Next Article in Journal
Crystal Structures of Xenon(VI) Salts: XeF5Ni(AsF6)3, XeF5AF6 (A = Nb, Ta, Ru, Rh, Ir, Pt, Au), and XeF5A2F11 (A = Nb, Ta)
Previous Article in Journal
The Dichotomy of Mn–H Bond Cleavage and Kinetic Hydricity of Tricarbonyl Manganese Hydride Complexes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cell-Penetrating Peptides for Use in Development of Transgenic Plants

1
Department of Laboratory Medicine and Biotechnology, College of Medicine, Tzu Chi University, Hualien 970374, Taiwan
2
Department of Life Science, College of Science and Engineering, National Dong Hwa University, Hualien 974301, Taiwan
3
Department of Biological Sciences, College of Arts, Sciences, and Education, Missouri University of Science and Technology, Rolla, MO 65409, USA
4
Department of Natural Resources and Environmental Studies, College of Environmental Studies and Oceanography, National Dong Hwa University, Hualien 974301, Taiwan
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(8), 3367; https://doi.org/10.3390/molecules28083367
Submission received: 18 February 2023 / Revised: 24 March 2023 / Accepted: 7 April 2023 / Published: 11 April 2023

Abstract

:
Genetically modified plants and crops can contribute to remarkable increase in global food supply, with improved yield and resistance to plant diseases or insect pests. The development of biotechnology introducing exogenous nucleic acids in transgenic plants is important for plant health management. Different genetic engineering methods for DNA delivery, such as biolistic methods, Agrobacterium tumefaciens-mediated transformation, and other physicochemical methods have been developed to improve translocation across the plasma membrane and cell wall in plants. Recently, the peptide-based gene delivery system, mediated by cell-penetrating peptides (CPPs), has been regarded as a promising non-viral tool for efficient and stable gene transfection into both animal and plant cells. CPPs are short peptides with diverse sequences and functionalities, capable of agitating plasma membrane and entering cells. Here, we highlight recent research and ideas on diverse types of CPPs, which have been applied in DNA delivery in plants. Various basic, amphipathic, cyclic, and branched CPPs were designed, and modifications of functional groups were performed to enhance DNA interaction and stabilization in transgenesis. CPPs were able to carry cargoes in either a covalent or noncovalent manner and to internalize CPP/cargo complexes into cells by either direct membrane translocation or endocytosis. Importantly, subcellular targets of CPP-mediated nucleic acid delivery were reviewed. CPPs offer transfection strategies and influence transgene expression at subcellular localizations, such as in plastids, mitochondria, and the nucleus. In summary, the technology of CPP-mediated gene delivery provides a potent and useful tool to genetically modified plants and crops of the future.

Graphical Abstract

1. Introduction

The current trends in crop yield fall short of meeting the demand, as the global requirement for food is projected to double in the next 30 years [1]. Modern agriculture is facing major global challenges, such as loss of biodiversity, chemical contamination of soils, plant pests, and diseases [2], all of which can directly affect plant health and productivity. Genetically modified plants and crops provide one of the solutions to increase global food production with improved gains in yield and resistance to plant diseases or insect pests. Several successful cases of genetically modified plants have conferred phytoprotection against insects, pests, and pathogens, such as overexpression of proteinase inhibitor genes from legumes [3], recombinant Bt toxic proteins from soil bacteria Bacillus thuringiensis [4], α-amylase inhibitors, and plant lectins [5]. In spite of these successful examples, there is a need to develop alternative strategies of phytoprotection.
Transgenic plants are defined as plants containing gene modifications and expressing recombinant proteins or products from foreign genes [6]. The success of transgenic plants depend on favorable methods of gene delivery. The first transgenic plant was reported in 1983 when an antibiotic-resistant Ti plasmid was delivered into tobacco, mediated by A. tumefaciens [7]. Subsequently, tremendous gene delivery strategies, such as particle bombardment (biolistics), were applied in plants. Flourishing developments of biotechnology in exogenous nucleic acid delivery have brought a great improvement in transgenic plants [8,9].
Genetic transformation methods in plants were generally divided into two types: direct and indirect gene delivery methods [9]. DNAs or RNAs can be introduced into plants either directly or packaged by specific viruses or bacteria, then transferred into plants via an indirect method [9]. Additional gene delivery classifications include physical, chemical, and biological methods (Figure 1) [10]. Physical methods, such as microinjection, biolistics, electroporation, silicon carbide fibers, laser-mediated DNA delivery, sonoporation, hydrodynamic force, etc., facilitate nucleic acids to penetrate cell membrane directly [11]. The advantages of physical methods for gene delivery are plant- and genome-type independence. Large- or small-sized plasmid DNAs can be delivered by these methods, while DNA transformations in some recalcitrant plants, such as cereals and legumes, are wildly applied [12]. However, the criticisms of physical methods are irreversible tissue damage and integration of genes into host genomes [10]. In microinjection and electroporation methods, plant sample preparations become protoplasts or a single cell, and this complicated procedure makes the drawback for transgenic plant applications [10].
Polyethylene glycol (PEG)-, diethylaminoethanol (DEAE)-dextran-, calcium phosphate-, dendrimer-, and liposome-mediated gene transfers were categorized as chemical methods [11,13]. Hygromycin resistance gene was introduced into the protoplasts of rice, and these transgenic rice plants generated viable seeds [14]. The benefits of liposome-mediated nucleic acids delivery include DNA protection from nuclease digestion, as well as suitability for multiple types of plant cells, such as protoplasts and plasmodesmata [15]. Calcium phosphate is a cheap and easily handled DNA delivery method that possesses economic benefits. DNAs are able to interact with positively charged calcium ions via electrostatics, form precipitates, and enter cells by endocytosis [11]. Endocytosis is the major route for DNA delivery in chemical method-mediated plant transformation [16]. However, nucleic acids, which might be digested in lysosomes, and the limitation on plasmid DNA size reduce the potential of application in transgenic plants [16,17].
A. tumefaciens is the earliest and most common biological method for DNA delivery in transgenic plants (Figure 1) [7,12]. This bacterium has the ability to deliver genes into the host genome, leading to the establishment of Agrobacterium-mediated transformation as one of the standard protocols in transgenic plants [8,12]. However, there are some disadvantages in Agrobacterium-mediated transformations [18,19]. First of all, low transformation yield was reported in certain plant species, which might be due to the specificity of hosts [19]. This reduces the enthusiasm in transgenic engineering in plants. Additionally, it took long tissue culture periods to recover from transformation, and low yield of stably transformed plants was observed in Agrobacterium-mediated transformation [8,18,19]. Due to these limitations, virus-based vectors become an alternative (Figure 1). Protein expressions by virus-based infected genes are able to be produced more quickly and have a greater yield [20]. However, safety concerns and continuous re-infection are the drawbacks for this method [7]. Recently, non-viral peptide-based gene delivery systems have become popular in transgenic engineering. However, not all peptides possess the same abilities for gene delivery. The membrane-active peptides, cell-penetrating peptides (CPPs), were reported as a good tool for nucleic acid transfection [21,22]. We devote the sections below to introduce CPPs and their roles in transgenic plants in detail.

