke 2007, 38, 1354–1361. [Google Scholar] [CrossRef] [PubMed]
  23. Foged, C.; Franzyk, H.; Bahrami, S.; Frokjaer, S.; Jaroszewski, J.W.; Nielsen, H.M.; Olsen, C.A. Cellular uptake and membrane-destabilising properties of alpha-peptide/beta-peptoid chimeras: Lessons for the design of new cell-penetrating peptides. Biochim. Biophys. Acta 2008, 1778, 2487–2495. [Google Scholar] [CrossRef] [PubMed]
  24. Sun, J.; Yan, Y.; Wang, X.T.; Liu, X.W.; Peng, D.J.; Wang, M.; Tian, J.; Zong, Y.Q.; Zhang, Y.H.; Noteborn, M.H.; et al. PTD4-apoptin protein therapy inhibits tumor growth in vivo. Int. J. Cancer 2009, 124, 2973–2981. [Google Scholar] [CrossRef] [PubMed]
  25. Mai, J.C.; Mi, Z.; Kim, S.H.; Ng, B.; Robbins, P.D. A proapoptotic peptide for the treatment of solid tumors. Cancer Res. 2001, 61, 7709–7712. [Google Scholar] [PubMed]
  26. Arrighi, R.B.; Ebikeme, C.; Jiang, Y.; Ranford-Cartwright, L.; Barrett, M.P.; Langel, U.; Faye, I. Cell-penetrating peptide TP10 shows broad-spectrum activity against both Plasmodium falciparum and Trypanosoma brucei brucei. Antimicrob. Agents Chemother. 2008, 52, 3414–3417. [Google Scholar] [CrossRef] [PubMed]
  27. El-Andaloussi, S.; Johansson, H.J.; Holm, T.; Langel, U. A novel cell-penetrating peptide, M918, for efficient delivery of proteins and peptide nucleic acids. Mol. Ther. 2007, 15, 1820–1826. [Google Scholar] [CrossRef] [PubMed]
  28. Aussedat, B.; Sagan, S.; Chassaing, G.; Bolbach, G.; Burlina, F. Quantification of the efficiency of cargo delivery by peptidic and pseudo-peptidic Trojan carriers using MALDI-TOF mass spectrometry. Biochim. Biophys. Acta 2006, 1758, 375–383. [Google Scholar] [CrossRef] [PubMed]
  29. Choi, J.M.; Ahn, M.H.; Chae, W.J.; Jung, Y.G.; Park, J.C.; Song, H.M.; Kim, Y.E.; Shin, J.A.; Park, C.S.; Park, J.W.; et al. Intranasal delivery of the cytoplasmic domain of CTLA-4 using a novel protein transduction domain prevents allergic inflammation. Nat. Med. 2006, 12, 574–579. [Google Scholar] [CrossRef] [PubMed]
  30. Kameyama, S.; Horie, M.; Kikuchi, T.; Omura, T.; Takeuchi, T.; Nakase, I.; Sugiura, Y.; Futaki, S. Effects of cell-permeating peptide binding on the distribution of 125I-labeled Fab fragment in rats. Bioconj. Chem. 2006, 17, 597–602. [Google Scholar] [CrossRef] [PubMed]
  31. Johnson, L.N.; Cashman, S.M.; Read, S.P.; Kumar-Singh, R. Cell penetrating peptide POD mediates delivery of recombinant proteins to retina, cornea and skin. Vis. Res. 2010, 50, 686–697. [Google Scholar] [CrossRef] [PubMed]
  32. Mano, M.; Teodosio, C.; Paiva, A.; Simoes, S.; Pedroso de Lima, M.C. On the mechanisms of the internalization of S4(13)-PV cell-penetrating peptide. Biochem. J. 2005, 390, 603–612. [Google Scholar] [CrossRef] [PubMed]
  33. Sheng, J.; Oyler, G.; Zhou, B.; Janda, K.; Shoemaker, C.B. Identification and characterization of a novel cell-penetrating peptide. Biochem. Biophys. Res. Commun. 2009, 382, 236–240. [Google Scholar] [CrossRef] [PubMed]
  34. Elmquist, A.; Hansen, M.; Langel, U. Structure-activity relationship study of the cell-penetrating peptide pVEC. Biochim. Biophys. Acta 2006, 1758, 721–729. [Google Scholar] [CrossRef] [PubMed]
