Abstract

The potential use of antisense and siRNA oligonucleotides as therapeutic agents has elicited a great deal of interest. However, a major issue for oligonucleotide-based therapeutics involves effective intracellular delivery of the active molecules. In this Survey and Summary, we review recent reports on delivery strategies, including conjugates of oligonucleotides with various ligands, as well as use of nanocarrier approaches. These are discussed in the context of intracellular trafficking pathways and issues regarding in vivo biodistribution of molecules and nanoparticles. Molecular-sized chemical conjugates and supramolecular nanocarriers each display advantages and disadvantages in terms of effective and nontoxic delivery. Thus, choice of an optimal delivery modality will likely depend on the therapeutic context.

OVERVIEW

Antisense and siRNA oligonucleotides hold great promise as therapeutic agents. Several first generation (phosphorothioate) antisense oligonucleotides are in late phase clinical testing (1–4), while newer oligonucleotide chemistries are providing antisense molecules with higher binding affinities, greater stability and lower toxicity as clinical candidates (5–7). The rapid development of mammalian RNA interference (RNAi) opens the path to a powerful new strategy for therapeutic regulation of gene expression (8–12). Promising results have been attained with small interfering RNAs (siRNAs) in animal models (13–15) and several clinical trials are underway (13). However, despite abundant promise, a number of problems and hurdles remain for oligonucleotide-based therapeutics. Perhaps the most important issue concerns the effective delivery of antisense or siRNA oligonucleotides to their respective sites of action in the nucleus or cytoplasm. In studies of cells in culture, delivery agents such as cationic lipids or polymers are required in order to attain significant antisense or siRNA effects. However, the large size and/or considerable toxicity (16,17) of cationic lipid particles and cationic polymers may render them problematic candidates for in vivo utilization. In contrast, many animal studies and virtually all of the clinical studies thus far have used ‘free’ antisense or siRNA compounds (without a delivery agent), thereby demonstrating that oligonucleotides can function in that form. However, many investigators believe that appropriate delivery platforms could be very helpful for oligonucleotide-based therapeutics (18–20). In this Survey and Summary, we review and analyze chemically based approaches to oligonucleotide delivery, including use of nanocarriers and molecular conjugates. These approaches will be considered both in terms of intracellular delivery to cultured cells and in terms of in vivo biodistribution. Obviously, another important therapeutic strategy will be to use viral vectors for siRNA expression (10,12,21–24), but we will not further consider the viral approach in this review.

BACKGROUND

Antisense and siRNA mechanisms

Here, we briefly summarize aspects of the chemistry and biology of antisense and siRNA oligonucleotides that are salient to their potential as therapeutic agents.

Antisense

RNaseH-mediated degradation of complementary mRNA is the major mode of action of antisense oligonucleotides. However, oligonucleotides that do not support RNaseH can affect gene expression by translation arrest or by altering splicing (25). Target site selection in the mRNA is an important issue and remains rather empirical. Simple phosphodiester oligonucleotides are unstable in the biological milieu; thus, a number of chemically modified oligonucleotides have been developed to enhance stability and to confer other desirable properties (3,5,6). Substitution of sulfur for oxygen forms phosphorothioate oligonucleotides, the most common modification. However, several other highly improved oligonucleotide chemistries have emerged including 2′-OH modifications, locked nucleic acids (LNAs), peptide nucleic acids (PNAs), morpholino compounds and hexitol nucleic acids (HNAs). All of these entities have high affinities for RNA and are more stable than phosphorothioates; however, they do not support RNaseH activity (5–7). Thus oligomers based entirely on these chemistries cannot be used as ‘classic’ antisense agents (although they may be very effective for modification of splicing or translation arrest). Introduction of several central phosphodiester residues into these agents, thereby creating ‘gapmers’, results in antisense oligonucleotides that activate RNaseH but that also retain many of the desirable properties of the parent compounds (7).

siRNA

Suppression by double-stranded RNA (dsRNA) is an important endogenous mechanism of gene regulation, acting through pathways involving mRNA degradation and/or sequestration, translation arrest and effects on chromatin and transcription (26). The mRNA cleaving action of interfering dsRNA in mammals involves two enzymatic steps. First, the ‘Dicer’ enzyme and its co-factors cleaves dsRNA to 21- to 23-mer segments (siRNA) and assists its’ loading on to the Argonaute 2 (Ago 2)-containing ‘RISC’ complex. RISC removes the sense strand, pairs the antisense (guide) strand with a complementary region in the cognate mRNA and initiates cleavage (‘slicing’) at a site between bases 10 and 11, relative to the 5′ end of the antisense strand (21,27–29). The resulting 5′ and 3′ mRNA fragments are subsequently fully degraded by several cellular nucleases (26,30).

