Skip to main content
The AAPS Journal logoLink to The AAPS Journal
. 2009 Oct 16;11(4):671. doi: 10.1208/s12248-009-9143-y

Nonviral Gene Delivery: Principle, Limitations, and Recent Progress

Mohammed S Al-Dosari 1,, Xiang Gao 2
PMCID: PMC2782077  PMID: 19834816

Abstract

Gene therapy is becoming a promising therapeutic modality for the treatment of genetic and acquired disorders. Nonviral approaches as alternative gene transfer vehicles to the popular viral vectors have received significant attention because of their favorable properties, including lack of immunogenicity, low toxicity, and potential for tissue specificity. Such approaches have been tested in preclinical studies and human clinical trials over the last decade. Although therapeutic benefit has been demonstrated in animal models, gene delivery efficiency of the nonviral approaches remains to be a key obstacle for clinical applications. This review focuses on existing and emerging concepts of chemical and physical methods for delivery of therapeutic nucleic acid molecules in vivo. The emphasis is placed on discussion about problems associated with current nonviral methods and recent efforts toward refinement of nonviral approaches.

Key words: gene delivery, gene therapy, lipoplex, nonviral vectors, polyplex, transfection

INTRODUCTION

Gene transfer, the technique to introduce new genetic materials to hosts, has become an invaluable experimental tool to study gene function and its regulation, to establish various disease models, to acquire DNA-based immunization, and finally, to explore potential therapeutic applications to various acquired or inherited diseases. Naked DNA molecules do not enter cells efficiently because of their large size and hydrophilic nature due to negatively charged phosphate groups. In addition, they are very susceptible to nuclease-mediated degradation. Therefore, the primary challenge for gene therapy is to develop carriers (commonly called vectors) and physical methods that facilitate gene transfer to targeted cells without degradation of the delivered gene.

Recombinant viruses such as retrovirus, lentivirus, adenovirus, adeno-associated virus, and herpes simplex virus have been widely utilized as vectors for gene transfer (1). Viruses mediate efficient gene transfer through their favorable cell uptake and intracellular trafficking machineries. However, viral vectors have several intrinsic drawbacks including difficulty in production, limited opportunity for repeated administrations due to acute inflammatory response, and delayed humeral or cellular immune responses. Insertional mutagenesis is also a potential issue for some viral vectors that integrate foreign DNA into the genome.

The nonviral gene delivery methods, on the other hand, use synthetic or natural compounds or physical forces to deliver a piece of DNA into a cell. The materials used are generally less toxic and immunogenic than the viral counterparts. In addition, cell or tissue specificity can be achieved by harnessing cell-specific functionality in the design of chemical or biological vectors, while physical procedures can provide spatial precision. Other practical advantages of nonviral approaches include ease of production and the potential for repeat administration. Nonviral methods are generally viewed as less efficacious than the viral methods, and in many cases, the gene expression is short-lived. However, recent developments suggest that gene delivery by some physical methods has reached the efficiency and expression duration that is clinically meaningful. The purpose of this article is to provide an update and concise review in the field of nonviral gene delivery. Particular emphasis will be on the rate-limiting steps that affect the overall transfection and current efforts and strategies to overcome these limitations.

EXTRA- AND INTRACELLULAR BARRIERS FOR GENE DELIVERY

Several anatomical and cellular barriers limit the overall efficiency of gene transfer by nonviral methods (Fig. 1). Anatomical barriers are epithelial, endothelial cell linings and the extracellular matrix surrounding the cells that prevent direct access of macromolecules to the target cells. Professional phagocytes such as Kupffer cells in the liver and residential macrophages in the spleen are largely responsible for the clearance of DNA-loaded colloidal particles administered through blood circulation. In addition, various nucleases existing in blood and extracellular matrix can rapidly degrade free and unprotected nucleic acids following systemic administration. Crossing plasma membrane is considered the most critical limiting step for an efficient DNA transfection. Nucleic acids typically cannot pass through cell membrane unless their entry is facilitated by creating transient holes by physical meanings (2), or through various active cell uptake mechanisms such as endocytosis, pinocytosis, or phagocytosis (3).

Fig. 1.

Fig. 1

Schematic representation of barriers limiting nonviral gene transfer in vivo

Upon being taken up via endocytosis, macromolecules captured within the endosomes usually transform into digestive lysosomes unless some escape mechanisms are used to intercept this maturation process. Two escape mechanisms have been explored. The first involves the use of membrane active or fusogenic molecules such as fusion peptides (4) or lipid components with acid-sensitive bond and large hydrophobic portion of the molecules to rupture the endosome membrane (5). The other mechanism acts on building up osmotic pressure within the endosomal compartment that eventually triggers the swelling or burst of endosomal vesicles (6). Weak amine compounds such as chloroquine and cationic polymers (polyethyleneimines (PEI) and partially degraded polyamidoamine dendrimers) absorb protons and slow down the acidification process that is essential for endosome–lysosome transition (7). Consequently, the influx of chloride counter ions builds up osmotic pressure inside the endosomes. For polyester-based carriers such as poly(lactic-co-glycolic acid), the breakdown products by hydrolysis can also build up the osmotic pressure inside the endosome which leads to the release of the contents trapped therein. Several other attempts have been used to increase the rate of endosome release, among which are codelivery of inactivated viral particles or recombinant viral capsule proteins that possess endosomolytic activities (8), and the use of photochemically generated free radicals to cause membrane damage (9).

Upon their release from endosomes, DNA molecules in their free form or as complexes must travel through cytoplasmic space filled with viscous protein solution and a network of cytoskeleton matrix toward the nucleus where transcription takes place. Observations made by direct intracellular microinjection of naked DNA prove that the movement by diffusion is slow and inefficient, and the resulting levels of gene expression are very weak (10). The nuclear envelope represents an important barrier for the entry of DNA. This double-membrane envelope is interrupted by large protein structures called nuclear pore complexes (NPC) which regulate transport through nuclear envelope. NPC have diameters of ∼9 nm, which allow free diffusion of ions and molecules of small to medium sizes, such as proteins of up to 40–60 kDa, or nucleic acids of up to ∼300 bp, but restrict larger macromolecules passing through freely (11). For resting cells, nuclear uptake of large proteins is an active transport process mediated through sequence-specific recognition of nuclear localization signal peptide (NLS) sequence in their structures by importin proteins (12). Protein-NLS/importin complexes dock at the NPC to allow nuclear entry of DNA. The entry is achieved indirectly through the NLS sequences of transcription factors that are associated with the DNA molecules. For replicating cells, most DNA molecules enter the nuclei through the process of dissolution and reorganization of the nuclear envelope during mitosis (13). Finally, the unpacking of DNA–carrier complexes could constitute yet another rate-limiting step after transfection. Cationic lipids dissociate from DNA through lipid mixing and exchange with host cell lipids at the cytoplasm entry step, while DNA complexes formed with cationic polymers, such as PEI, remain stable after endosome escape. An interesting concept has been reported recently, under which the intracellular trafficking of DNA-loaded nanoparticles is coupled with microtubule-directed transport mechanism (14). The polymer–DNA complexes disintegrate later in nucleus (15).

PERSISTENCE OF GENE EXPRESSION

Generally, the duration of gene expression from plasmid DNA delivered by cationic lipids or polymers is biphasic in nature, with a dramatic, yet transient expression that only lasts for a few days, followed by a prolonged gene expression at a lower level. Several reasons may contribute to the initial rapid decline. The plasmid-based transgene in most cases stays in the nucleus as an episomal DNA molecules without substantial chance to integrate into the host genome. Such molecules in dividing cells do not replicate and will eventually be diluted away as the population of replicating cells grows. Another possibility could be that the recipient cells are killed due to injuries occurred during the transfection process or that transfected cells undergo programmed cell death due to exposure of the transfection agent, or the degradation products of DNA (16). Inflammation response after transfection may also contribute to shortened duration of transgene expression in vivo, either by promoting the clearance of injured cells or through promoter downregulation (17). Immune response to cells expressing foreign proteins may also lead to the elimination of transfected cells. On the other hand, plasmids being physically intact and persistent in the nuclei for a few months to 1 or 2 years have been reported in animals after direct DNA injection in skeletal muscles (18), or through hydrodynamic injection to the liver (19), suggesting that long-term expression from simple plasmids in resting cell populations is possible under certain circumstances. In these cases, promoter shutdown has been suggested as the major cause of a transient nature of gene expression. Certain ubiquitous strong viral promoter sequences such as the cytomegalovirus immediate early promoter/enhancer sequence are silenced over time (20). This could be due to the status of the key transcriptional factors that are most active after cell injury or under an inflammatory response, but not very active at the resting state. Possible gene silencing due to DNA methylation in the promoter region has also been proposed (21). The use of tissue-specific promoters in addition to native or viral scaffold/matrix attachment region has resulted in sustained gene expression (22). The plasmids that provided long-term expression typically have large sizes in comparison to the regular plasmids (23). However, delivery of linearized short DNA fragments has been shown to be beneficial in some studies (24). It was shown that these sequences can be ligated by the host DNA repair mechanisms into large-sized oligomers. Unmethylated CpG sequences, a characteristic of DNA of bacteria origin, could stimulate innate immune response. Plasmids deplete of these CpG sequences (25) or minicircles that contain only the functional part of the plasmid with less CpG contents have been shown to prolong the expression duration (26).