  35. Jiang, Y.; Li, M.; Zhang, Z.; Gong, T.; Sun, X. Cell-penetrating peptides as delivery enhancers for vaccine. Curr. Pharm. Biotechnol. 2014, 15, 256–266. [Google Scholar] [CrossRef] [PubMed]
  36. Funhoff, A.M.; van Nostrum, C.F.; Lok, M.C.; Fretz, M.M.; Crommelin, D.J.; Hennink, W.E. Poly(3-guanidinopropyl methacrylate): A novel cationic polymer for gene delivery. Bioconj. Chem. 2004, 15, 1212–1220. [Google Scholar] [CrossRef] [PubMed]
  37. Li, H.; Tsui, T. Six-cell penetrating peptide-based fusion proteins for siRNA delivery. Drug Deliv. 2015, 22, 436–443. [Google Scholar] [CrossRef] [PubMed]
  38. Jo, J.; Hong, S.; Choi, W.Y.; Lee, D.R. Cell-penetrating peptide (CPP)-conjugated proteins is an efficient tool for manipulation of human mesenchymal stromal cells. Sci. Rep. 2014, 4, 4378. [Google Scholar] [CrossRef] [PubMed]
  39. 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] [PubMed]
  40. Cardoso, A.M.; Trabulo, S.; Cardoso, A.L.; Maia, S.; Gomes, P.; Jurado, A.S.; Pedroso de Lima, M.C. Comparison of the efficiency of complexes based on S4(13)-PV cell-penetrating peptides in plasmid DNA and siRNA delivery. Mol. Pharm. 2013, 10, 2653–2666. [Google Scholar] [CrossRef] [PubMed]
  41. Li, H.; Zheng, X.; Koren, V.; Vashist, Y.K.; Tsui, T.Y. Highly efficient delivery of siRNA to a heart transplant model by a novel cell penetrating peptide-dsRNA binding domain. Int. J. Pharm. 2014, 469, 206–213. [Google Scholar] [CrossRef] [PubMed]
  42. Farkhani, S.M.; Valizadeh, A.; Karami, H.; Mohammadi, S.; Sohrabi, N.; Badrzadeh, F. Cell penetrating peptides: Efficient vectors for delivery of nanoparticles, nanocarriers, therapeutic and diagnostic molecules. Peptides 2014, 57, 78–94. [Google Scholar] [CrossRef] [PubMed]
  43. Sharma, G.; Modgil, A.; Zhong, T.; Sun, C.; Singh, J. Influence of short-chain cell-penetrating peptides on transport of doxorubicin encapsulating receptor-targeted liposomes across brain endothelial barrier. Pharm. Res. 2014, 31, 1194–1209. [Google Scholar] [CrossRef] [PubMed]
  44. Sharma, G.; Modgil, A.; Layek, B.; Arora, K.; Sun, C.; Law, B.; Singh, J. Cell penetrating peptide tethered bi-ligand liposomes for delivery to brain in vivo: Biodistribution and transfection. J. Control. Release: Off. J. Control. Release Soc. 2013, 167, 1–10. [Google Scholar] [CrossRef] [PubMed]
  45. Zhou, J.; Liu, W.; Pong, R.C.; Hao, G.; Sun, X.; Hsieh, J.T. Analysis of oligo-arginine cell-permeable peptides uptake by prostate cells. Amino Acids 2012, 42, 1253–1260. [Google Scholar] [CrossRef] [PubMed]
  46. Patel, L.N.; Wang, J.; Kim, K.J.; Borok, Z.; Crandall, E.D.; Shen, W.C. Conjugation with cationic cell-penetrating peptide increases pulmonary absorption of insulin. Mol. Pharm. 2009, 6, 492–503. [Google Scholar] [CrossRef] [PubMed]
  47. Endoh, T.; Ohtsuki, T. Cellular siRNA delivery using cell-penetrating peptides modified for endosomal escape. Adv. Drug Deliv. Rev. 2009, 61, 704–709. [Google Scholar] [CrossRef] [PubMed]