2. Cell-Penetrating Peptides (CPPs)

CPPs are short and membrane-active peptides [21]. In general, macromolecules, such as DNAs, RNAs, and proteins, are impermeable to cell membranes. The cell membrane is a natural barrier to prevent harmful exogenous or pathological molecules from entering the cell freely, and to maintain the osmotic balance within the cell. Some functional proteins are able to enter cells via specific receptors or channels, while nucleic acids alone are generally not [23]. Not only can CPPs enter cells by themselves, but also deliver various cargoes, including nucleic acids, into living cells [24]. However, nucleic acids are not the only macromolecules that CPPs are able to deliver. CPPs also can serve as a Trojan horse, while peptides/proteins [25], nanoparticles [26], pharmaceutical molecules, and small drugs [24,27,28] play the role of Achilles. The selectivity and efficiency of drug/molecules delivery are significantly improved when CPPs cooperate with liposomes/micelles [29,30]. Most CPPs have been shown to be nontoxic, and do not interfere with functionality of the delivered biomacromolecule [31].
There are various interactions between CPPs and their cargoes. CPPs and cargoes can form complexes with covalent bonds [32], noncovalent interactions [21], and covalent- and noncovalent-synchronous linkages [33]. The potential of bio-membrane penetration in CPPs is amazing. Until now, studies indicated that CPPs were able to penetrate different targets, including mammalian cells [21], plant cells/tissues [32,33,34,35,36,37,38,39], rodent skin and intestinal mucosa [25], prokaryotes [40,41], fungi [42], insect cells [43], paramecia [44], and rotifers which were individual organisms containing thick cuticles [45]. Recently, different modifications on CPPs, such as D-form amino acid applications, branches on backbone sequences, cyclic structures alterations, and non-standard amino acid substitutions were designed to increase their internalization efficiency and stability [21,46,47,48]. Highly cellular penetration efficiencies and non-cytotoxic properties make CPPs an ideal delivery system for therapeutic drugs, gene therapies, and transgenic plants [22,25,49,50,51].

2.1. Categories of Cell-Penetrating Peptides

Since the first CPP, trans-activator of transcription (Tat) protein of the human immunodeficiency virus type 1 (HIV-1), was identified, 1855 CPP entries with sequence information have been deposited and annotated in a database repository named the CPPsite 2.0 [52]. In this database, CPPs were categorized according to various characterizations, such as peptide lengths, chemical and physical properties, and structures [52]. Additionally, it offered further predicting of CPPs. Taking peptide length as an example, there were 60 CPPs containing five or less residues in this CPPsite 2.0 online database. L5a (Table 1), the peptide composed of only five amino acid residues, was one of them [53]. Some novel peptides or protein segments were also predicted as CPPs by the machine learning web-server of KELM-CPPpred [54]. Classification of CPPs can help identify a suitable CPP for a specific purpose. Three types of CPPs were categorized based upon their chemical and physical properties [55].

2.1.1. Cationic Type

The Tat protein transduction domain (PTD) is the most well-known representative of a cationic CPP. When Tat protein, consisting of 86 amino acid residues, was first discovered by two independent groups in 1988 [84,85], researchers found that the Tat PTD (amino acid residues 48–60) was composed of eight cationic amino acids [56]. These cationic amino acids were regarded as the key for cellular translocation. Replacing the positively charged residues with other amino acids decreases cell entry efficiency [59,86]. Later, many derivatives from the Tat PTD were made. Among them, poly-arginines, such as octa-arginine (R8) and nona-arginine (R9), display more amazing cellular uptake efficiency [21,59,86]. Besides the Tat PTD and poly-arginines, penetratin, the third helix of the homeodomain of Antennapedia protein (pAntp), diatos peptide vector 1047 (DPV1047), PTD-5, and poly-lysines were among the cationic CPPs [55]. Studies indicated that peptides containing much richer arginines correspond with higher cellular penetration efficiencies than the peptides with lysine-rich residues [55,59]. Peptide structures were studied and the guanidinium groups on arginines were considered as the crucial factor for membrane disturbance and peptide entry [55].

2.1.2. Amphipathic Type

Amphipathic CPPs are peptides that contain both polar and nonpolar regions of amino acids. VP22, KALA, GALA, model amphipathic peptide (MAP), Pep-1, VE-cadherin-derived peptide (pVEC), alternative reading frame (ARF), and MPG are the examples of amphipathic CPPs [29,55]. They may contain valine, leucine, isoleucine, and alanine to form the hydrophobic portions on their secondary structures, and their hydrophilic portions are composed of charged residues like lysine, arginine, and histidine [55]. Positively charged portions on amphipathic CPPs increase the solubility and offer high affinity to cell membrane with electrostatic interaction, while the hydrophobic portions are able to fuse with lipids easily and allow peptides to enter cells by membrane-inverted fusions [54]. Due to these two portions, α-helix (hydrophilic propensities) and β-sheet (hydrophobic propensities) are easily observed in amphipathic CPPs [55].

2.1.3. Hydrophobic Type

The number of hydrophobic CPPs is fewer compared to two other types of CPPs. In this hydrophobic family, peptides contain only non-polar residues and limited net charges [55]. C105Y, PFVYLI (derived from the c-term of C105Y), and Pep-7 were representatives [55]. However, some studies suggested that MAP, which contained both lysine and non-polar residues and was categorized in the amphipathic group, should belong to the hydrophobic CPP [22,87]. The reason for this sorting was based upon the lack of cell selectivity. Hydrophobic CPPs usually display a wide penetrating ability in various tissue types by membrane disorganization [87]. As shown, there are many pros of hydrophobic CPPs and MAP containing these properties. Unfortunately, the cons come from the same properties. Their entry mechanisms lead to higher hemolysis and membrane pore formations, resulting in adverse side effects and limitations in applications [87].

2.2. Mechanisms of Cellular Internalization of Cell-Penetrating Peptide/Cargo Complexes

Although mechanisms of cellular internalization of CPPs have been continuously studied, the discrepancies in understanding their entry routes remain considerable. Many factors, such as primary sequences of CPPs, modifications on peptides, types of cargoes, concentrations of CPPs, linkages between CPPs and cargoes, as well as entry targets of cell lines, can influence penetrating mechanisms [21,22]. Up to date, energy-dependent endocytic pathway and energy-independent direct membrane translocation seem to be the major uptake routes [22]. As shown in Figure 2, in direct membrane translocation, low temperature and endocytic inhibitors cannot stop the entry of CPPs or CPP/cargo complexes [66]. The process of this mechanism starts with attaching CPPs or CPP/cargo complexes to lipid membrane utilizing electrostatics or hydrophobic interactions, followed by membrane destabilization [55,87]. Direct membrane translocation can be further divided into three different pathways: inverted micelles, carpet model, and pore formation [57,88]. For the inverted micelles pathway, CPPs or CPP/cargo complexes attach to membrane surface, followed by invagination of lipid bilayers [60]. Liposome-like structure will form later to package the CPPs or CPP/cargo complexes, and phospholipid inversion flips CPPs or CPP/cargo complexes to enter the cytoplasm (Figure 2). The carpet model is a concentration-dependent model, which is usually applied in α-helical cationic CPPs [89]. CPPs remain parallel to the surface without insertion at low concentrations, while changing the membrane fluidity by redirection of phospholipids and making micelles and pores at high concentrations. Pore formation, also called the Barrel-Stave model, occurs in CPPs with strong positive charge [88], such as HR9 (charge: +14 calculated by CellPPD website [90]) or amphipathic CPPs with an α-helical structure, such as mastoparan [91]. CPPs or CPP/cargo complexes enter cells by direct membrane translocation to avoid the trap from lysosomes, but this action causes highly disturbed membranes, leading to a high risk of cellular injury. Energy-dependent endocytosis is also reported in some CPPs or CPP/cargo complexes [21,22,55].
Endocytosis is a natural process involving at least four subtypes: clathrin-mediated endocytosis, caveolae-mediated endocytosis, clathrin- and caveolae-independent endocytosis, and macropinocytosis [21,38]. Both clathrin-mediated and caveolae-mediated endocytic pathways involve receptor absorptions, microtubule and actin rearrangements, while macropinocytosis (Figure 3), an atypical endocytic route, only involves actin rearrangements [21,69,92,93]. It is hard to conclude whether the types of CPPs or CPP/cargo complexes highly correspond to the types of endocytosis. Studies indicated that CPPs or CPP/cargo complexes enter cells primarily through endocytosis [22,55]; however, SR9 and PR9 enter cells by endocytosis as well [21,69]. Furthermore, SR9 carrying nanoparticles enter cells by multiple endocytic pathways [26].
The relationship between the types of CPPs and their entry mechanisms is unclear. Most studies indicated that cationic CPPs interact with cell membranes by electrostatics and cause disruption of lipid bilayer [94]. This process either thins the cell membrane, according to the membrane thinning model, or makes some pores in cellular membranes, according to the pore formation model [94]. This led Xie et al. to propose that most cationic CPPs enter cells by direct membrane translocation [88]. Pore formations easily perturb the membrane stability, which raises the concern of cytotoxicity. However, not all cationic CPPs enter cells by direct membrane translocation. For instance, SR9 and PR9, two cationic CPPs, use endocytosis for cellular entry (Table 1).
Amphipathic CPPs possess both hydrophobic and hydrophilic properties, which are considered to be a key factor for intracellular internalization [95]. As shown in Table 1, most amphipathic CPPs, such as VP22, KALA, and GALA, follow the endocytic pathway. However, there are some amphipathic CPPs, like MPG and Pep-1, which use direct membrane translocation. Derakhshankhah and Jafari proposed that the hydrophobic portions of amphipathic CPPs insert into cellular membranes by hydrophobic force and polymerization, while the hydrophilic portions of amphipathic CPPs form pores in cellular membranes [94]. However, this hypothesis cannot fully explain the complexity of cellular internalization as other factors may also be involved in cellular entry. Target cells/tissues/species are also determinants in cellular internalization. SR9/quantum dot complexes entered A549 cells and prokaryotes by using multiple pathways [26] and macropinocytosis [68], respectively. Kauffman et al. indicated that cationic CPPs enter cells by endocytosis at low concentrations (<10 µM), and switch to direct membrane translocation at high concentrations [96]. The authors also suggested that high concentrations of cationic CPPs might have caused membrane disruption and possibly increased toxicity [96].