In addition to episomal form of expressive DNA, stable transfection by using DNA integration into genomic DNA has been explored to prolong the duration of gene expression. Retroviral integrase (27), sleeping beauty (28), or phage-derived recombinase (29) have been studied for the use of site-specific integration to the host chromosomes. Favorable low insertion mutation rate has been demonstrated using these systems when compared to retroviral integrases.

PHYSICAL NONVIRAL GENE DELIVERY METHODS

Over the past decade, many physical methods have been investigated for gene transfer. These methods facilitate the transfer of genes from extracellular to nucleus by creating transient membrane holes/defects using physical forces such as local or rapid systemic injection, particle impact, electric pulse, ultrasound, or laser irradiation.

Needle and Jet Injection

Localized injection of naked DNA was first demonstrated intramuscularly in 1990 (30) and afterward in several other tissues, including liver, skin, and brain. DNA uptake is mostly localized in the area where needle track is, suggesting that physical damage induced by needle insertion is responsible in part for the uptake of DNA. Different agents such as transferrin, water-immiscible solvents, nonionic polymers, surfactants, or nuclease inhibitors have been tested to enhance the overall gene expression by this procedure (3134).

From application standpoint, this procedure is particularly attractive because of its simplicity and lack of safety concerns. Direct injection of therapeutic genes into muscle or skin has been a very useful tool for evaluation of various aspects of DNA-based vaccination. Intramuscular injection of plasmid DNAs coding for interleukin-12 and carcinoembryonic antigen gene was able to improve the antitumor immunity (35). A phase I/II clinical trial examining the therapeutic effect of direct myocardial injection of VEGF-2 gene to patients suffering a severe ischemic heart disease has generated some positive results including improved exercise capacity (36). As of March 2009, more than 18.1% (n = 278) of clinical trials in gene therapy were conducted using this method (http://www.wiley.co.uk/genetherapy/clinical/).

The jet injection was first described in 1947 as a needle-free drug delivery method in contrast to the conventional needle injection (37). Jet injection is accomplished through a high-speed, ultrafine stream of DNA solution driven by a pressurized gas, usually CO2. The injection generates pores on membranes of target cells and allows intracellular gene transfer. The penetration power in this method can be controlled by the gas pressure depending on the mechanical properties of the target tissues. The standard procedure for jet injection includes loading of DNA solution (usually 3–5 µl), choosing the strength of the pressure (usually 1–3 bars), aiming the injector directly to the target tissue, and finally, pulling the trigger. Several jet injectors varying in their capacities of injection volume and the modes of repetitive injection have been used for gene transfer. A direct comparison between needle injection and various types of jet injectors showed that the levels of gene expression by jet injection are 50-folds higher than conventional needle injection (38). The models with low-volume injectors are more suitable than those with high-volume injectors in carrying out multiple injections (39).

The jet injection gene transfer is well tolerated by the target tissues, and no serious side effects have been reported. However, localized pain, edema, and bleeding at the injection site have been reported, particularly when old models of injectors were used (40).

The jet injection-based gene transfer is ideal for DNA-based vaccine development and for topical immunization purpose. In addition, this method has been used to directly transfect skin cancer cells to facilitate conventional chemotherapy. Stein et al. have demonstrated that after in vivo intratumoral transfer of short hairpin RNA expression vector against multidrug resistance gene 1 (MDR1), a complete reversal of the phenotype of MDR1 and subsequently enhanced the efficacy of chemotherapy for tumor growth inhibition were achieved (41). Furthermore, jet injection has shown its potential for in vivo gene delivery studies and is currently under investigation in phase I clinical trial for treatment of skin metastases of breast cancer and malignant melanoma (42).

Hydrodynamic Gene Transfer

The hydrodynamic procedure was reported in 1999 (43). When rapid injection of large volume of DNA solution into a mouse via the tail vein was performed, efficient transfection in liver, lung, kidney, and heart was achieved. The hydrodynamic method employs the high pressure as a driving force for gene transfer. The injection of large DNA volume, 8–12% of body weight in short time (3–5 s), leads to a reversible permeability change in the endothelial lining and the generation of transient pores in hepatocyte membranes allowing the DNA molecules to diffuse internally (44). Up to 30–40% of the hepatocytes can be efficiently transfected (45). Currently, this method is considered to be the most efficient nonviral gene transfer method for in vivo gene delivery in rodents. Using this method, it was possible to provide levels of transgene expression close to average levels of physiological gene expression. By using catheter-assisted perfusion, efficient gene transfer can also be achieved in kidney, muscle, or a specific lobe in the liver (46). The simplicity and safety of the hydrodynamic gene delivery allows a wide range of use of this technique for in vivo transfection of hepatocytes to study promoter function, gene function, and therapeutic effects of liver-generated secreted proteins in established disease models (47,48).

Until recently, the idea of employing this procedure for in vivo human gene delivery was ruled out basically due to the fact that proportionally large injection volume should be used that is beyond an acceptable level for patients. Several modifications have been made to make this procedure less invasive. For instance, inserting a balloon catheter into a vasculature in the targeted tissue followed by injection with lower injection speed and volume has resulted in efficient gene transfer in large animals (49). This suggests that clinical application of the hydrodynamic gene transfer is feasible, particularly after the recent development of computer-controlled injection device (50).

Gene Gun

Gene gun delivery, also called ballistic DNA transfer or DNA-coated particle bombardment, was first used in 1987 for gene transfer in plants (51). This method depends on the impact of heavy metal particles on target tissues and delivery of coated DNA on particles in passing. The particles are accelerated to sufficient velocity by highly pressurized inert gas, usually helium. Macroparticles made of gold, tungsten, or silver have been used for gene delivery through gene gun. Gas pressure, particle size, and dosing frequency are critical factors that determine penetration efficiency to the tissues, the degree of tissue injury, and overall gene transfer levels (52). Gene gun-based gene transfer has been extensively tested for intramuscular, intradermal, and intratumor genetic immunization. It was demonstrated that this approach is able to produce more immune response with lower doses comparing to needle injection in large animal models and in clinical human trials. Goudy et al. have shown that vaccination in an animal model, at late preclinical stage of type 1 diabetes, with glutamic acid decarboxylase gene has induced type 2 immunity that consequently resulted in blocking β cell autoimmunity (53). As of March 2009, about 0.3% (n = 5) of clinical trials in gene therapy were conducted using gene gun (http://www.wiley.co.uk/genetherapy/clinical/).

Electroporation

The use of an electric field to alter the cell permeability was known since 1960s. However, the first in vitro and in vivo attempts to utilize electroporation in gene transfer were demonstrated in 1982 (54) and 1991 (55), respectively. In vivo electroporation depends on electric pulses to drive gene transfer. These pulses generated transient pores in cell membranes followed by intracellular electrophoretic DNA movement.

Typically, in vivo electroporation is conducted by first injecting DNA to the target tissue followed by electric pulses, with varied voltage, pulse duration, and number of cycles, from two electrodes applied. In vivo electroporation technique is generally safe, efficient, and can produce good reproducibility compared to other nonviral methods. When parameters are optimized, this method can generate transfection efficiency equal to that achieved by viral vectors (56). In addition to local injection and electroporation, Sakai et al. have reported that in vivo electroporation can be performed in localized manner after a systemic injection of plasmid DNA. They demonstrated that a localized electroporation at particular lobe of rat liver after systemic injection of plasmid DNA resulted in a widespread transfection in hepatocytes in the treated lobe, but not in the surrounding lobes (57). One of the encouraging applications of electroporation was reported recently by Marti et al. who, using the in vivo electroporation, demonstrated an enhanced wound healing after transfection of the affected area with keratinocyte growth factor-1 (58). Despite the progress made recently, limitation of the in vivo electroporation-mediated gene transfer to solid tissues is the accessibility of the electrodes to the internal organs. More basic research and technological developments are likely to speed its clinical applications.