  48. Meade, B.R.; Dowdy, S.F. Enhancing the cellular uptake of siRNA duplexes following noncovalent packaging with protein transduction domain peptides. Adv. Drug Deliv. Rev. 2008, 60, 530–536. [Google Scholar] [CrossRef] [PubMed]
  49. Turner, J.J.; Jones, S.; Fabani, M.M.; Ivanova, G.; Arzumanov, A.A.; Gait, M.J. RNA targeting with peptide conjugates of oligonucleotides, siRNA and PNA. Blood Cells Mol. Dis. 2007, 38, 1–7. [Google Scholar] [CrossRef] [PubMed]
  50. Torchilin, V.P. Recent advances with liposomes as pharmaceutical carriers. Nat. Rev. Drug Discov. 2005, 4, 145–160. [Google Scholar] [CrossRef] [PubMed]
  51. Zhang, Q.; Tang, J.; Fu, L.; Ran, R.; Liu, Y.; Yuan, M.; He, Q. A pH-responsive alpha-helical cell penetrating peptide-mediated liposomal delivery system. Biomaterials 2013, 34, 7980–7993. [Google Scholar] [CrossRef] [PubMed]
  52. Marty, C.; Meylan, C.; Schott, H.; Ballmer-Hofer, K.; Schwendener, R.A. Enhanced heparan sulfate proteoglycan-mediated uptake of cell-penetrating peptide-modified liposomes. Cell. Mol. Life Sci. 2004, 61, 1785–1794. [Google Scholar] [CrossRef] [PubMed]
  53. Fretz, M.M.; Koning, G.A.; Mastrobattista, E.; Jiskoot, W.; Storm, G. OVCAR-3 cells internalize TAT-peptide modified liposomes by endocytosis. Biochim. Biophys. Acta 2004, 1665, 48–56. [Google Scholar] [CrossRef] [PubMed]
  54. Cryan, S.A.; Devocelle, M.; Moran, P.J.; Hickey, A.J.; Kelly, J.G. Increased intracellular targeting to airway cells using octaarginine-coated liposomes: In vitro assessment of their suitability for inhalation. Mol. Pharm. 2006, 3, 104–112. [Google Scholar] [CrossRef] [PubMed]
  55. Zhang, C.; Tang, N.; Liu, X.; Liang, W.; Xu, W.; Torchilin, V.P. siRNA-containing liposomes modified with polyarginine effectively silence the targeted gene. J. Control. Release 2006, 112, 229–239. [Google Scholar] [CrossRef] [PubMed]
  56. Nakamura, T.; Moriguchi, R.; Kogure, K.; Shastri, N.; Harashima, H. Efficient MHC class I presentation by controlled intracellular trafficking of antigens in octaarginine-modified liposomes. Mol. Ther. 2008, 16, 1507–1514. [Google Scholar] [CrossRef] [PubMed]
  57. Torchilin, V.P.; Rammohan, R.; Weissig, V.; Levchenko, T.S. TAT peptide on the surface of liposomes affords their efficient intracellular delivery even at low temperature and in the presence of metabolic inhibitors. Proc. Natl. Acad. Sci. USA 2001, 98, 8786–8791. [Google Scholar] [CrossRef] [PubMed]
  58. Saw, P.E.; Ko, Y.T.; Jon, S. Efficient Liposomal Nanocarrier-mediated Oligodeoxynucleotide Delivery Involving Dual Use of a Cell-Penetrating Peptide as a Packaging and Intracellular Delivery Agent. Macromol. Rapid Commun. 2010, 31, 1155–1162. [Google Scholar] [CrossRef] [PubMed]
  59. Koren, E.; Apte, A.; Jani, A.; Torchilin, V.P. Multifunctional PEGylated 2C5-immunoliposomes containing pH-sensitive bonds and TAT peptide for enhanced tumor cell internalization and cytotoxicity. J. Control. Release 2012, 160, 264–273. [Google Scholar] [CrossRef] [PubMed]
  60. Qin, Y.; Chen, H.; Yuan, W.; Kuai, R.; Zhang, Q.; Xie, F.; Zhang, L.; Zhang, Z.; Liu, J.; He, Q. Liposome formulated with TAT-modified cholesterol for enhancing the brain delivery. Int. J. Pharm. 2011, 419, 85–95. [Google Scholar] [CrossRef] [PubMed]