3. Subcellular Targets for Gene Delivery

CPPs have demonstrated remarkable ability to deliver diverse biomacromolecules into various plant species. The plasmid DNA delivery mediated by CPPs displayed a high potential and efficiency in plant root cells [67], embryos [97], and leaf cells [82] without protoplast preparations. Positively charged CPPs possess the abilities to interact with, condense, and package plasmid DNAs. The combination ratio between CPP and nucleic acid (Figure 2), also called nitrogen (NH3+)/phosphate (PO4) (N/P) ratio [98], is key to DNA condensation and packaging. It further affects gene delivery efficiency [53,67]. An optimal N/P ratio makes CPP/DNA complexes more stable and is able to raise gene delivery efficiency. A good transgenic efficiency also depends on other factors, such as long-term stability of CPP/DNA complexes in cytosol, evasion from the endosome–lysosome system, targeted site of gene expression, and DNA releasing from CPP/DNA complexes [34,38,81,83,99]. The efficiency of cytoplasmic delivery by the predominant endosomal pathway is typically very low. A study showed that glutathione-responsive CPPs are able to escape from endosome entrapment and release DNAs at a higher rate to achieve gene transfer in plants [34]. Aside from efficiency, targeted delivery is also crucial in transgenic plant development [100]. Various DNA plasmids were designed and applied to the genes that were successfully achieved for development of transgenic plants (Table 2). Here, we discussed three major subcellular targets for CPP/DNA complex delivery: nucleus, plastids, and mitochondria (Figure 2).

3.1. Nucleus

Nuclear localization signal (NLS) is a small and basic peptide containing four lysines, one arginine, and several nonpolar residues [76] commonly found in CPP sequences (Table 1). Proteins or peptides containing this short signal are recognized by importin, and are transported into nucleus through the classical nuclear import pathway [76]. MPG, a chimeric CPP composed of HIV glycoprotein 41 and SV40 T antigen, is an example of NLS. NLS enters nucleus not only by itself but also with its cargoes [55]. According to the chemical and physical properties of NLS, it is considered a cationic peptide [27]. However, many studies suggested this NLS to be categorized as an amphipathic peptide, because its primary sequence contains both cationic and hydrophobic residues [22,55]. Fagerlund et al. suggested that lysine/arginine-rich NLS on signal transducers and activators of transcription 1 (STAT1) homodimer proteins and STAT1-STAT2 heterodimer proteins is key to both DNA binding and importin interaction [102]. Mutations of the conserved arginine/lysine-rich portions were able to prohibit nuclear import. Furthermore, R9-based CPPs (without NLS) affirmed the principal role of arginine in nuclear entry [66,69]. Both R9-green fluorescent protein (GFP) and NLS-R9-GFP displayed nuclear targeting in mung bean roots [67]. SR9 and PR9 entered cells via multiple pathways and classical endocytosis, respectively [21]. However, they all escaped from the endosome–lysosome system and entered nuclei [69]. Recent data studied by Kurnaeva et al. demonstrated that arginine residues are much more critical than lysines in NLS actions [103]. Therefore, NLS-tagged CPPs or arginine-rich CPPs play an important role in the nucleus delivery of nucleic acids, and their nuclear targeting abilities dramatically increase successful results in transgenic plants.

3.2. Chloroplasts (Plastids)

Chloroplasts (a.k.a., plastids in plants) contain their own genomes and are the core components for photosynthesis. According to the membrane structure, plastids are divided into two groups: primary and secondary plastids [104]. Primary plastids are found in most algae and plants, while plankton typically belongs to the category of secondary plastid organisms. Plastid genomes are essential, as genes in plastids regulate not only metabolism of photosynthesis, but also energy transfer and storage [105]. Plastids also influence the expression of nuclear genes via plastid-to-nucleus signaling pathways, which regulate plastidic and extraplastidic processes to cope with environmental changes [105].
In recent years, transgenic plastids are gaining more attraction in biotechnology for the following reasons: (1) the genome in plastids is smaller than chromosomes in nuclei, only contains about 150 kb in molecular mass, and is easily manufactured by humans [106]; (2) a mature chloroplast contains a high copy number of circular double-stranded DNA, which is able to produce large amounts of recombinant proteins, which is very important for vaccine or economic production [107]; and (3) plastids are the maternal inheritance in most plant species. Plastid genetic engineering, such as in transplastomic plants, manipulates organellar DNA without changing the nuclear genes. Extranuclear genetic engineering prevents genetic pollution from the nucleus and protects wild-type plants or relative wild species [108].
Macromolecules tagged with a specific signal are essential for organelle-targeted delivery. Chloroplast transit peptides (CTPs, a.k.a. chloroplast-targeting peptides) are special peptides containing 33–35% hydrophobic, 22–23% hydroxylated, and 14–15% positively charged amino acids [109]. Shen et al. indicated that the most efficient CTPs in rice is RC2, and its sequence also follows the similar percentage of hydrophobic, hydroxylated, and cationic residues [77]. Thagun et al. successfully combined plasmid DNA, CTP (KH9-AtOEP34) [78], and CPP (BP100) [81] as a complex system to deliver DNAs into chloroplasts [99]. Further, they used the above complexes as nanocarriers, transfecting the plasmid DNA into chloroplasts after spraying on leaf surfaces [101]. The CTP/DNA complexes were transported from the extracellular space to the chloroplast stroma in Arabidopsis leaves [79]. These studies on the CPP(CTP)/DNA complex system dramatically enhanced the transgenesis without protoplast preparations nor callus formations, and provided a useful tool for rapid and effective plastids engineering in plants.