Sonoporation

The first indication that ultrasound might enhance the transdermal penetration of drugs was demonstrated in 1954 (59). Sonoporation, as the term suggested, is a technique that uses ultrasound waves to create plasma membrane defects by acoustic cavitation. With each ultrasonic cycle, a fraction of the energy of the propagating wave is absorbed by the tissue resulting in local heating which affects the structure of cell membranes. Tissue absorption to ultrasound waves depends on tissue type and ultrasound frequency and intensity. Most gene delivery studies use the therapeutic ultrasound at frequency of 1–3 MHz with intensity of 0.5–2.5 W/cm2 (60).

A major improvement in ultrasound-based gene transfer was the combination of ultrasound irradiation with contrast agents or microbubbles (61). Microbubbles are air-filled vesicles stabilized by surface active molecules such as albumin, polymers, or phospholipids. Upon absorption of ultrasound waves, microbubbles cavitate, oscillate, break up, and release local shock waves in the form of high-velocity microjet that can disrupt the nearby cell membranes. This promotes the transient pore formation which, consequently, facilitates local DNA transfer. One of the most commonly used contrast agent is Optison (Molecular Biosystems, San Diego, CA, USA) which consists of gas-filled human albumin microspheres. The size of microbubbles, usually 1–6 μm in diameter, is critical for the efficient transfection and for not being removed by the reticular endothelial system. Smaller-sized nanobubbles have been also employed but were found to require higher frequency ultrasound exposure to be rendered effective. Modification of microbubbles through lipid or polymer coating resulted in enhanced transfection efficiency (62). In general, the efficiency of sonoporation-based gene transfer is dependent on the frequency and intensity of ultrasound irradiation, the presence of contrast agent, DNA concentration, and the duration of exposure. However, enhancement of fluidity of the cell membrane by feeding cells with long-chain unsaturated fatty acids, which facilitates its flexibility and minimizes cellular resistance to sonication, was also suggested to improve the effectiveness of sonoporation (63). The major advantage for sonoporation is its safety, noninvasiveness, and being able to reach internal organs without surgical procedure. Recently, ultrasound has been shown to be able to enhance the permeability of blood–brain barrier (64). Interestingly, targeted gene delivery can be achieved through focused sonoporation using nontargeted microbubbles or through microbubbles equipped with site-specific ligands, such as antibodies or biotin–streptavidin that helps in transferring of microbubbles to certain tissue or organ. Tsunoda et al. demonstrated successful in vivo gene transfer to injured myocardial tissues using sonoporation after an intraventricular injection (65). Sheyn et al. have found the use of ultrasound-mediated gene therapy to facilitate bone tissue regeneration and subsequently bone formation upon transfer of recombinant human bone morphogenetic protein-9 (66).

CHEMICAL-BASED NONVIRAL VECTORS

Chemical vectors such as cationic lipids and cationic polymers form condensed complexes with negatively charged DNA through electrostatic interactions. The complexes protect DNA and facilitate cell uptake and intracellular delivery. Cationic lipids and cationic polymers as gene delivery tools have been well studied, and the subject has been covered by many review articles in great details (67). Below is a brief summary of chemical vectors employed for gene transfer.

Cationic Lipids

Felgner and colleagues pioneered cationic lipid-based gene transfer in 1987 (68). Cationic liposome-mediated gene transfer or lipofection represents the most extensively investigated and commonly used nonviral gene delivery method. Currently, hundreds of lipids have been developed and tested for gene transfer. They share the common structure of positively charged hydrophilic head and hydrophobic tail that are connected via a linker structure. The most commonly seen hydrophilic head groups are primary, secondary, tertiary amines, or quaternary ammonium salts. However, guanidino, imidazole, pyridinium, phosphorus, and arsenic groups have also been developed. The positively charged head group is necessary for binding with negatively charged phosphate groups in nucleic acids. The hydrophobic tails are usually made of two types of hydrophobic moieties, aliphatic chains, cholesterol, or other types of steroid rings. Most of the linkage between the hydrophilic and hydrophobic moiety are ether, ester, carbamate, or amide bonds that can affect the rate of biodegradability. Table I summarizes some cationic lipids used in gene transfer. Transfection efficiency of cationic lipids varies dramatically depending on the structure of cationic lipids (the overall geometric shape, the number of charged groups per molecules, the nature of lipid anchor, and linker bondages), the charge ratio used to form DNA–lipid complexes, and the properties of the colipid (69). When mixed with the negatively charged DNA, the positive-charged liposomes spontaneously form uniquely compacted structures called lipoplexes. In lipoplex structure, DNA molecules are surrounded with positively charged lipids which grant them protection against extracellular or intracellular nucleases. Furthermore, lipoplexes, due to their positive charges, tend to electrostatically interact with the negatively charged molecules of the cell membrane (glycoproteins and proteoglycans) that may facilitate their cellular uptake.

Table I.

Lipids Commonly Used for Gene Transfer

Lipid Abbreviation Feature
1,2-Dioleoyl-sn-glycero-3-phosphatidylcholine DOPC Helper
1,2-Dioleoyl-sn-glycero-3-phosphatidylethanolamine DOPE Helper
Cholesterol Helper
N-[1-(2,3-Dioleyloxy)propyl]N,N,N-trimethylammonium chloride DOTMA Cationic
1,2-Dioleoyloxy-3-trimethylammonium-propane DOTAP Cationic
Dioctadecylamidoglycylspermine DOGS Cationic
N-(3-Aminopropyl)-N,N-dimethyl-2,3-bis(dodecyloxy)-1-propanaminium bromide GAP-DLRIE Cationic
Cetyltrimethylammonium bromide CTAB Cationic
6-Lauroxyhexyl ornithinate LHON Cationic
1-(2,3-Dioleoyloxypropyl)-2,4,6-trimethylpyridinium 2Oc Cationic
2,3-Dioleyloxy-N-[2(sperminecarboxamido)-ethyl]-N,N-dimethyl-1-propanaminium trifluoroacetate DOSPA Cationic
1,2-Dioleyl-3-trimethylammonium-propane DOPA Cationic
N-(2-Hydroxyethyl)-N,N-dimethyl-2,3-bis(tetradecyloxy)-1-propanaminium bromide MDRIE Cationic
Dimyristooxypropyl dimethyl hydroxyethyl ammonium bromide DMRI Cationic
3β-[N-(N′,N′-Dimethylaminoethane)-carbamoyl]cholesterol DC-Chol Cationic
Bis-guanidium-tren-cholesterol BGTC Cationic
1,3-Dioleoxy-2-(6-carboxy-spermyl)-propylamide DOSPER Cationic
Dimethyloctadecylammonium bromide DDAB Cationic
Dioctadecylamidoglicylspermidin DSL Cationic
rac-[(2,3-Dioctadecyloxypropyl)(2-hydroxyethyl)]-dimethylammonium chloride CLIP-1 Cationic
rac-[2(2,3-Dihexadecyloxypropyl-oxymethyloxy)ethyl]trimethylammonium bromide CLIP-6 Cationic
Ethyldimyristoylphosphatidylcholine EDMPC Cationic
1,2-Distearyloxy-N,N-dimethyl-3-aminopropane DSDMA Cationic
1,2-Dimyristoyl-trimethylammoniumpropane DMTAP Cationic
O,O′-Dimyristyl-N-lysyl aspartate DMKE Cationic
1,2-Distearoyl-sn-glycero-3-ethylphosphocholine DSEPC Cationic
N-Palmitoyl d-erythro-sphingosyl carbamoyl-spermine CCS Cationic
N-t-Butyl-N0-tetradecyl-3-tetradecylaminopropionamidine diC14-amidine Cationic
Octadecenolyoxy[ethyl-2-heptadecenyl-3 hydroxyethyl] imidazolinium chloride DOTIM Cationic
N1-Cholesteryloxycarbonyl-3,7-diazanonane-1,9-diamine CDAN Cationic
2-(3-[Bis-(3-amino-propyl)-amino]propylamino)-N-ditetradecylcarbamoylme-ethyl-acetamide RPR209120 Cationic

The most commonly used colipids are cholesterol and dioleoylphosphatidylethanolamine (DOPE). The contribution of these “helper” lipids on cationic liposome-mediated gene transfer to a large extent depends on the structure of cationic lipid. Some lipids absolutely require DOPE for an appreciate level of transfection, while some cationic lipids, particularly those with double fatty chains capable of forming bilayer or micellar structures, have transfection activity without any helper lipids. For the latter one, the presence of DOPE in most cases reduces the charge ratio of lipid to DNA needed to reach the maximal transfection in vitro in the absence of serum. Most cationic lipids are more or less toxic to cells, and the presence of DOPE could lead to reduced charge ratio, thus less toxicity. Cholesterol-containing cationic liposomes were developed for in vivo application. The presence of cholesterol stabilizes the cationic lipidic membrane structures against the destructive activity of serum components and can provide better activity for in vivo transfection when serum components are present (70).