  61. Qin, Y.; Chen, H.; Zhang, Q.; Wang, X.; Yuan, W.; Kuai, R.; Tang, J.; Zhang, L.; Zhang, Z.; Zhang, Q.; et al. Liposome formulated with TAT-modified cholesterol for improving brain delivery and therapeutic efficacy on brain glioma in animals. Int. J. Pharm. 2011, 420, 304–312. [Google Scholar] [CrossRef] [PubMed]
  62. Wang, H.; Zhang, S.; Liao, Z.; Wang, C.; Liu, Y.; Feng, S.; Jiang, X.; Chang, J. PEGlated magnetic polymeric liposome anchored with TAT for delivery of drugs across the blood-spinal cord barrier. Biomaterials 2010, 31, 6589–6596. [Google Scholar] [CrossRef] [PubMed]
  63. Crombez, L.; Divita, G. A non-covalent peptide-based strategy for siRNA delivery. Methods Mol. Biol. 2011, 683, 349–360. [Google Scholar] [PubMed]
  64. Walker, L.; Perkins, E.; Kratz, F.; Raucher, D. Cell penetrating peptides fused to a thermally targeted biopolymer drug carrier improve the delivery and antitumor efficacy of an acid-sensitive doxorubicin derivative. Int. J. Pharm. 2012, 436, 825–832. [Google Scholar] [CrossRef] [PubMed]
  65. Xia, H.; Gao, X.; Gu, G.; Liu, Z.; Hu, Q.; Tu, Y.; Song, Q.; Yao, L.; Pang, Z.; Jiang, X.; et al. Penetratin-functionalized PEG-PLA nanoparticles for brain drug delivery. Int. J. Pharm. 2012, 436, 840–850. [Google Scholar] [CrossRef] [PubMed]
  66. Kanazawa, T.; Taki, H.; Tanaka, K.; Takashima, Y.; Okada, H. Cell-penetrating peptide-modified block copolymer micelles promote direct brain delivery via intranasal administration. Pharm. Res. 2011, 28, 2130–2139. [Google Scholar] [CrossRef] [PubMed]
  67. Nam, H.Y.; Kim, J.; Kim, S.; Yockman, J.W.; Kim, S.W.; Bull, D.A. Cell penetrating peptide conjugated bioreducible polymer for siRNA delivery. Biomaterials 2011, 32, 5213–5222. [Google Scholar] [CrossRef] [PubMed]
  68. Sakuma, S.; Suita, M.; Yamamoto, T.; Masaoka, Y.; Kataoka, M.; Yamashita, S.; Nakajima, N.; Shinkai, N.; Yamauchi, H.; Hiwatari, K.; et al. Performance of cell-penetrating peptide-linked polymers physically mixed with poorly membrane-permeable molecules on cell membranes. Eur. J. Pharm. Biopharm. 2012, 81, 64–73. [Google Scholar] [CrossRef] [PubMed]
  69. Hashida, H.; Miyamoto, M.; Cho, Y.; Hida, Y.; Kato, K.; Kurokawa, T.; Okushiba, S.; Kondo, S.; Dosaka-Akita, H.; Katoh, H. Fusion of HIV-1 Tat protein transduction domain to poly-lysine as a new DNA delivery tool. Br. J. Cancer 2004, 90, 1252–1258. [Google Scholar] [CrossRef] [PubMed]
  70. Sakuma, S.; Suita, M.; Masaoka, Y.; Kataoka, M.; Nakajima, N.; Shinkai, N.; Yamauchi, H.; Hiwatari, K.; Tachikawa, H.; Kimura, R.; et al. Oligoarginine-linked polymers as a new class of penetration enhancers. J. Control. Release 2010, 148, 187–196. [Google Scholar] [CrossRef] [PubMed]
  71. Kanazawa, T.; Sugawara, K.; Tanaka, K.; Horiuchi, S.; Takashima, Y.; Okada, H. Suppression of tumor growth by systemic delivery of anti-VEGF siRNA with cell-penetrating peptide-modified MPEG-PCL nanomicelles. Eur. J. Pharm. Biopharm. 2012, 81, 470–477. [Google Scholar] [CrossRef] [PubMed]