3.3. Mitochondria

Mitochondrion is another valuable target in plant genetic modification. Mitochondria are the energy-producing organelles that contain the plasma membrane-like double membranes, their own genome (i.e., mitochondrial DNA; mtDNA), and a transcription-translation system [110]. mtDNA is a small and circular double-stranded DNA, similar to a plasmid DNA. This characteristic challenges scientists studying mtDNA modifications [110]. However, many factors, such as low transgenic efficiency, poor cytosolic entry, complicated preparation protocols, high mobility of mitochondria, and a limited number of cargo types, remain to be resolved before gene delivery into the mitochondria of plant cells can be widely used [28,83].
Foreign DNAs or cargoes tagged with mitochondrial targeting sequence (MTS) have a higher chance to be transported to mitochondria [28]. MTS, like CTP, is a short peptide signal, which is recognized by mitochondrial outer membrane receptor complex and interacts with components in mitochondrial protein import pathway [111]. The exciting mitochondrial transgenic results were first published by Chuah et al. [81]. They fused MTS with cationic lysine/histidine repeat residues ((KH)9), becoming the fused peptide MTPKH, and this MTPKH formed complexes with CPPs (BP100). CPP/plasmid DNA complexes penetrated plasma membrane through CPPs, while MTPKH/DNA complexes were found to be localized into mitochondria [81]. The Renilla luciferase gene expression in mitochondria of A. thaliana illustrated that CPP-mediated gene delivery can be applied in mitochondrial transgenic engineering [81]. Recently, Xiao et al. developed two novel cell-penetrating mitochondrial-targeting MitoLigand ligands (miniCPM3 and SeSe-TPP) that contain 2~3 hydrophobic aromatic amino acids and 3~4 arginine residues [28]. This ligand design included the MTS conserved sequence, a hydrophobic-, and a cationic-rich amphipathic helix [111]. They found that MitoLigand-delivered cargoes were predominantly localized inside mitochondria after cellular uptake and endosomal escape [28]. Artificial peptide (LURL)3 was another novel cell-penetrating MTS that demonstrates the importance of hydrophobicity and helicity for mitochondrial localization [83]. Together, these effective peptide-based methods provide a starting point for the development of more sophisticated plant mitochondrial transfection strategies.

4. Conclusions

Delivery and expression of exogenous genes in plants have economic potential in biotechnology and industry. Physical, chemical, and biological methods for gene deliveries have been developed for more than three decades. CPPs-based gene delivery systems bring a bright prospect for transgenesis in agriculture. In addition, this review sums up the classification of CPPs, cellular entry mechanisms, cytotoxicity, and various genes applied for development of transgenic plants. Varying designs on primary sequences of CPPs result in different cellular entry routes with different transfection efficiencies. CPPs can be modified for specific organelle-targeted delivery in plant cells. By targeting nuclei, chloroplasts, and mitochondria, CPPs/DNAs complexes elevate certain gene expressions of interest, which may increase higher agricultural yields.

Author Contributions

Conceptualization, B.R.L. and H.-J.L.; writing—original draft preparation, B.R.L.; writing—review and editing, Y.-W.H., C.-W.C. and 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 author.

Acknowledgments

We are grateful to Mikaela Ritchie from the Missouri University of Science and Technology for editing.

Conflicts of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Abbreviations

CPP: cell-penetrating peptide; CTP, chloroplast transit peptide; DEAE, diethylaminoethanol; GFP, green fluorescent protein; HIV, human immunodeficiency virus; MAP, model amphipathic peptide; mtDNA, mitochondrial DNA; MTS, mitochondrial targeting sequence; NLS, nuclear localization signal; N/P, nitrogen (NH3+)/phosphate (PO4); PEG, polyethylene glycol; PTD, protein transduction domain; STAT, signal transducers and activators of transcription; Tat, trans-activator of transcription.