The choice of a colipid has a quite dramatic effect on the overall performance of cationic liposomes. In general, a lipid composition that provides high degree of membrane fluidity allows better transfection. It has been found that DOPE as a colipid promotes hexagonal phase lipid polymorphism that is in favor of membrane fusion, lipid mixing, and boost of transfection efficiency in vitro (71). The fusogenic properties of DOPE facilitate the endosomal escape of lipoplexes through membrane destabilization. In addition, DOPE induces the displacement of the anionic lipids from the cytoplasm-facing monolayer of the endosomal membrane to the opposite direction via a flip-flop mechanism (5,72). Displaced anionic lipids then interact with the endosomally engaged cationic lipids causing the formation of neutral lipids that subsequently help in decomplexation of the DNA from the cationic lipid leading to the release of the transgene in the cytoplasm as a free or as a lipoplex. Although DOPE-containing formulations are among the best for in vitro transfection in the absence of serum, the serum components are known to inactivate and destabilize the lipoplex structures that contain DOPE (73).

Due to the excessive surface charge, the circulation half-life of cationic lipids in blood is very short. Systemic elimination of cationic lipids takes place upon formation of larger aggregates via their interactions with the negatively charged serum molecules or cellular components (primarily erythrocytes and platelets). Lipoplexes tend to initially accumulate in the pulmonary vasculature because of the so-called first passage effect (74). However, the lipoplexes redistribute to the liver in about 60 min after the injection owing to active uptake by Kupffer cells. As a result, pulmonary vascular endothelial cells and some airway epithelial cells are predominant cells that are transfected (75). The drawback of a fast clearance of cationic lipids from the circulation limits their utility in gene delivery to cells located beyond vascular endothelial cells. Surface shielding through the use of hydrophilic and charge neutral polymers such as polyethylene glycol (PEG) to reduce excessive charge–charge interaction appears very effective in prolonging the circulation half-life of lipoplexes (76).

It is important to note that the presence of the bulky PEG moiety on the surface prevents an intimate interaction between lipoplexes and cell membrane, thus reduces the overall transfection efficiency. Several strategies have been used to make the PEG shielding conditional and nonpermanent. PEG–lipid conjugates with single lipid anchor of a shorter fatty chain length have been used to make these molecules diffusible over time to achieve a conditional deshielding by kinetic means. Various forms of pH-sensitive or reduction-sensitive chemical bonds have been developed to allow the deshielding within acidified endosome compartments.

Prolonged circulation time due to surface modification makes targeted gene delivery to interstitially located cell populations possible. Successfully targeted delivery of DNA and siRNA to tumor cells using lipoplexes coated with antitransferrin receptor monoclonal antibodies to bind to overexpressed transferrin receptors has been reported (77). Gene transfer of anticancer therapeutic genes, either with tumor-suppresser or enzyme-directed prodrug or by siRNA-based gene knockdown approaches, has been explored as a possible therapeutic intervention for cancer gene therapy. In addition, monoclonal antibody-mediated transcytosis for lipoplex-based brain gene delivery has also been studied for gene therapy of neuronal diseases and brain tumor (78).

In addition to systemic application, significant efforts have been made for topical and regional gene and siRNA delivery with lipoplexes to respiratory mucosa and airway epithelial cells for the treatment of cystic fibrosis, acute lung injuries (79), as well as to cornea tissues for degenerative ocular diseases (80).

Major hurdles for practical use of lipoplex-mediated transfection are acute toxicity and short duration of gene expression. Cationic lipids, when combined with unmethylated CpG-containing plasmid DNA or other nucleic acid compositions, can stimulate potent inflammatory response in the hosts. Rapid production of cytokines followed by clearance of transfected cells after the administration of uncoated DNA lipoplexes contributes to both treatment-related toxicity and shortened duration of gene expression. In some extreme cases, a subsequent transfection of the same vectors is less effective because of the cytokine effect from previous transfection. Systematic deletions and mutations of CpG sequences from plasmid sequences have generated some promising data in suppressing the levels of cytokine production after the administration of lipoplexes (81). Surface shielding with PEG–lipid can also minimize the inflammatory response (76,82).

In general, cationic lipids have the advantages of being inexpensive to produce and can be engineered to have targeted specificity. However, their transfection efficiency needs to be further improved, and the significant toxicities such as formation of aggregates in blood and the tendency to induce inflammatory response have to be solved for in vivo application. As of March 2009, lipoplexes have been used in clinical trials and represent 7.1% (n = 109) of total human gene therapy trials mainly for cancer and cystic fibrosis studies (http://www.wiley.co.uk/genetherapy/clinical/).

Cationic Polymers

Cationic polymers have also been used extensively for gene transfer. Upon mixing with DNA, these polymers form nanosized complexes, often called polyplexes. Typically, polyplexes are more stable than lipoplexes. Table II summarizes some polymers that are commonly used in gene delivery.

Table II.

Polymers Commonly Used for Gene Transfer

Polymer Abbreviation Unique feature
Poly(ethylene)glycol PEG Inert
Polyethylenimine PEI Cationic
Dithiobis(succinimidylpropionate) DSP Biodegradable PEI
Dimethyl-3,3′-dithiobispropionimidate DTBP Biodegradable PEI
Poly(ethylene imine) biscarbamate PEIC Biodegradable PEI
Poly(l-lysine) PLL Cationic
Histidine modified PLL Biodegradable
Poly(N-vinylpyrrolidone) PVP Neutral
Poly(propylenimine) PPI Dendromer
Poly(amidoamine) PAMAM Dendromer
Poly(amido ethylenimine) SS-PAEI Biodegradable
Triethylenetetramine TETA Cationic
Poly(β-aminoester) Biodegradable
Poly(4-hydroxy-l-proline ester) PHP Biodegradable
Poly(allylamine) Cationic
Poly(α-[4-aminobutyl]-l-glycolic acid) PAGA Biodegradable
Poly(d,l-lactic-co-glycolic acid) PLGA Biodegradable
Poly(N-ethyl-4-vinylpyridinium bromide) Cationic
Poly(phosphazene)s PPZ Biodegradable
Poly(phosphoester)s PPE Biodegradable
Poly(phosphoramidate)s PPA Biodegradable
Poly(N-2-hydroxypropylmethacrylamide) pHPMA Cationic
Poly (2-(dimethylamino)ethyl methacrylate) pDMAEMA Cationic
Poly(2-aminoethyl propylene phosphate) PPE-EA Biodegradable
Chitosan Polysaccharide
Galactosylated chitosan Synthetic chitosan
N-Dodecylated chitosan Synthetic chitosan
Histone Natural
Collagen Natural
Dextran–spermine D-SPM Polysaccharide

Among cationic polymers, PEI is considered one of the most effective polymer-based transfection agents. PEI was first used in gene transfer in 1995 (83). It exists in either branched or linear structures. PEI has a high density of amine groups of which majority are nonprotonated at the physiological pH. The nonprotonated amines in the PEI exert the so-called proton sponge effect, which effectively stops the acidification of the endosomal pH by neutralizing the protons that are pumped by an active membrane transporter, ATPase (6,84). Ultimately, it leads to an influx of chloride counter ions within the compartment and a buildup of osmotic pressure that causes the swelling and rupture of the endosomal membrane.

Transfection efficiency and toxicity of PEI depends on its molecular weight (MW), configuration, and the charge ratio of polymer to DNA used. Several studies showed that high MW PEI (greater than 25,000 Da) is toxic to cells while polymers with medium to low MW (5,000–25,000 Da) are more efficient and less toxic (85). Synthetic, high MW PEI derived from low MW PEIs by polymerization either through biodegradable disulfide linkages or cross-linked with inert polymers, particularly those biodegradable polymers, have been shown to be more efficacious and less toxic than the PEI of the same MW (86). Polyplexes between DNA and linear PEI are active in vivo when administrated intravenously (87). The degree of free amine content, MW, and the polymer-to-DNA ratios and the solution used to prepare the polyplexes are important factors for in vivo transfection. Branched PEI, on the other hand, has high toxicity and low transfection efficiency than the polyplexes prepared from linear PEI (88). However, they are useful for forming stable nanoparticles that are suitable for airway gene delivery through aerosol for treating airway epithelial cells or lung cancers (89).