  72. Lee, J.Y.; Bae, K.H.; Kim, J.S.; Nam, Y.S.; Park, T.G. Intracellular delivery of paclitaxel using oil-free, shell cross-linked HSA--multi-armed PEG nanocapsules. Biomaterials 2011, 32, 8635–8644. [Google Scholar] [CrossRef] [PubMed]
  73. Shrestha, R.; Shen, Y.; Pollack, K.A.; Taylor, J.S.; Wooley, K.L. Dual peptide nucleic acid- and peptide-functionalized shell cross-linked nanoparticles designed to target mRNA toward the diagnosis and treatment of acute lung injury. Bioconj. Chem. 2012, 23, 574–585. [Google Scholar] [CrossRef] [PubMed]
  74. Tanaka, K.; Kanazawa, T.; Shibata, Y.; Suda, Y.; Fukuda, T.; Takashima, Y.; Okada, H. Development of cell-penetrating peptide-modified MPEG-PCL diblock copolymeric nanoparticles for systemic gene delivery. Int. J. Pharm. 2010, 396, 229–238. [Google Scholar] [CrossRef] [PubMed]
  75. Rudolph, C.; Schillinger, U.; Ortiz, A.; Tabatt, K.; Plank, C.; Muller, R.H.; Rosenecker, J. Application of novel solid lipid nanoparticle (SLN)-gene vector formulations based on a dimeric HIV-1 TAT-peptide in vitro and in vivo. Pharm. Res. 2004, 21, 1662–1669. [Google Scholar] [CrossRef] [PubMed]
  76. Ishihara, T.; Goto, M.; Kodera, K.; Kanazawa, H.; Murakami, Y.; Mizushima, Y.; Higaki, M. Intracellular delivery of siRNA by cell-penetrating peptides modified with cationic oligopeptides. Drug Deliv. 2009, 16, 153–159. [Google Scholar] [CrossRef] [PubMed]
  77. Mo, R.H.; Zaro, J.L.; Shen, W.C. Comparison of cationic and amphipathic cell penetrating peptides for siRNA delivery and efficacy. Mol. Pharm. 2012, 9, 299–309. [Google Scholar] [CrossRef] [PubMed]
  78. Palm-Apergi, C.; Eguchi, A.; Dowdy, S.F. PTD-DRBD siRNA delivery. Methods Mol. Biol. 2011, 683, 339–347. [Google Scholar] [PubMed]
  79. Eguchi, A.; Meade, B.R.; Chang, Y.C.; Fredrickson, C.T.; Willert, K.; Puri, N.; Dowdy, S.F. Efficient siRNA delivery into primary cells by a peptide transduction domain-dsRNA binding domain fusion protein. Nat. Biotechnol. 2009, 27, 567–571. [Google Scholar] [CrossRef] [PubMed]
  80. Endoh, T.; Sisido, M.; Ohtsuki, T. Cellular siRNA delivery mediated by a cell-permeant RNA-binding protein and photoinduced RNA interference. Bioconj. Chem. 2008, 19, 1017–1024. [Google Scholar] [CrossRef] [PubMed]
  81. Zhang, P.; Cheetham, A.G.; Lin, Y.A.; Cui, H. Self-assembled Tat nanofibers as effective drug carrier and transporter. ACS Nano 2013, 7, 5965–5977. [Google Scholar] [CrossRef] [PubMed]
  82. Asai, T.; Tsuzuku, T.; Takahashi, S.; Okamoto, A.; Dewa, T.; Nango, M.; Hyodo, K.; Ishihara, H.; Kikuchi, H.; Oku, N. Cell-penetrating peptide-conjugated lipid nanoparticles for siRNA delivery. Biochem. Biophys. Res. Commun. 2014, 444, 599–604. [Google Scholar] [CrossRef] [PubMed]
  83. Kanazawa, T.; Akiyama, F.; Kakizaki, S.; Takashima, Y.; Seta, Y. Delivery of siRNA to the brain using a combination of nose-to-brain delivery and cell-penetrating peptide-modified nano-micelles. Biomaterials 2013, 34, 9220–9226. [Google Scholar] [CrossRef] [PubMed]