References

  1. Ray, D.K.; Mueller, N.D.; West, P.C.; Foley, J.A. Yield trends are insufficient to double global crop production by 2050. PLoS ONE 2013, 8, e66428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Akanmu, A.O.; Babalola, O.O.; Venturi, V.; Ayilara, M.S.; Adeleke, B.S.; Amoo, A.E.; Sobowale, A.A.; Fadiji, A.E.; Glick, B.R. Plant disease management: Leveraging on the plant-microbe-soil interface in the biorational use of organic amendments. Front. Plant Sci. 2021, 12, 700507. [Google Scholar] [CrossRef] [PubMed]
  3. Sultana, M.S.; Millwood, R.J.; Mazarei, M.; Stewart, C.N., Jr. Proteinase inhibitors in legume herbivore defense: From natural to genetically engineered protectants. Plant Cell Rep. 2022, 41, 293–305. [Google Scholar] [CrossRef] [PubMed]
  4. Then, C.; Miyazaki, J.; Bauer-Panskus, A. Deficiencies in the risk assessment of genetically engineered Bt cowpea approved for cultivation in Nigeria: A critical review. Plants 2022, 11, 380. [Google Scholar] [CrossRef]
  5. Haq, S.K.; Atif, S.M.; Khan, R.H. Protein proteinase inhibitor genes in combat against insects, pests, and pathogens: Natural and engineered phytoprotection. Arch. Biochem. Biophys. 2004, 431, 145–159. [Google Scholar] [CrossRef]
  6. Key, S.; Ma, J.K.; Drake, P.M. Genetically modified plants and human health. J. R. Soc. Med. 2008, 101, 290–298. [Google Scholar] [CrossRef] [Green Version]
  7. Nandy, D.; Maity, A.; Mitra, A.K. Target-specific gene delivery in plant systems and their expression: Insights into recent developments. J. Biosci. 2020, 45, 30. [Google Scholar] [CrossRef]
  8. Vain, P. Thirty years of plant transformation technology development. Plant Biotechnol. J. 2007, 5, 221–229. [Google Scholar] [CrossRef]
  9. Chen, Q.; Lai, H. Gene delivery into plant cells for recombinant protein production. Biomed. Res. Int. 2015, 2015, 932161. [Google Scholar] [CrossRef] [Green Version]
  10. Peng, L.H.; Gu, T.W.; Xu, Y.; Dad, H.A.; Liu, J.X.; Lian, J.Z.; Huang, L.Q. Gene delivery strategies for therapeutic proteins production in plants: Emerging opportunities and challenges. Biotechnol. Adv. 2022, 54, 107845. [Google Scholar] [CrossRef]
  11. Holl, N.J.; Lee, H.J.; Huang, Y.W. Evolutionary timeline of genetic delivery and gene therapy. Curr. Gene Ther. 2021, 21, 89–111. [Google Scholar] [CrossRef] [PubMed]
  12. Imani, J.; Kogel, K.H. Plant transformation techniques: Agrobacterium- and microparticle-mediated gene transfer in cereal plants. Methods Mol. Biol. 2020, 2124, 281–294. [Google Scholar]
  13. Ozyigit, I.I. Gene transfer to plants by electroporation: Methods and applications. Mol. Biol. Rep. 2020, 47, 3195–3210. [Google Scholar] [CrossRef] [PubMed]
  14. Hayashimoto, A.; Li, Z.; Murai, N. A polyethylene glycol-mediated protoplast transformation system for production of fertile transgenic rice plants. Plant Physiol. 1990, 93, 857–863. [Google Scholar] [CrossRef] [Green Version]
  15. Gad, A.E.; Rosenberg, N.; Altman, A. Liposome-mediated gene delivery into plant cells. Physiol. Plant. 1990, 79, 177–183. [Google Scholar] [CrossRef]
  16. Alghuthaymi, M.A.; Ahmad, A.; Khan, Z.; Khan, S.H.; Ahmed, F.K.; Faiz, S.; Nepovimova, E.; Kuca, K.; Abd-Elsalam, K.A. Exosome/liposome-like nanoparticles: New carriers for CRISPR genome editing in plants. Int. J. Mol. Sci. 2021, 22, 7456. [Google Scholar] [CrossRef]
  17. McLenachan, S.; Sarsero, J.P.; Ioannou, P.A. Flow-cytometric analysis of mouse embryonic stem cell lipofection using small and large DNA constructs. Genomics 2007, 89, 708–720. [Google Scholar] [CrossRef] [Green Version]
  18. Davey, M.R.; Soneji, J.R.; Rao, M.N.; Kourmpetli, S.; Bhattacharya, A.; Kole, C. Generation and deployment of transgenic crop plants: An overview. In Transgenic Crop Plants: Principles and Development; Kole, C., Michler, C.H., Abbott, A.G., Hall, T.C., Eds.; Springer: Berlin/Heidelberg, Germany, 2010; Volume 1, pp. 1–29. [Google Scholar]
  19. Hwang, H.H.; Yu, M.; Lai, E.M. Agrobacterium-mediated plant transformation: Biology and applications. Arab. Book 2017, 15, e0186. [Google Scholar] [CrossRef] [Green Version]
  20. Yusibov, V.; Rabindran, S.; Commandeur, U.; Twyman, R.M.; Fischer, R. The potential of plant virus vectors for vaccine production. Drugs R D 2006, 7, 203–217. [Google Scholar] [CrossRef]
  21. 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. [Google Scholar] [CrossRef]
  22. Taylor, R.E.; Zahid, M. Cell penetrating peptides, novel vectors for gene therapy. Pharmaceutics 2020, 12, 225. [Google Scholar] [CrossRef] [Green Version]
  23. Sokolova, V.; Epple, M. Inorganic nanoparticles are carriers of nucleic acids. Angew. Chem. Int. Ed. Engl. 2008, 47, 1382–1395. [Google Scholar] [CrossRef]
  24. Hingorani, D.V.; Camargo, M.F.; Quraishi, M.A.; Adams, S.R.; Advani, S.J. Tumor activated cell penetrating peptides to selectively deliver immune modulatory drugs. Pharmaceutics 2021, 13, 365. [Google Scholar] [CrossRef] [PubMed]
  25. Korivi, M.; Huang, Y.W.; Liu, B.R. Cell-penetrating peptides as a potential drug delivery system for effective treatment of diabetes. Curr. Pharm. Des. 2021, 27, 816–825. [Google Scholar] [CrossRef] [PubMed]
  26. 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] [PubMed] [Green Version]
  27. Habault, J.; Poyet, J.L. Recent advances in cell penetrating peptide-based anticancer therapies. Molecules 2019, 24, 927. [Google Scholar] [CrossRef] [Green Version]
  28. Xiao, Q.; Du, W.; Dong, X.; Du, S.; Ong, S.Y.; Tang, G.; Zhang, C.; Yang, F.; Li, L.; Gao, L.; et al. Cell-penetrating mitochondrion-targeting ligands for the universal delivery of small molecules, proteins and nanomaterials. Chemistry 2021, 27, 12207–12214. [Google Scholar] [CrossRef]
  29. 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]
  30. Kardani, K.; Milani, A.; Shabani, S.H.; Bolhassani, A. Cell penetrating peptides: The potent multi-cargo intracellular carriers. Expert Opin. Drug Deliv. 2019, 16, 1227–1258. [Google Scholar] [CrossRef]
  31. Chugh, A.; Eudes, F.; Shim, Y.S. Cell-penetrating peptides: Nanocarrier for macromolecule delivery in living cells. IUBMB Life 2010, 62, 183–193. [Google Scholar] [CrossRef]
  32. 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]
  33. 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] [PubMed]
  34. Chuah, J.A.; Numata, K. Stimulus-responsive peptide for effective delivery and release of DNA in plants. Biomacromolecules 2018, 19, 1154–1163. [Google Scholar] [CrossRef]
  35. Numata, K.; Horii, Y.; Oikawa, K.; Miyagi, Y.; Demura, T.; Ohtani, M. Library screening of cell-penetrating peptide for BY-2 cells, leaves of Arabidopsis, tobacco, tomato, poplar, and rice callus. Sci Rep 2018, 8, 10966. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Guo, B.; Itami, J.; Oikawa, K.; Motoda, Y.; Kigawa, T.; Numata, K. Native protein delivery into rice callus using ionic complexes of protein and cell-penetrating peptides. PLoS ONE 2019, 14, e0214033. [Google Scholar] [CrossRef] [Green Version]
  37. Terada, K.; Gimenez-Dejoz, J.; Miyagi, Y.; Oikawa, K.; Tsuchiya, K.; Numata, K. Artificial cell-penetrating peptide containing periodic α-aminoisobutyric acid with long-term internalization efficiency in human and plant cells. ACS Biomater. Sci. Eng. 2020, 6, 3287–3298. [Google Scholar] [CrossRef] [Green Version]
  38. Miyamoto, T.; Tsuchiya, K.; Numata, K. Endosome-escaping micelle complexes dually equipped with cell-penetrating and endosome-disrupting peptides for efficient DNA delivery into intact plants. Nanoscale 2021, 13, 5679–5692. [Google Scholar] [CrossRef]
  39. Watanabe, K.; Odahara, M.; Miyamoto, T.; Numata, K. Fusion peptide-based biomacromolecule delivery system for plant cells. ACS Biomater. Sci. Eng. 2021, 7, 2246–2254. [Google Scholar] [CrossRef]
  40. Islam, M.M.; Odahara, M.; Yoshizumi, T.; Oikawa, K.; Kimura, M.; Su’etsugu, M.; Numata, K. Cell-penetrating peptide-mediated transformation of large plasmid DNA into Escherichia coli. ACS Synth. Biol. 2019, 8, 1215–1218. [Google Scholar] [CrossRef] [Green Version]
  41. Higuchi-Takeuchi, M.; Miyamoto, T.; Foong, C.P.; Goto, M.; Morisaki, K.; Numata, K. Peptide-mediated gene transfer into marine purple photosynthetic bacteria. Int. J. Mol. Sci. 2020, 21, 8625. [Google Scholar] [CrossRef]
  42. Gong, Z.; Ikonomova, S.P.; Karlsson, A.J. Secondary structure of cell-penetrating peptides during interaction with fungal cells. Protein Sci. 2018, 27, 702–713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. 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] [PubMed]