Upon systemic administration, these polyplexes of small particle size tend to aggregate to form larger complexes and accumulate in major tissues including lung and liver. The levels of cytokine induction by polyplexes appear less severe than that of lipoplexes (90). Additional improvement can be achieved through conjugation to an inert polymer such as PEG (5,000 Da), pluronic triblock polymers (P127), or dextran to reduce the nonspecific interactions. Nanoparticles coformulated from biodegradable anionic polymers such as albumin, dextran sulfate, or other anionic polymers to modify the surface charge of the polyplexes are another interesting idea to overcome the toxicity and nonspecific binding issues. Along the same direction, encapsulating the polyplexes inside neutral or anionic liposomes, solid biodegradable nanoparticles, or polymer-based hydrogels has shown promising improvement in either enhancing the transfection activity or reducing cytotoxicity of PEI-based polyplexes.

Recently, more polymers with improved biocompatibility and biodegradability have been reported demonstrating equal or superior performance comparing to nondegradable PEIs. Among these are aminoesters or oligoamines polymerized through disulfide linkages or polyamino acid derivatives with proton absorption capacities (91). Besides PEI and more recent polyamines of varied structures, synthetic or natural polypeptides and their derivatives have been explored as gene delivery vehicles. These include poly(l-lysine) (PLL), polyornithine, polyarginine, histones, and protamines that have excellent ability to condense DNA.

PLL is among the first synthetic polymers being used for constructing target-specific gene carrier by Wu and Wu and several other groups for liver, lung, and tumor-specific gene delivery (92). Recently, PEG–PLL conjugates with defined chemical composition have been shown to be improved vector for DNA formulation and delivery to various organs (93). Such polyplexes are being investigated in phase I and phase II clinical trials for potential treatment of cystic fibrosis and ocular degenerative diseases (94). Improved transfection has been shown using PLL with dendritic configuration and imidazole modified linear PLL (95). Other polymers such as dendromers, chitosans, synthetic amino derivatives of dextran, and cationic acrylic polymers have been shown to possess significant levels of gene transfer activity (96).

Inorganic Nanoparticles

Inorganic nanoparticles are usually prepared from metals (e.g., iron, gold, silver), inorganic salts, or ceramics (e.g., phosphate or carbonate salts of calcium, magnesium, or silicon) (97). The metal ion-based salts produce complexes with typical size range of 10–100 nm in diameter. The surface of these nanoparticles can be coated to facilitate DNA binding or targeted gene delivery. The small particle size offers several advantages including that they usually bypass most of the physiological and cellular barriers and produce higher gene expression (98). They can also be transported through the cellular membranes via specific membrane receptor or nucleolin which delivers nanoparticles directly to the nucleus skipping the endosomal–lysosomal degradation (99). Nanoparticles have the ability to efficiently transfect postmitotic cells in vivo and in vitro. Additionally, they tend to show no or low toxicity and are inert to immune responses. Supraparamagnetic iron oxide-based nanoparticles can also provide magnetic responsiveness in a magnetic field and can provide magnetic field guided targeted delivery (100). Progress in the in vivo application of the inorganic nanoparticles has picked up considerable speed recently. However, extensive studies are still required to assess the effect of their types, sizes, and shapes on transfection efficiency. It is certain that further studies focusing on long-term safety and surface functionalization can accelerate their clinical applications.

FUTURE PERSPECTIVES

Nonviral approaches were developed to facilitate transfer of exogenous genes into target cells without the complication of immunogenicity or insertion mutation commonly seen in viral vectors. These methods differ widely in their transfection efficiency and toxicity. In the past few years, the work continued in developing new nonviral methods, particularly in the area of chemical vectors. However, the last few years showed broad successful applications of the physical methods for in vivo gene transfer. As a whole, the transfection efficiency reported so far for the nonviral approaches is still below that of the highly efficient viral vectors. Further improvements to increase the efficiency and reduce the toxicity of nonviral vectors are needed before their clinical implication can be met. These improvement will rely on our better understanding of the limiting steps that nonviral vector must overcome. Developing new vectors that are more target specific will also be necessary. The strategies that merge nonviral and viral vectors might be helpful to achieve more, efficient, long-lasting, and nontoxic gene delivery systems.

Acknowledgments

We are grateful to Dr. Fowzan S. Alkuraya for his critical reading of the manuscript.