  84. Liu, L.; Guo, K.; Lu, J.; Venkatraman, S.S.; Luo, D.; Ng, K.C.; Ling, E.A.; Moochhala, S.; Yang, Y.Y. Biologically active core/shell nanoparticles self-assembled from cholesterol-terminated PEG-TAT for drug delivery across the blood-brain barrier. Biomaterials 2008, 29, 1509–1517. [Google Scholar] [CrossRef] [PubMed]
  85. Tiwari, P.M.; Eroglu, E.; Bawage, S.S.; Vig, K.; Miller, M.E.; Pillai, S.; Dennis, V.A.; Singh, S.R. Enhanced intracellular translocation and biodistribution of gold nanoparticles functionalized with a cell-penetrating peptide (VG-21) from vesicular stomatitis virus. Biomaterials 2014, 35, 9484–9494. [Google Scholar] [CrossRef] [PubMed]
  86. Xia, H.; Gao, X.; Gu, G.; Liu, Z.; Zeng, N.; Hu, Q.; Song, Q.; Yao, L.; Pang, Z.; Jiang, X.; et al. Low molecular weight protamine-functionalized nanoparticles for drug delivery to the brain after intranasal administration. Biomaterials 2011, 32, 9888–9898. [Google Scholar] [CrossRef] [PubMed]
  87. Susumu, K.; Oh, E.; Delehanty, J.B.; Blanco-Canosa, J.B.; Johnson, B.J.; Jain, V.; Hervey, W.J.T.; Algar, W.R.; Boeneman, K.; Dawson, P.E.; et al. Multifunctional compact zwitterionic ligands for preparing robust biocompatible semiconductor quantum dots and gold nanoparticles. J. Am. Chem. Soc. 2011, 133, 9480–9496. [Google Scholar] [CrossRef] [PubMed]
  88. Veiseh, O.; Kievit, F.M.; Mok, H.; Ayesh, J.; Clark, C.; Fang, C.; Leung, M.; Arami, H.; Park, J.O.; Zhang, M. Cell transcytosing poly-arginine coated magnetic nanovector for safe and effective siRNA delivery. Biomaterials 2011, 32, 5717–5725. [Google Scholar] [CrossRef] [PubMed]
  89. Nativo, P.; Prior, I.A.; Brust, M. Uptake and intracellular fate of surface-modified gold nanoparticles. ACS Nano 2008, 2, 1639–1644. [Google Scholar] [CrossRef] [PubMed]
  90. Nakamura, Y.; Kogure, K.; Futaki, S.; Harashima, H. Octaarginine-modified multifunctional envelope-type nano device for siRNA. J. Control. Release 2007, 119, 360–367. [Google Scholar] [CrossRef] [PubMed]
  91. Hayashi, Y.; Yamauchi, J.; Khalil, I.A.; Kajimoto, K.; Akita, H.; Harashima, H. Cell penetrating peptide-mediated systemic siRNA delivery to the liver. Int. J. Pharm. 2011, 419, 308–313. [Google Scholar] [CrossRef] [PubMed]
  92. Zhao, M.; Kircher, M.F.; Josephson, L.; Weissleder, R. Differential conjugation of tat peptide to superparamagnetic nanoparticles and its effect on cellular uptake. Bioconj. Chem. 2002, 13, 840–844. [Google Scholar] [CrossRef]
  93. Zhang, K.; Fang, H.; Chen, Z.; Taylor, J.S.; Wooley, K.L. Shape effects of nanoparticles conjugated with cell-penetrating peptides (HIV Tat PTD) on CHO cell uptake. Bioconj. Chem. 2008, 19, 1880–1887. [Google Scholar] [CrossRef] [PubMed]
  94. Jiang, Q.Y.; Lai, L.H.; Shen, J.; Wang, Q.Q.; Xu, F.J.; Tang, G.P. Gene delivery to tumor cells by cationic polymeric nanovectors coupled to folic acid and the cell-penetrating peptide octaarginine. Biomaterials 2011, 32, 7253–7262. [Google Scholar] [CrossRef] [PubMed]
  95. Oh, E.; Delehanty, J.B.; Sapsford, K.E.; Susumu, K.; Goswami, R.; Blanco-Canosa, J.B.; Dawson, P.E.; Granek, J.; Shoff, M.; Zhang, Q.; et al. Cellular uptake and fate of PEGylated gold nanoparticles is dependent on both cell-penetration peptides and particle size. ACS Nano 2011, 5, 6434–6448. [Google Scholar] [CrossRef] [PubMed]