  44. 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] [PubMed]
  45. 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. N. Y. 2013, 15, 584–595. [Google Scholar] [CrossRef] [PubMed]
  46. 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]
  47. Park, S.E.; Sajid, M.I.; Parang, K.; Tiwari, R.K. Cyclic cell-penetrating peptides as efficient intracellular drug delivery tools. Mol. Pharm. 2019, 16, 3727–3743. [Google Scholar] [CrossRef]
  48. Szabó, I.; Yousef, M.; Soltész, D.; Bató, C.; Mező, G.; Bánóczi, Z. Redesigning of cell-penetrating peptides to improve their efficacy as a drug delivery system. Pharmaceutics 2022, 14, 907. [Google Scholar] [CrossRef]
  49. Lehto, T.; Kurrikoff, K.; Langel, U. Cell-penetrating peptides for the delivery of nucleic acids. Expert Opin. Drug Deliv. 2012, 9, 823–836. [Google Scholar] [CrossRef]
  50. Ziemienowicz, A.; Pepper, J.; Eudes, F. Applications of CPPs in genome modulation of plants. Methods Mol. Biol. 2015, 1324, 417–434. [Google Scholar]
  51. Bilichak, A.; Sastry-Dent, L.; Sriram, S.; Simpson, M.; Samuel, P.; Webb, S.; Jiang, F.; Eudes, F. Genome editing in wheat microspores and haploid embryos mediated by delivery of ZFN proteins and cell-penetrating peptide complexes. Plant Biotechnol. J. 2020, 18, 1307–1316. [Google Scholar] [CrossRef]
  52. 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] [PubMed]
  53. 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]
  54. Pandey, P.; Patel, V.; George, N.V.; Mallajosyula, S.S. KELM-CPPpred: Kernel extreme learning machine based prediction model for cell-penetrating peptides. J. Proteome Res. 2018, 17, 3214–3222. [Google Scholar] [CrossRef] [PubMed]
  55. Guidotti, G.; Brambilla, L.; Rossi, D. Cell-penetrating peptides: From basic research to clinics. Trends Pharmacol Sci 2017, 38, 406–424. [Google Scholar] [CrossRef]
  56. 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] [Green Version]
  57. Herce, H.D.; Garcia, A.E. Molecular dynamics simulations suggest a mechanism for translocation of the HIV-1 TAT peptide across lipid membranes. Proc. Natl. Acad. Sci. USA 2007, 104, 20805–20810. [Google Scholar] [CrossRef] [Green Version]
  58. Tietz, O.; Cortezon-Tamarit, F.; Chalk, R.; Able, S.; Vallis, K.A. Tricyclic cell-penetrating peptides for efficient delivery of functional antibodies into cancer cells. Nat. Chem. 2022, 14, 284–293. [Google Scholar] [CrossRef]
  59. Wender, P.A.; Mitchell, D.J.; Pattabiraman, K.; Pelkey, E.T.; Steinman, L.; Rothbard, J.B. The design, synthesis, and evaluation of molecules that enable or enhance cellular uptake: Peptoid molecular transporters. Proc. Natl. Acad. Sci. USA 2000, 97, 13003–13008. [Google Scholar] [CrossRef] [Green Version]
  60. Derossi, D.; Calvet, S.; Trembleau, A.; Brunissen, A.; Chassaing, G.; Prochiantz, A. Cell internalization of the third helix of the Antennapedia homeodomain is receptor-independent. J. Biol. Chem. 1996, 271, 18188–18193. [Google Scholar] [CrossRef] [Green Version]
  61. Drin, G.; Mazel, M.; Clair, P.; Mathieu, D.; Kaczorek, M.; Temsamani, J. Physico-chemical requirements for cellular uptake of pAntp peptide. Role of lipid-binding affinity. Eur. J. Biochem. 2001, 268, 1304–1314. [Google Scholar] [CrossRef]
  62. Lundin, P.; Johansson, H.; Guterstam, P.; Holm, T.; Hansen, M.; Langel, U.; Andaloussi, S.E.L. Distinct uptake routes of cell-penetrating peptide conjugates. Bioconjug. Chem. 2008, 19, 2535–2542. [Google Scholar] [CrossRef] [PubMed]
  63. Nishi, K.; Saigo, K. Cellular internalization of green fluorescent protein fused with herpes simplex virus protein VP22 via a lipid raft-mediated endocytic pathway independent of caveolae and Rho family GTPases but dependent on dynamin and Arf6. J. Biol. Chem. 2007, 282, 27503–27517. [Google Scholar] [CrossRef] [Green Version]
  64. Rhee, M.; Davis, P. Mechanism of uptake of C105Y, a novel cell-penetrating peptide. J. Biol. Chem. 2006, 281, 1233–1240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Oehlke, J.; Scheller, A.; Wiesner, B.; Krause, E.; Beyermann, M.; Klauschenz, E.; Melzig, M.; Bienert, M. Cellular uptake of an α-helical amphipathic model peptide with the potential to deliver polar compounds into the cell interior non-endocytically. Biochim. Biophys. Acta 1998, 1414, 127–139. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. 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]
  67. 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]
  68. 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]
  69. 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] [Green Version]
  70. Suresh, A.; Kim, Y.C. Translocation of cell penetrating peptides on Chlamydomonas reinhardtii. Biotechnol. Bioeng. 2013, 110, 2795–2801. [Google Scholar] [CrossRef]
  71. Simeoni, F.; Morris, M.C.; Heitz, F.; Divita, G. Insight into the mechanism of the peptide-based gene delivery system MPG: Implications for delivery of siRNA into mammalian cells. Nucleic Acids Res. 2003, 31, 2717–2724. [Google Scholar] [CrossRef] [Green Version]
  72. Wyman, T.B.; Nicol, F.; Zelphati, O.; Scaria, P.V.; Plank, C.; Szoka, F.C. Design, synthesis, and characterization of a cationic peptide that binds to nucleic acids and permeabilizes bilayers. Biochemistry 1997, 36, 3008–3017. [Google Scholar] [CrossRef] [PubMed]
  73. Henriques, S.T.; Costa, J.; Castanho, M.A.R.B. Translocation of β-galactosidase mediated by the cell-penetrating peptide pep-1 into lipid vesicles and human HeLa cells is driven by membrane electrostatic potential. Biochemistry 2005, 44, 10189–10198. [Google Scholar] [CrossRef] [PubMed]
  74. Henriques, S.T.; Quintas, A.; Bagatolli, L.A.; Homblé, F.; Castanho, M.A.R.B. Energy-independent translocation of cell-penetrating peptides occurs without formation of pores. A biophysical study with pep-1. Mol. Membr. Biol. 2007, 24, 282–293. [Google Scholar] [CrossRef] [PubMed]
  75. Ragin, A.D.; Morgan, R.A.; Chmielewski, J. Cellular import mediated by nuclear localization signal peptide sequences. Chem. Biol. 2002, 9, 43–948. [Google Scholar] [CrossRef] [Green Version]
  76. Lange, A.; Mills, R.E.; Lange, C.J.; Stewart, M.; Devine, S.E.; Corbett, A.H. Classical nuclear localization signals: Definition, function, and interaction with importin alpha. J. Biol. Chem. 2007, 282, 5101–5105. [Google Scholar] [CrossRef] [Green Version]
  77. Shen, B.R.; Zhu, C.H.; Yao, Z.; Cui, L.L.; Zhang, J.J.; Yang, C.W.; He, Z.H.; Peng, X.X. An optimized transit peptide for effective targeting of diverse foreign proteins into chloroplasts in rice. Sci. Rep. 2017, 7, 46231. [Google Scholar] [CrossRef] [Green Version]
  78. Yoshizumi, T.; Oikawa, K.; Chuah, J.A.; Kodama, Y.; Numata, K. Selective gene delivery for integrating exogenous DNA into plastid and mitochondrial genomes using peptide-DNA complexes. Biomacromolecules 2018, 19, 1582–1591. [Google Scholar] [CrossRef] [Green Version]
  79. Oikawa, K.; Tateishi, A.; Odahara, M.; Kodama, Y.; Numata, K. Imaging of the entry pathway of a cell-penetrating peptide-DNA complex from the extracellular space to chloroplast nucleoids across multiple membranes in Arabidopsis leaves. Front. Plant Sci. 2021, 12, 759871. [Google Scholar] [CrossRef]
  80. Eggenberger, K.; Mink, C.; Wadhwani, P.; Ulrich, A.S.; Nick, P. Using the peptide BP100 as a cell-penetrating tool for the chemical engineering of actin filaments within living plant cells. Chembiochem 2011, 12, 132–137. [Google Scholar] [CrossRef]
  81. Chuah, J.A.; Yoshizumi, T.; Kodama, Y.; Numata, K. Gene introduction into the mitochondria of Arabidopsis thaliana via peptide-based carriers. Sci. Rep. 2015, 5, 7751. [Google Scholar] [CrossRef] [Green Version]
  82. Lakshmanan, M.; Kodama, Y.; Yoshizumi, T.; Sudesh, K.; Numata, K. Rapid and efficient gene delivery into plant cells using designed peptide carriers. Biomacromolecules 2013, 14, 10–16. [Google Scholar] [CrossRef] [PubMed]
  83. Terada, K.; Gimenez-Dejoz, J.; Kurita, T.; Oikawa, K.; Uji, H.; Tsuchiya, K.; Numata, K. Synthetic mitochondria-targeting peptides incorporating α-aminoisobutyric acid with a stable amphiphilic helix conformation in plant cells. ACS Biomater. Sci. Eng. 2021, 7, 1475–1484. [Google Scholar] [CrossRef] [PubMed]
  84. 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] [PubMed]
  85. Frankle, A.D.; Pabo, C.O. Cellular uptake of the tat protein from human immunodeficiency virus. Cell 1988, 55, 1189–1193. [Google Scholar] [CrossRef] [PubMed]