References

  • 1.Walther W, Stein U. Viral vectors for gene transfer: a review of their use in the treatment of human diseases. Drugs. 2000;60:249–271. doi: 10.2165/00003495-200060020-00002. [DOI] [PubMed] [Google Scholar]
  • 2.Villemejane J, Mir LM. Physical methods of nucleic acid transfer: general concepts and applications. Br J Pharmacol. 2009;157:207–219. doi: 10.1111/j.1476-5381.2009.00032.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Medina-Kauwe LK, Xie J, Hamm-Alvarez S. Intracellular trafficking of nonviral vectors. Gene Ther. 2005;12:1734–1751. doi: 10.1038/sj.gt.3302592. [DOI] [PubMed] [Google Scholar]
  • 4.Li W, Nicol F, Szoka FC., Jr GALA: a designed synthetic pH-responsive amphipathic peptide with applications in drug and gene delivery. Adv Drug Deliv Rev. 2004;23:967–985. doi: 10.1016/j.addr.2003.10.041. [DOI] [PubMed] [Google Scholar]
  • 5.Xu Y, Szoka FC., Jr Mechanism of DNA release from cationic liposome/DNA complexes used in cell transfection. Biochemistry. 1996;35:5616–5623. doi: 10.1021/bi9602019. [DOI] [PubMed] [Google Scholar]
  • 6.Akinc A, Thomas M, Klibanov AM, Langer R. Exploring polyethylenimine-mediated DNA transfection and the proton sponge hypothesis. J Gene Med. 2005;7:657–663. doi: 10.1002/jgm.696. [DOI] [PubMed] [Google Scholar]
  • 7.Sonawane ND, Szoka FC, Jr, Verkman AS. Chloride accumulation and swelling in endosomes enhances DNA transfer by polyamine–DNA polyplexes. J Biol Chem. 2003;278:44826–44831. doi: 10.1074/jbc.M308643200. [DOI] [PubMed] [Google Scholar]
  • 8.Curiel DT, Agarwal S, Wagner E, Cotten M. Adenovirus enhancement of transferrin–polylysine-mediated gene delivery. Proc Natl Acad Sci USA. 1991;88:8850–8854. doi: 10.1073/pnas.88.19.8850. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Kloeckner J, Prasmickaite L, Hogset A, Berg K, Wagner E. Photochemically enhanced gene delivery of EGF receptor-targeted DNA polyplexes. J Drug Target. 2004;12:205–213. doi: 10.1080/10611860410001723090. [DOI] [PubMed] [Google Scholar]
  • 10.Lukacs GL, Haggie P, Seksek O, Lechardeur D, Freedman N, Verkman AS. Size-dependent DNA mobility in cytoplasm and nucleus. J Biol Chem. 2000;275:1625–1629. doi: 10.1074/jbc.275.3.1625. [DOI] [PubMed] [Google Scholar]
  • 11.Bastos R, Pante N, Burke B. Nuclear pore complex proteins. Int Rev Cytol. 1995;162B:257–302. doi: 10.1016/s0074-7696(08)62619-4. [DOI] [PubMed] [Google Scholar]
  • 12.Wente SR. Gatekeepers of the nucleus. Science. 2000;288:1374–1377. doi: 10.1126/science.288.5470.1374. [DOI] [PubMed] [Google Scholar]
  • 13.Dean DA, Strong DD, Zimmer WE. Nuclear entry of nonviral vectors. Gene Ther. 2005;12:881–890. doi: 10.1038/sj.gt.3302534. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Bachand M, Trent AM, Bunker BC, Bachand GD. Physical factors affecting kinesin-based transport of synthetic nanoparticle cargo. J Nanosci Nanotechnol. 2005;5:718–722. doi: 10.1166/jnn.2005.112. [DOI] [PubMed] [Google Scholar]
  • 15.Chen HH, Ho YP, Jiang X, Mao HQ, Wang TH, Leong KW. Quantitative comparison of intracellular unpacking kinetics of polyplexes by a model constructed from quantum dot-FRET. Mol Ther. 2008;16:324–332. doi: 10.1038/sj.mt.6300392. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Nguyen LT, Atobe K, Barichello JM, Ishida T, Kiwada H. Complex formation with plasmid DNA increases the cytotoxicity of cationic liposomes. Biol Pharm Bull. 2007;30:751–757. doi: 10.1248/bpb.30.751. [DOI] [PubMed] [Google Scholar]
  • 17.Yew NS, Wang KX, Przybylska M, Bagley RG, Stedman M, Marshall J, et al. Contribution of plasmid DNA to inflammation in the lung after administration of cationic lipid:pDNA complexes. Hum Gene Ther. 1999;10:223–234. doi: 10.1089/10430349950019011. [DOI] [PubMed] [Google Scholar]
  • 18.Wolff JA, Ludtke JJ, Acsadi G, Williams P, Jani A. Long-term persistence of plasmid DNA and foreign gene expression in mouse muscle. Hum Mol Genet. 1992;1:363–369. doi: 10.1093/hmg/1.6.363. [DOI] [PubMed] [Google Scholar]
  • 19.Al-Dosari MS, Knapp JE, Liu D. Hydrodynamic delivery. Adv Genet. 2005;54:65–82. doi: 10.1016/S0065-2660(05)54004-5. [DOI] [PubMed] [Google Scholar]
  • 20.Loser P, Jennings GS, Strauss M, Sandig V. Reactivation of the previously silenced cytomegalovirus major immediate-early promoter in the mouse liver: involvement of NF-kappa B. J Virol. 1998;72:180–190. doi: 10.1128/jvi.72.1.180-190.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Newell-Price J, Clark AJ, King P. DNA methylation and silencing of gene expression. Trends Endocrinol Metab. 2000;11:142–148. doi: 10.1016/S1043-2760(00)00248-4. [DOI] [PubMed] [Google Scholar]
  • 22.Argyros O, Wong SP, Niceta M, Waddington SN, Howe SJ, Coutelle C, et al. Persistent episomal transgene expression in liver following delivery of a scaffold/matrix attachment region containing non-viral vector. Gene Ther. 2008;15:1593–1605. doi: 10.1038/gt.2008.113. [DOI] [PubMed] [Google Scholar]
  • 23.Hibbitt OC, Harbottle RP, Waddington SN, Bursill CA, Coutelle C, Channon KM, et al. Delivery and long-term expression of a 135 kb LDLR genomic DNA locus in vivo by hydrodynamic tail vein injection. J Gene Med. 2007;9:488–497. doi: 10.1002/jgm.1041. [DOI] [PubMed] [Google Scholar]
  • 24.Chen ZY, Yant SR, He CY, Meuse L, Shen S, Kay MA. Linear DNAs concatemerize in vivo and result in sustained transgene expression in mouse liver. Mol Ther. 2001;3:403–410. doi: 10.1006/mthe.2001.0278. [DOI] [PubMed] [Google Scholar]
  • 25.Hodges BL, Taylor KM, Joseph MF, Bourgeois SA, Scheule RK. Long-term transgene expression from plasmid DNA gene therapy vectors is negatively affected by CpG dinucleotides. Mol Ther. 2004;10:269–278. doi: 10.1016/j.ymthe.2004.04.018. [DOI] [PubMed] [Google Scholar]
  • 26.Chen ZY, He CY, Ehrhardt A, Kay MA. Minicircle DNA vectors devoid of bacterial DNA result in persistent and high-level transgene expression in vivo. Mol Ther. 2003;8:495–500. doi: 10.1016/S1525-0016(03)00168-0. [DOI] [PubMed] [Google Scholar]
  • 27.Tanaka AS, Tanaka M, Komuro K. A highly efficient method for the site-specific integration of transfected plasmids into the genome of mammalian cells using purified retroviral integrase. Gene. 1998;216:67–76. doi: 10.1016/S0378-1119(98)00312-6. [DOI] [PubMed] [Google Scholar]
  • 28.Yant SR, Meuse L, Chiu W, Ivics Z, Izsvak Z, Kay MA. Somatic integration and long-term transgene expression in normal and haemophilic mice using a DNA transposon system. Nat Genet. 2000;25:35–41. doi: 10.1038/75568. [DOI] [PubMed] [Google Scholar]
  • 29.Calos MP. The phi C31 integrase system for gene therapy. Curr Gene Ther. 2006;6:633–645. doi: 10.2174/156652306779010642. [DOI] [PubMed] [Google Scholar]
  • 30.Wolff JA, Malone RW, Williams P, Chong W, Acsadi G, Jani A, et al. Direct gene transfer into mouse muscle in vivo. Science. 1990;247:1465–1468. doi: 10.1126/science.1690918. [DOI] [PubMed] [Google Scholar]
  • 31.Sato Y, Yamauchi N, Takahashi M, Sasaki K, Fukaura J, Neda H, et al. In vivo gene delivery to tumor cells by transferrin–streptavidin–DNA conjugate. Faseb J. 2000;14:2108–2118. doi: 10.1096/fj.99-1052com. [DOI] [PubMed] [Google Scholar]
  • 32.Desigaux L, Gourden C, Bello-Roufai M, Richard P, Oudrhiri N, Lehn P, et al. Nonionic amphiphilic block copolymers promote gene transfer to the lung. Hum Gene Ther. 2005;16:821–829. doi: 10.1089/hum.2005.16.821. [DOI] [PubMed] [Google Scholar]
  • 33.Freeman DJ, Niven RW. The influence of sodium glycocholate and other additives on the in vivo transfection of plasmid DNA in the lungs. Pharm Res. 1996;13:202–209. doi: 10.1023/A:1016078728202. [DOI] [PubMed] [Google Scholar]
  • 34.Glasspool-Malone J, Malone RW. Marked enhancement of direct respiratory tissue transfection by aurintricarboxylic acid. Hum Gene Ther. 1999;10:1703–1713. doi: 10.1089/10430349950017707. [DOI] [PubMed] [Google Scholar]
  • 35.Song K, Chang Y, Prud'homme GJ. Regulation of T-helper-1 versus T-helper-2 activity and enhancement of tumor immunity by combined DNA-based vaccination and nonviral cytokine gene transfer. Gene Ther. 2000;7:481–492. doi: 10.1038/sj.gt.3301123. [DOI] [PubMed] [Google Scholar]