  96. Benihoud, K.; Yeh, P.; Perricaudet, M. Adenovirus vectors for gene delivery. Curr. Opin. Biotechnol. 1999, 10, 440–447. [Google Scholar] [CrossRef]
  97. Kovesdi, I.; Brough, D.E.; Bruder, J.T.; Wickham, T.J. Adenoviral vectors for gene transfer. Curr. Opin. Biotechnol. 1997, 8, 583–589. [Google Scholar] [CrossRef]
  98. Wang, L.; Yao, B.; Li, Q.; Mei, K.; Xu, J.R.; Li, H.X.; Wang, Y.S.; Wen, Y.J.; Wang, X.D.; Yang, H.S.; et al. Gene therapy with recombinant adenovirus encoding endostatin encapsulated in cationic liposome in coxsackievirus and adenovirus receptor-deficient colon carcinoma murine models. Hum. Gene Ther. 2011, 22, 1061–1069. [Google Scholar] [CrossRef] [PubMed]
  99. Sharma, P.; Kolawole, A.O.; Wiltshire, S.M.; Frondorf, K.; Excoffon, K.J. Accessibility of the coxsackievirus and adenovirus receptor and its importance in adenovirus gene transduction efficiency. J. Gen. Virol. 2012, 93 Pt 1, 155–158. [Google Scholar] [CrossRef] [PubMed]
  100. Kida, S.; Eto, Y.; Maeda, M.; Yoshioka, Y.; Nakagawa, S.; Hojo, K.; Tsuda, Y.; Mayumia, T.; Kawasaki, K. Preparation of a Tat-related transporter peptide for carrying the adenovirus vector into cells. Protein Pept. Lett. 2008, 15, 219–222. [Google Scholar] [PubMed]
  101. Maeda, M.; Kida, S.; Hojo, K.; Eto, Y.; Gaob, J.Q.; Kurachi, S.; Sekiguchi, F.; Mizuguchi, H.; Hayakawa, T.; Mayumi, T.; et al. Design and synthesis of a peptide-PEG transporter tool for carrying adenovirus vector into cells. Bioorg. Med. Chem. Lett. 2005, 15, 621–624. [Google Scholar] [CrossRef] [PubMed]
  102. Nigatu, A.S.; Vupputuri, S.; Flynn, N.; Ramsey, J.D. Effects of cell-penetrating peptides on transduction efficiency of PEGylated adenovirus. Biomed. Pharmacother. 2015, 71, 153–160. [Google Scholar] [CrossRef] [PubMed]
  103. Kurachi, S.; Tashiro, K.; Sakurai, F.; Sakurai, H.; Kawabata, K.; Yayama, K.; Okamoto, H.; Nakagawa, S.; Mizuguchi, H. Fiber-modified adenovirus vectors containing the TAT peptide derived from HIV-1 in the fiber knob have efficient gene transfer activity. Gene Ther. 2007, 14, 1160–1165. [Google Scholar] [CrossRef] [PubMed]
  104. Yu, D.; Jin, C.; Leja, J.; Majdalani, N.; Nilsson, B.; Eriksson, F.; Essand, M. Adenovirus with hexon Tat-protein transduction domain modification exhibits increased therapeutic effect in experimental neuroblastoma and neuroendocrine tumors. J. Virol. 2011, 85, 13114–13123. [Google Scholar] [CrossRef] [PubMed]
  105. Bottger, M.; Zaitsev, S.V.; Otto, A.; Haberland, A.; Vorob’ev, V.I. Acid nuclear extracts as mediators of gene transfer and expression. Biochim. Biophys. Acta 1998, 1395, 78–87. [Google Scholar] [CrossRef]
  106. Shiraishi, T.; Pankratova, S.; Nielsen, P.E. Calcium ions effectively enhance the effect of antisense peptide nucleic acids conjugated to cationic tat and oligoarginine peptides. Chem. Biol. 2005, 12, 923–929. [Google Scholar] [CrossRef] [PubMed]