  86. 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] [PubMed]
  87. Zahid, M.; Robbins, P.D. Cell-type specific penetrating peptides: Therapeutic promises and challenges. Molecules 2015, 20, 13055–13070. [Google Scholar] [CrossRef] [Green Version]
  88. Xie, J.; Bi, Y.; Zhang, H.; Dong, S.; Teng, L.; Lee, R.J.; Yang, Z. Cell-penetrating peptides in diagnosis and treatment of human diseases: From preclinical research to clinical application. Front. Pharmacol. 2020, 11, 697. [Google Scholar] [CrossRef]
  89. Pouny, Y.; Rapaport, D.; Mor, A.; Nicolas, P.; Shai, Y. Interaction of antimicrobial dermaseptin and its fluorescently labeled analogues with phospholipid membranes. Biochemistry 1992, 31, 12416–12423. [Google Scholar] [CrossRef]
  90. Gautam, A.; Chaudhary, K.; Kumar, R.; Sharma, A.; Kapoor, P.; Tyagi, A.; Open Source Drug Discovery Consortium; Raghava, G.P.S. In silico approaches for designing highly effective cell penetrating peptides. J. Transl. Med. 2013, 11, 74. [Google Scholar] [CrossRef] [Green Version]
  91. Matsuzaki, K.; Yoneyama, S.; Murase, O.; Miyajima, K. Transbilayer transport of ions and lipids coupled with mastoparan X translocation. Biochemistry 1996, 35, 8450–8456. [Google Scholar] [CrossRef]
  92. 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] [PubMed]
  93. Huang, Z.; Chen, C.W.; Buj, R.; Tangudu, N.K.; Fang, R.S.; Leon, K.E.; Dahl, E.S.; Varner, E.L.; von Krusenstiern, E.; Cole, A.R.; et al. ATM inhibition drives metabolic adaptation via induction of macropinocytosis. J. Cell Biol. 2023, 222, e202007026. [Google Scholar] [CrossRef] [PubMed]
  94. Derakhshankhah, H.; Jafari, S. Cell penetrating peptides: A concise review with emphasis on biomedical applications. Biomed. Pharmacother. 2018, 108, 1090–1096. [Google Scholar] [CrossRef]
  95. Deshayes, S.; Morris, M.C.; Divita, G.; Heitz, F. Interactions of amphipathic CPPs with model membranes. Biochim. Biophys. Acta 2006, 1758, 328–335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Kauffman, W.B.; Guha, S.; Wimley, W.C. Synthetic molecular evolution of hybrid cell penetrating peptides. Nat. Commun. 2018, 9, 2568. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Chugh, A.; Eudes, F. Study of uptake of cell penetrating peptides and their cargoes in permeabilized wheat immature embryos. FEBS J. 2008, 275, 2403–2414. [Google Scholar] [CrossRef]
  98. Holl, N.J.; Dey, M.; Huang, Y.W.; Chiou, S.H.; Lee, H.J. Lactoferricin-derived L5a cell-penetrating peptide for delivery of DNA into cells. Methods Mol. Biol. 2021, 2211, 113–121. [Google Scholar]
  99. Thagun, C.; Chuah, J.A.; Numata, K. Targeted gene delivery into various plastids mediated by clustered cell-penetrating and chloroplast-targeting peptides. Adv. Sci. Weinh. 2019, 6, 1902064. [Google Scholar] [CrossRef] [Green Version]
  100. Thagun, C.; Motoda, Y.; Kigawa, T.; Kodama, Y.; Numata, K. Simultaneous introduction of multiple biomacromolecules into plant cells using a cell-penetrating peptide nanocarrier. Nanoscale 2020, 12, 18844–18856. [Google Scholar] [CrossRef]
  101. Thagun, C.; Horii, Y.; Mori, M.; Fujita, S.; Ohtani, M.; Tsuchiya, K.; Kodama, Y.; Odahara, M.; Numata, K. Non-transgenic gene modulation via spray delivery of nucleic acid/peptide complexes into plant nuclei and chloroplasts. ACS Nano. 2022, 16, 3506–3521. [Google Scholar] [CrossRef]
  102. Fagerlund, R.; Mélen, K.; Kinnunen, L.; Julkunen, I. Arginine/lysine-rich nuclear localization signals mediate interactions between dimeric STATs and importin alpha 5. J. Biol. Chem. 2002, 277, 30072–30078. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Kurnaeva, M.A.; Zalevsky, A.O.; Arifulin, E.A.; Lisitsyna, O.M.; Tvorogova, A.V.; Shubina, M.Y.; Bourenkov, G.; Tikhomirova, M.A.; Potashnikova, D.M.; Kachalova, A.I.; et al. Molecular coevolution of nuclear and nucleolar localization signals inside the basic domain of HIV-1 Tat. J. Virol. 2022, 96, e0150521. [Google Scholar] [CrossRef] [PubMed]
  104. Kroth, P.G. Protein transport into secondary plastids and the evolution of primary and secondary plastids. Int. Rev. Cytol. 2002, 221, 191–255. [Google Scholar] [PubMed]
  105. Larkin, R.M. Influence of plastids on light signalling and development. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2014, 369, 20130232. [Google Scholar] [CrossRef] [Green Version]
  106. Maliga, P. Transplastomic technology for safer and better transgenic crops. Nat. Biotechnol. 1999, 17, 28. [Google Scholar] [CrossRef]
  107. Waheed, M.T.; Ismail, H.; Gottschamel, J.; Mirza, B.; Lössl, A.G. Plastids: The green frontiers for vaccine production. Front. Plant Sci. 2015, 6, 1005. [Google Scholar] [CrossRef] [Green Version]
  108. De Marchis, F.; Bellucci, M. Plastid transformation in sugar beet: An important industrial crop. Methods Mol. Biol. 2021, 2317, 283–290. [Google Scholar]
  109. Bhushan, S.; Kuhn, C.; Berglund, A.K.; Roth, C.; Glaser, E. The role of the N-terminal domain of chloroplast targeting peptides in organellar protein import and miss-sorting. FEBS Lett. 2006, 580, 3966–3972. [Google Scholar] [CrossRef] [Green Version]
  110. Scarpulla, R.C. Transcriptional paradigms in mammalian mitochondrial biogenesis and function. Physiol. Rev. 2008, 88, 611–638. [Google Scholar] [CrossRef] [Green Version]
  111. Omura, T. Mitochondria-targeting sequence, a multi-role sorting sequence recognized at all steps of protein import into mitochondria. J. Biochem. 1998, 123, 1010–1016. [Google Scholar] [CrossRef]
Figure 1. Methodologies of gene delivery in plants. Various methods including physical, chemical, and biological manners were applied in gene delivery. Plant cells were prepared as protoplasts for gene uptake. Plant tissues (callus) and zygotic embryos also served as the transgenic hosts. PEG: polyethylene glycol; DEAE: diethylaminoethanol.
Figure 1. Methodologies of gene delivery in plants. Various methods including physical, chemical, and biological manners were applied in gene delivery. Plant cells were prepared as protoplasts for gene uptake. Plant tissues (callus) and zygotic embryos also served as the transgenic hosts. PEG: polyethylene glycol; DEAE: diethylaminoethanol.
Molecules 28 03367 g001
Figure 2. Cellular internalization mechanisms of CPP/DNA complexes and their subcellular targets in plant cells. There are two main routes of cellular internalization: endocytosis and direct membrane translocation. The internalized CPP/DNA complexes stay in the nucleus, plastids, or mitochondria, depending on signal sequences on CPPs. With endocytic pathway, CPP/DNA complexes have to escape from lysosomes eventually.
Figure 2. Cellular internalization mechanisms of CPP/DNA complexes and their subcellular targets in plant cells. There are two main routes of cellular internalization: endocytosis and direct membrane translocation. The internalized CPP/DNA complexes stay in the nucleus, plastids, or mitochondria, depending on signal sequences on CPPs. With endocytic pathway, CPP/DNA complexes have to escape from lysosomes eventually.
Molecules 28 03367 g002
Figure 3. Schematic diagram of macropinocytosis.
Figure 3. Schematic diagram of macropinocytosis.
Molecules 28 03367 g003
Table 1. Comprehensive information of major cell-penetrating peptides and others.
Table 1. Comprehensive information of major cell-penetrating peptides and others.
CPP/OthersPrimary SequenceCategoriesTarget Cells/Tissues/SpeciesEntry MechanismsCytotoxicityReferences
L5aRRWQWAmphipathicA549 cellsDirect membrane translocationUp to 10 μM is not toxic[53]
Tat PTD (48–60)GRKKRRQRRRPPQCationicHeLa, HL116, CCL39 cellsDirect membrane translocationUp to 100 μM is not toxic[56]
HeLa, CHO cellsDirect membrane translocation
(pore formation)
1 μM, 3-day treatment without cytotoxicity[57,58]
R9RRRRRRRRR *CationicJurkat, murine B, human PBL cellsEndocytosis[59]
Plant tissuesEnergy-independent pathway (link GFP with covalent manner)Up to 2 μg is not toxic[32]
pAntpRQIKIWFQNRRMKWKKCationicK562, HeLa cellsDirect membrane translocation
(inverted micelle)
[60,61]
CHO cellsEndocytosis
(macropinocytosis)
Up to 10 μM is not toxic[62]
VP22DAATATRGRSAASRPTERPRAPARSASRPRRPVDAmphipathicCHO-K1, HeLa cellsEndocytosis
(macropinocytosis)
[29,63]
C105YCSIPPEVKFNKPFVYLIHydrophobicHuH7 cellsEnergy-independent pathway
(cell entry)
[55,64]
Energy-dependent pathway
(nucleolar entry)
MAPKLALKLALKALKAALKLAAmphipathicHeLa, endothelial cellsEnergy-dependent pathway