  • 36.Losordo DW, Vale PR, Hendel RC, Milliken CE, Fortuin FD, Cummings N, et al. Phase 1/2 placebo-controlled, double-blind, dose-escalating trial of myocardial vascular endothelial growth factor 2 gene transfer by catheter delivery in patients with chronic myocardial ischemia. Circulation. 2002;105:2012–2018. doi: 10.1161/01.CIR.0000015982.70785.B7. [DOI] [PubMed] [Google Scholar]
  • 37.Wendell DM, Hemond BD, Hogan NC, Taberner AJ, Hunter IW. The effect of jet parameters on jet injection. Conf Proc IEEE Eng Med Biol Soc. 2006;1:5005–5008. doi: 10.1109/IEMBS.2006.260369. [DOI] [PubMed] [Google Scholar]
  • 38.Ren S, Li M, Smith JM, DeTolla LJ, Furth PA. Low-volume jet injection for intradermal immunization in rabbits. BMC Biotechnol. 2002;2:1–6. doi: 10.1186/1472-6750-2-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Cartier R, Ren SV, Walther W, Stein U, Lewis A, Schlag PM, et al. In vivo gene transfer by low-volume jet injection. Anal Biochem. 2000;282:262–265. doi: 10.1006/abio.2000.4619. [DOI] [PubMed] [Google Scholar]
  • 40.Lysakowski C, Dumont L, Tramer MR, Tassonyi E. A needle-free jet-injection system with lidocaine for peripheral intravenous cannula insertion: a randomized controlled trial with cost-effectiveness analysis. Anesth Analg. 2003;96:215–219. doi: 10.1097/00000539-200301000-00044. [DOI] [PubMed] [Google Scholar]
  • 41.Stein U, Walther W, Stege A, Kaszubiak A, Fichtner I, Lage H. Complete in vivo reversal of the multidrug resistance phenotype by jet-injection of anti-MDR1 short hairpin RNA-encoding plasmid DNA. Mol Ther. 2008;16:178–186. doi: 10.1038/sj.mt.6300304. [DOI] [PubMed] [Google Scholar]
  • 42.Walther W, Siegel R, Kobelt D, Knosel T, Dietel M, Bembenek A, et al. Novel jet-injection technology for nonviral intratumoral gene transfer in patients with melanoma and breast cancer. Clin Cancer Res. 2008;14:7545–7553. doi: 10.1158/1078-0432.CCR-08-0412. [DOI] [PubMed] [Google Scholar]
  • 43.Liu F, Song Y, Liu D. Hydrodynamics-based transfection in animals by systemic administration of plasmid DNA. Gene Ther. 1999;6:1258–1266. doi: 10.1038/sj.gt.3300947. [DOI] [PubMed] [Google Scholar]
  • 44.Zhang G, Gao X, Song YK, Vollmer R, Stolz DB, Gasiorowski JZ, et al. Hydroporation as the mechanism of hydrodynamic delivery. Gene Ther. 2004;11:675–682. doi: 10.1038/sj.gt.3302210. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Zhang G, Song YK, Liu D. Long-term expression of human alpha1-antitrypsin gene in mouse liver achieved by intravenous administration of plasmid DNA using a hydrodynamics-based procedure. Gene Ther. 2000;7:1344–1349. doi: 10.1038/sj.gt.3301229. [DOI] [PubMed] [Google Scholar]
  • 46.Eastman SJ, Baskin KM, Hodges BL, Chu Q, Gates A, Dreusicke R, et al. Development of catheter-based procedures for transducing the isolated rabbit liver with plasmid DNA. Hum Gene Ther. 2002;13:2065–2077. doi: 10.1089/10430340260395910. [DOI] [PubMed] [Google Scholar]
  • 47.Suda T, Liu D. Hydrodynamic gene delivery: its principles and applications. Mol Ther. 2007;15:2063–2069. doi: 10.1038/sj.mt.6300314. [DOI] [PubMed] [Google Scholar]
  • 48.Al-Dosari MS, Knapp JE, Liu D. Activation of human CYP2C9 promoter and regulation by CAR and PXR in mouse liver. Mol Pharm. 2006;3:322–328. doi: 10.1021/mp0500824. [DOI] [PubMed] [Google Scholar]
  • 49.Fabre JW, Grehan A, Whitehorne M, Sawyer GJ, Dong X, Salehi S, et al. Hydrodynamic gene delivery to the pig liver via an isolated segment of the inferior vena cava. Gene Ther. 2008;15:452–462. doi: 10.1038/sj.gt.3303079. [DOI] [PubMed] [Google Scholar]
  • 50.Suda T, Suda K, Liu D. Computer-assisted hydrodynamic gene delivery. Mol Ther. 2008;16:1098–1104. doi: 10.1038/mt.2008.66. [DOI] [PubMed] [Google Scholar]
  • 51.Klein RM, Wolf ED, Wu R, Sanford JC. High-velocity microprojectiles for delivering nucleic acids into living cells. Biotechnology. 1992;24:384–386. [PubMed] [Google Scholar]
  • 52.Uchida M, Natsume H, Kobayashi D, Sugibayashi K, Morimoto Y. Effects of particle size, helium gas pressure and microparticle dose on the plasma concentration of indomethacin after bombardment of indomethacin-loaded poly-L-lactic acid microspheres using a Helios gun system. Biol Pharm Bull. 2002;25:690–693. doi: 10.1248/bpb.25.690. [DOI] [PubMed] [Google Scholar]
  • 53.Goudy KS, Wang B, Tisch R. Gene gun-mediated DNA vaccination enhances antigen-specific immunotherapy at a late preclinical stage of type 1 diabetes in nonobese diabetic mice. Clin Immunol. 2008;129:49–57. doi: 10.1016/j.clim.2008.06.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Neumann E, Schaefer-Ridder M, Wang Y, Hofschneider PH. Gene transfer into mouse lyoma cells by electroporation in high electric fields. Embo J. 1982;1:841–845. doi: 10.1002/j.1460-2075.1982.tb01257.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Titomirov AV, Sukharev S, Kistanova E. In vivo electroporation and stable transformation of skin cells of newborn mice by plasmid DNA. Biochim Biophys Acta. 1991;1088:131–134. doi: 10.1016/0167-4781(91)90162-f. [DOI] [PubMed] [Google Scholar]
  • 56.Andre F, Mir LM. DNA electrotransfer: its principles and an updated review of its therapeutic applications. Gene Ther. 2004;11:S33–S42. doi: 10.1038/sj.gt.3302367. [DOI] [PubMed] [Google Scholar]
  • 57.Sakai M, Nishikawa M, Thanaketpaisarn O, Yamashita F, Hashida M. Hepatocyte-targeted gene transfer by combination of vascularly delivered plasmid DNA and in vivo electroporation. Gene Ther. 2005;12:607–616. doi: 10.1038/sj.gt.3302435. [DOI] [PubMed] [Google Scholar]
  • 58.Marti G, Ferguson M, Wang J, Byrnes C, Dieb R, Qaiser R, et al. Electroporative transfection with KGF-1 DNA improves wound healing in a diabetic mouse model. Gene Ther. 2004;11:1780–1785. doi: 10.1038/sj.gt.3302383. [DOI] [PubMed] [Google Scholar]
  • 59.ter Haar G. Therapeutic applications of ultrasound. Prog Biophys Mol Biol. 2007;93:111–129. doi: 10.1016/j.pbiomolbio.2006.07.005. [DOI] [PubMed] [Google Scholar]
  • 60.Kim HJ, Greenleaf JF, Kinnick RR, Bronk JT, Bolander ME. Ultrasound-mediated transfection of mammalian cells. Hum Gene Ther. 1996;7:1339–1346. doi: 10.1089/hum.1996.7.11-1339. [DOI] [PubMed] [Google Scholar]
  • 61.Endoh M, Koibuchi N, Sato M, Morishita R, Kanzaki T, Murata Y, et al. Fetal gene transfer by intrauterine injection with microbubble-enhanced ultrasound. Mol Ther. 2002;5:501–508. doi: 10.1006/mthe.2002.0577. [DOI] [PubMed] [Google Scholar]
  • 62.Bekeredjian R, Grayburn PA, Shohet RV. Use of ultrasound contrast agents for gene or drug delivery in cardiovascular medicine. J Am Coll Cardiol. 2005;45:329–335. doi: 10.1016/j.jacc.2004.08.067. [DOI] [PubMed] [Google Scholar]
  • 63.Ogawa R, Kagiya G, Feril LB, Jr, Nakaya N, Nozaki T, Fuse H, et al. Ultrasound mediated intravesical transfection enhanced by treatment with lidocaine or heat. J Urol. 2004;172:1469–1473. doi: 10.1097/01.ju.0000139589.52415.3d. [DOI] [PubMed] [Google Scholar]
  • 64.Sheikov N, McDannold N, Sharma S, Hynynen K. Effect of focused ultrasound applied with an ultrasound contrast agent on the tight junctional integrity of the brain microvascular endothelium. Ultrasound Med Biol. 2008;34:1093–1104. doi: 10.1016/j.ultrasmedbio.2007.12.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Tsunoda S, Mazda O, Oda Y, Iida Y, Akabame S, Kishida T, et al. Sonoporation using microbubble BR14 promotes pDNA/siRNA transduction to murine heart. Biochem Biophys Res Commun. 2005;336:118–127. doi: 10.1016/j.bbrc.2005.08.052. [DOI] [PubMed] [Google Scholar]
  • 66.Sheyn D, Kimelman-Bleich N, Pelled G, Zilberman Y, Gazit D, Gazit Z. Ultrasound-based nonviral gene delivery induces bone formation in vivo. Gene Ther. 2008;15:257–266. doi: 10.1038/sj.gt.3303070. [DOI] [PubMed] [Google Scholar]