  107. Haberland, A.; Knaus, T.; Zaitsev, S.V.; Stahn, R.; Mistry, A.R.; Coutelle, C.; Haller, H.; Bottger, M. Calcium ions as efficient cofactor of polycation-mediated gene transfer. Biochim. Biophys. Acta 1999, 1445, 21–30. [Google Scholar] [CrossRef]
  108. Baoum, A.; Ovcharenko, D.; Berkland, C. Calcium condensed cell penetrating peptide complexes offer highly efficient, low toxicity gene silencing. Int. J. Pharm. 2012, 427, 134–142. [Google Scholar] [CrossRef] [PubMed]
  109. Khondee, S.; Baoum, A.; Siahaan, T.J.; Berkland, C. Calcium condensed LABL-TAT complexes effectively target gene delivery to ICAM-1 expressing cells. Mol. Pharm. 2011, 8, 788–798. [Google Scholar] [CrossRef] [PubMed]
  110. Baoum, A.A.; Berkland, C. Calcium condensation of DNA complexed with cell-penetrating peptides offers efficient, noncytotoxic gene delivery. J. Pharm. Sci. 2011, 100, 1637–1642. [Google Scholar] [CrossRef] [PubMed]
  111. Khosravi-Darani, K.; Mozafari, M.R.; Rashidi, L.; Mohammadi, M. Calcium based non-viral gene delivery: An overview of methodology and applications. Acta Medica Iran 2010, 48, 133–141. [Google Scholar]
  112. Zaitsev, S.; Buchwalow, I.; Haberland, A.; Tkachuk, S.; Zaitseva, I.; Haller, H.; Bottger, M. Histone H1-mediated transfection: Role of calcium in the cellular uptake and intracellular fate of H1-DNA complexes. Acta Histochem. 2002, 104, 85–92. [Google Scholar] [CrossRef] [PubMed]
  113. Hoyer, J.; Neundorf, I. Peptide vectors for the nonviral delivery of nucleic acids. Acc. Chem. Res. 2012, 45, 1048–1056. [Google Scholar] [CrossRef] [PubMed]
  114. 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] [PubMed]
  115. Kawabata, A.; Baoum, A.; Ohta, N.; Jacquez, S.; Seo, G.M.; Berkland, C.; Tamura, M. Intratracheal administration of a nanoparticle-based therapy with the angiotensin II type 2 receptor gene attenuates lung cancer growth. Cancer Res. 2012, 72, 2057–2067. [Google Scholar] [CrossRef] [PubMed]
  116. Nascimento, F.D.; Sancey, L.; Pereira, A.; Rome, C.; Oliveira, V.; Oliveira, E.B.; Nader, H.B.; Yamane, T.; Kerkis, I.; Tersariol, I.L.; et al. The natural cell-penetrating peptide crotamine targets tumor tissue in vivo and triggers a lethal calcium-dependent pathway in cultured cells. Mol. Pharm. 2012, 9, 211–221. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Li, H.; Tsui, T.Y.; Ma, W. Intracellular Delivery of Molecular Cargo Using Cell-Penetrating Peptides and the Combination Strategies. Int. J. Mol. Sci. 2015, 16, 19518-19536. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms160819518

AMA Style

Li H, Tsui TY, Ma W. Intracellular Delivery of Molecular Cargo Using Cell-Penetrating Peptides and the Combination Strategies. International Journal of Molecular Sciences. 2015; 16(8):19518-19536. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms160819518

Chicago/Turabian Style

Li, Hua, Tung Yu Tsui, and Wenxue Ma. 2015. "Intracellular Delivery of Molecular Cargo Using Cell-Penetrating Peptides and the Combination Strategies" International Journal of Molecular Sciences 16, no. 8: 19518-19536. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms160819518

Article Metrics

Back to TopTop