Energy-independent pathway
Up to 10 μM is not toxic
[29,62,65]
HR9CHHHHHRRRRRRRRRHHHHHCCationicA549, Sf9, plant cells, paramecia, rotifers, prokaryotesDirect membrane translocationUp to 60 μM is not toxic[21,66]
SR9RRRRRRRRRCationicA549 cells
(Link protein with noncovalent manners)
Up to 16 μM is not toxic[21]
A549 cellsMultiple energy-dependent pathways
(Link nanoparticles with noncovalent manners)
Up to 60 μM is not toxic[26]
Plant tissuesMacropinocytosis (link GFP or DNA with noncovalent manners)Up to 16.6 μM is not toxic[32,67]
ProkaryotesMacropinocytosis
(Link nanoparticles with noncovalent manners)
Up to 48 μM is not toxic[68]
PR9FFLIPKGRRRRRRRRRCationicA549 cellsEndocytosis
(Link nanoparticles with noncovalent manners)
Up to 60 μM is not toxic[69]
pVECLLIILRRRIRKQAHAHSKAmphipathicHeLa cellsEndocytosisUp to 10 μM is not toxic[62]
Green algaDirect membrane translocationNo toxicity[70]
MPGGALFLGFLGAAGSTMGAWSQPKKKRKVAmphipathicHS-68, Cos-7, HeLa cellsDirect membrane translocation[29,71]
KALAWEAKLAKALAKALAKHLAKALAKALKACEAAmphipathicCV-1, Hep G2, C2C12, K562, CaCo2 cellsEndocytosis Toxic at the concentration ≥ 25 μM[29,72]
GALAWEAALAEALAAEALAEHLAEALAEALEALAAAmphipathicCV-1, Hep G2, C2C12, K562, CaCo2 cellsEndocytosis[29,72]
Pep-1KETWWETWWTEWSQPKKKRKVAmphipathicHeLa cellsDirect membrane translocation[73,74]
NLSCGYGPKKKRKVGGCationic (or amphipathic)MCF-7, KB, HT29, MIAPACA2, PC3 cellsEnergy-independent pathway[75,76]
RC2MQVWPIEGIKKFETLSYLPPLChloroplast transit peptide (CTP; not a CPP)Rice chloroplasts[77]
KH9-AtOEP34KHKHKHKHKHKHKHKHKHMFAFQYLLVMCationic CPP combined with CTPSeedlings and leaves of A. thaliana and Nicotiana tabacumEndocytosis or direct membrane translocation (for cellular entry)
Unknown (for plastid targeting)
[78,79]
BP100KKLFKKILKYLAmphipathicLeaves of A. thaliana, BY-2 cellsEndocytosis[35,80,81]
MTP-KH9MLSLRQSIRFFKKHKHKHKHKHKHKHKHKHCationic CPP combined with MTSLeaves of A. thalianaEndocytosis (for cellular entry)
Unknown (for mitochondrial targeting)
[81,82]
(LURL)3LURLLURLLURLAmphipathic MTSOnion bulbsGold nanoparticle biolistics (for cellular entry)
Binding to mitochondrial import receptors Tom20 and Tom22 (for mitochondrial targeting)
Negligible toxicity[83]
* A segment of an E. coli-expressed fusion protein, while SR9 is a synthetic peptide.
Table 2. List of various genes applied in transgenic plants.
Table 2. List of various genes applied in transgenic plants.
Delivery MethodsGenesTargetsReferences
Non-CPP-based gene deliveryProteinase inhibitor genesTobacco[3]
Recombinant Bt toxic proteinsVigna ungiguiculata[4,5]
α-Amylase inhibitors, plant lectinsAdzuki bean[5]
Antibiotic-resistant Ti plasmid
(A. tumefaciens mediated transfection)
Tobacco[7]
Hygromycin resistance geneProtoplasts of rice[14]
CPP-based gene deliveryp35S-RLuc-tNOS and p35S-GFP-tNOS plasmidsLeaves of A. thaliana[34]
p35S-Nluc-tNOS or p35S-GFP-tNOS plasmidSeedlings of A. thaliana[38]
pHBT-sGFP(S65T)-NOS plasmidRoots of mung bean and soybean[67]
psbAp:GFP:SPECr:psbAt At plastid genome integration vector, cox2p: GFP:SPECr:cox2t At mitochondrial genome integration vector, and cox2t:SPECr:GFP:cox2p Nt mitochondrial genome integration vectorSeedlings and leaves of A. thaliana or N. tabacum[78]
PsbA-SPECr-sGFP-psbA, Prrn-aadA-sfGFP-Trps, PsbA-SPECr-sGFP-psbA, and Prrn-aadA-sfGFP-TrpsLeaves of A. thaliana[79]
pDONR-cox2:rluc and pDONR-cox2:gfp plasmidsLeaves of A. thaliana[81]
pAct-1GUS plasmidWheat immature embryos[97]
pPrrn::GFP(S65T)::TpsbA, pPrrn::DsRed::TpsbA, and pPpsbA::Rluc plasmidsLeaves of A. thaliana[99]
psfGN155-MxMT and psfGC155-MxMT plasmidsLeaves of N. benthamiana[100]
pBI221, pBI121, and pPpsbA::Rluc plasmidsLeaves of Arabidopsis, soybean, and tomato[101]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, B.R.; Chen, C.-W.; Huang, Y.-W.; Lee, H.-J. Cell-Penetrating Peptides for Use in Development of Transgenic Plants. Molecules 2023, 28, 3367. https://doi.org/10.3390/molecules28083367

AMA Style

Liu BR, Chen C-W, Huang Y-W, Lee H-J. Cell-Penetrating Peptides for Use in Development of Transgenic Plants. Molecules. 2023; 28(8):3367. https://doi.org/10.3390/molecules28083367

Chicago/Turabian Style

Liu, Betty Revon, Chi-Wei Chen, Yue-Wern Huang, and Han-Jung Lee. 2023. "Cell-Penetrating Peptides for Use in Development of Transgenic Plants" Molecules 28, no. 8: 3367. https://doi.org/10.3390/molecules28083367

Article Metrics

Back to TopTop