  • 67.Morille M, Passirani C, Vonarbourg A, Clavreul A, Benoit JP. Progress in developing cationic vectors for non-viral systemic gene therapy against cancer. Biomaterials. 2008;29:3477–3496. doi: 10.1016/j.biomaterials.2008.04.036. [DOI] [PubMed] [Google Scholar]
  • 68.Felgner PL, Gadek TR, Holm M, Roman R, Chan HW, Wenz M, et al. Lipofection: a highly efficient, lipid-mediated DNA-transfection procedure. Proc Natl Acad Sci USA. 1987;84:7413–7417. doi: 10.1073/pnas.84.21.7413. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Wasungu L, Hoekstra D. Cationic lipids, lipoplexes and intracellular delivery of genes. J Control Release. 2006;116:255–264. doi: 10.1016/j.jconrel.2006.06.024. [DOI] [PubMed] [Google Scholar]
  • 70.Li S, Rizzo MA, Bhattacharya S, Huang L. Characterization of cationic lipid–protamine–DNA (LPD) complexes for intravenous gene delivery. Gene Ther. 1998;5:930–937. doi: 10.1038/sj.gt.3300683. [DOI] [PubMed] [Google Scholar]
  • 71.Koltover I, Salditt T, Radler JO, Safinya CR. An inverted hexagonal phase of cationic liposome–DNA complexes related to DNA release and delivery. Science. 1998;281:78–81. doi: 10.1126/science.281.5373.78. [DOI] [PubMed] [Google Scholar]
  • 72.Kol MA, van Laak AN, Rijkers DT, Killian JA, de Kroon AI, de Kruijff B. Phospholipid flop induced by transmembrane peptides in model membranes is modulated by lipid composition. Biochemistry. 2003;42:231–237. doi: 10.1021/bi0268403. [DOI] [PubMed] [Google Scholar]
  • 73.Sakurai F, Nishioka T, Yamashita F, Takakura Y, Hashida M. Effects of erythrocytes and serum proteins on lung accumulation of lipoplexes containing cholesterol or DOPE as a helper lipid in the single-pass rat lung perfusion system. Eur J Pharm Biopharm. 2001;52:165–172. doi: 10.1016/S0939-6411(01)00165-5. [DOI] [PubMed] [Google Scholar]
  • 74.Liu F, Qi H, Huang L, Liu D. Factors controlling the efficiency of cationic lipid-mediated transfection in vivo via intravenous administration. Gene Ther. 1997;4:517–523. doi: 10.1038/sj.gt.3300424. [DOI] [PubMed] [Google Scholar]
  • 75.Barron LG, Gagne L, Szoka FC., Jr Lipoplex-mediated gene delivery to the lung occurs within 60 minutes of intravenous administration. Hum Gene Ther. 1999;10:1683–1694. doi: 10.1089/10430349950017680. [DOI] [PubMed] [Google Scholar]
  • 76.Harvie P, Wong FM, Bally MB. Use of poly(ethylene glycol)–lipid conjugates to regulate the surface attributes and transfection activity of lipid–DNA particles. J Pharm Sci. 2000;89:652–663. doi: 10.1002/(SICI)1520-6017(200005)89:5<652::AID-JPS11>3.0.CO;2-H. [DOI] [PubMed] [Google Scholar]
  • 77.Xu L, Huang CC, Huang W, Tang WH, Rait A, Yin YZ, et al. Systemic tumor-targeted gene delivery by anti-transferrin receptor scFv-immunoliposomes. Mol Cancer Ther. 2002;1:337–346. [PubMed] [Google Scholar]
  • 78.Pardridge WM. Re-engineering biopharmaceuticals for delivery to brain with molecular Trojan horses. Bioconjug Chem. 2008;19:1327–1338. doi: 10.1021/bc800148t. [DOI] [PubMed] [Google Scholar]
  • 79.Durcan N, Murphy C, Cryan SA. Inhalable siRNA: potential as a therapeutic agent in the lungs. Mol Pharm. 2008;5:559–566. doi: 10.1021/mp070048k. [DOI] [PubMed] [Google Scholar]
  • 80.Farjo R, Skaggs J, Quiambao AB, Cooper MJ, Naash MI. Efficient non-viral ocular gene transfer with compacted DNA nanoparticles. PLoS ONE. 2006;1e38:1–8. doi: 10.1371/journal.pone.0000038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Yew NS, Zhao H, Wu IH, Song A, Tousignant JD, Przybylska M, et al. Reduced inflammatory response to plasmid DNA vectors by elimination and inhibition of immunostimulatory CpG motifs. Mol Ther. 2000;1:255–262. doi: 10.1006/mthe.2000.0036. [DOI] [PubMed] [Google Scholar]
  • 82.Blume G, Cevc G. Liposomes for the sustained drug release in vivo. Biochim Biophys Acta. 1990;1029:91–97. doi: 10.1016/0005-2736(90)90440-Y. [DOI] [PubMed] [Google Scholar]
  • 83.Boussif O, Lezoualc'h F, Zanta MA, Mergny MD, Scherman D, Demeneix B, et al. A versatile vector for gene and oligonucleotide transfer into cells in culture and in vivo: polyethylenimine. Proc Natl Acad Sci USA. 1995;92:7297–7301. doi: 10.1073/pnas.92.16.7297. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Yamashiro DJ, Fluss SR, Maxfield FR. Acidification of endocytic vesicles by an ATP-dependent proton pump. J Cell Biol. 1983;97:929–934. doi: 10.1083/jcb.97.3.929. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Fischer D, Bieber T, Li Y, Elsässer HP, Kissel T. A novel non-viral vector for DNA delivery based on low molecular weight, branched polyethylenimine: effect of molecular weight on transfection efficiency and cytotoxicity. Pharm Res. 1999;16:1273–1279. doi: 10.1023/A:1014861900478. [DOI] [PubMed] [Google Scholar]
  • 86.Gosselin MA, Guo W, Lee RJ. Efficient gene transfer using reversibly cross-linked low molecular weight polyethylenimine. Bioconj Chem. 2001;12:989–994. doi: 10.1021/bc0100455. [DOI] [PubMed] [Google Scholar]
  • 87.Goula D, Benoist C, Mantero S, Merlo G, Levi G, Demeneix BA. Polyethylenimine-based intravenous delivery of transgenes to mouse lung. Gene Ther. 1998;5:1291–1295. doi: 10.1038/sj.gt.3300717. [DOI] [PubMed] [Google Scholar]
  • 88.Wightman L, Kircheis R, Rössler V, Carotta S, Ruzicka R, Kursa M, et al. Different behavior of branched and linear polyethylenimine for gene delivery in vitro and in vivo. J Gene Med. 2001;3:362–372. doi: 10.1002/jgm.187. [DOI] [PubMed] [Google Scholar]
  • 89.Jia SF, Worth LL, Densmore CL, Xu B, Duan X, Kleinerman ES. Aerosol gene therapy with PEI: IL-12 eradicates osteosarcoma lung metastases. Clin Cancer Res. 2003;9:3462–3468. [PubMed] [Google Scholar]
  • 90.Hwang SJ, Davis ME. Cationic polymers for gene delivery: designs for overcoming barriers to systemic administration. Curr Opin Mol Ther. 2001;3:183–191. [PubMed] [Google Scholar]
  • 91.Park TG, Jeong JH, Kim SW. Current status of polymeric gene delivery systems. Adv Drug Deliv Rev. 2006;58:467–486. doi: 10.1016/j.addr.2006.03.007. [DOI] [PubMed] [Google Scholar]
  • 92.Wu GY, Wu CH. Receptor-mediated in vitro gene transformation by a soluble DNA carrier system. J Biol Chem. 1987;262:4429–4432. [PubMed] [Google Scholar]
  • 93.Ziady AG, Gedeon CR, Miller T, Quan W, Payne JM, Hyatt SL, et al. Transfection of airway epithelium by stable PEGylated poly-L-lysine DNA nanoparticles in vivo. Mol Ther. 2003;8:936–947. doi: 10.1016/j.ymthe.2003.07.007. [DOI] [PubMed] [Google Scholar]
  • 94.Konstan MW, Davis PB, Wagener JS, Hilliard KA, Stern RC, Milgram LJ, et al. Compacted DNA nanoparticles administered to the nasal mucosa of cystic fibrosis subjects are safe and demonstrate partial to complete cystic fibrosis transmembrane regulator reconstitution. Hum Gene Ther. 2004;15:1255–1269. doi: 10.1089/hum.2004.15.1255. [DOI] [PubMed] [Google Scholar]
  • 95.Yamagata M, Kawano T, Shiba K, Mori T, Katayama Y, Niidome T. Structural advantage of dendritic poly(L-lysine) for gene delivery into cells. Bioorg Med Chem. 2007;15:526–532. doi: 10.1016/j.bmc.2006.09.033. [DOI] [PubMed] [Google Scholar]
  • 96.Gao X, Kim KS, Liu D. Nonviral gene delivery: what we know and what is next. AAPS J. 2007;9:92–104. doi: 10.1208/aapsj0901009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Sokolova V, Epple M. Inorganic nanoparticles as carriers of nucleic acids into cells. Angew Chem Int Ed Engl. 2008;47:1382–95. doi: 10.1002/anie.200703039. [DOI] [PubMed] [Google Scholar]
  • 98.Cai X, Conley S, Naash M. Nanoparticle applications in ocular gene therapy. Vision Res. 2008;48:319–324. doi: 10.1016/j.visres.2007.07.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Davis PB, Cooper MJ. Vectors for airway gene delivery. AAPS J. 2007;9:11–17. doi: 10.1208/aapsj0901002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Peng XH, Qian X, Mao H, Wang AY, Chen ZG, Nie S, et al. Targeted magnetic iron oxide nanoparticles for tumor imaging and therapy. Int J Nanomedicine. 2008;3:311–321. doi: 10.2147/ijn.s2824. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from The AAPS Journal are provided here courtesy of American Association of Pharmaceutical Scientists

RESOURCES