In addition to ‘slicer’ activity, which requires essentially complete complementarity between the siRNA guide strand and its target, short dsRNAs can also display miRNA activity against partially complementary sequences in the 3′-regulatory regions of mRNAs. While bound to Ago protein complexes, the ‘seed region’ of the antisense strand (positions 2–7, 8 from the 5′ end) pairs precisely with the target, while some mismatches are tolerated in other regions. This process can lead to arrest of translation, sequestration of the target mRNA in cytosolic P-bodies (which are key sites of RNA processing), and possibly to de-capping and degradation (26,31). Thus, miRNA-mediated actions can potentially lead to off-target effects of siRNAs. In addition to the ‘slicer’ and miRNA pathways, dsRNA can also regulate transcription at the chromatin level via processes that are not yet fully elucidated (32). There is also an interesting conjunction between siRNA effects and antisense mechanisms. Thus, antisense oligonucleotides can be designed to interrupt the function of endogenous miRNAs. Since miRNAs often reduce gene expression, these antisense agents (sometimes termed ‘antagomirs’) can cause upregulation of expression of some of the genes that are regulated by a particular miRNA (11,33–35).

In mammals, long dsRNAs elicit highly toxic responses related to the effects of viral infection and interferon production (8,28). To avoid this, Tuschl and colleagues initiated the use of short interfering RNAs (siRNAs), comprised of 19-mer duplexes with 2 base 3′ overhangs on each strand, that associate with Ago2 and selectively degrade targeted mRNAs (36). Short, chemically synthesized, siRNAs do not require the Dicer step, although Dicer-enhanced RISC loading may contribute to efficiency.

To enhance siRNA effectiveness a variety of chemical modifications have been pursued including alterations in the backbone chemistry, 2′-sugar modifications, altered ring structure, nucleobase modifications and others (37–40), and the importance of these modifications to potential clinical applications has been emphasized (41). In terms of overall design, understanding of the biochemical mechanism of RNA interference has provided important guidelines; first, the siRNA must maintain an A-form (RNA-like) duplex, second the 5-position on the antisense strand must be able to be phosphorylated, third to be effective siRNAs should have low thermodynamic stability in the 5′ antisense region (26,40,42). These general design principles can then be further refined through the use of appropriate chemical modifications (39).

It is important to note, however, that siRNA biology is complex, and that it is essential to validate the mechanisms underlying observed biological effects before attributing them to RNA interference. A dramatic example of this concerns a recent study of siRNAs designed to ameliorate macular degeneration by targeting VEGF or its receptor and thus suppressing angiogenesis in the eye. A comprehensive analysis of the situation revealed that the observed reduction of angiogenesis was not due to sequence-specific RNA interference at all, but rather was caused by sequence-independent stimulation of Toll-like Receptor 3 and its downstream signaling pathway by dsRNA-like molecules, leading to an interferon-γ and IL-12 mediated suppression of neovascularization (43). Thus, the possible interplay of oligonucleotides with various members of the Toll-like receptor family of cell surface proteins must be considered in the design of experiments.

Mechanisms of endocytosis and intracellular trafficking

In almost all instances, oligonucleotides (and their various delivery agents) are taken up by some form of endocytosis. Ultimately, the oligonucleotide must exit from the endosome to reach its site of action in the cytosol or nucleus. Thus, in order to understand key issues in the intracellular delivery of oligonucleotides it is important to consider the multiple routes of endocytosis and the complex trafficking pathways that exist in cells.

Endocytosis is a blanket term that covers multiple distinct uptake pathways including: (i) the ‘classic’ clathrin-coated pit pathway; (ii) the caveolar pathway; (iii) one or more noncaveolar, clathrin-independent pathways (CLIC pathways); (iv) macropinocytosis and (v) phagocytosis (that mainly takes place in ‘professional phagocytes’ such as macrophages and granulocytes) (44,45). When the molecule being internalized is simply dissolved in the ambient medium, the uptake process is usually termed pinocytosis. When the molecule is bound to a cell surface receptor the process is termed receptor-mediated endocytosis. Each of the endocytotic pathways is actin-dependent (with the possible exception of one CLIC pathway), and the clathrin and cavelolar pathways are dependent on the GTPase dynamin to pinch off budding vesicles, while macropinocytosis is not dynamin dependent. Caveosomes and CLIC vesicles are enriched in ‘lipid raft’ components including glycosphingolipids and cholesterol (46,47). Initial uptake of receptor and ligand is followed by sequential intracellular trafficking into a variety of low pH endomembrane compartments, including early/sorting endosomes, late endosomes/multi-vesicular bodies, and lysosomes; in some cases, receptors/ligands can traffic to the Golgi complex. In many instances, receptor and ligand are dissociated in the low pH endosome environment, and in some cases the receptor can recycle back to the cell surface via the sorting endosomes (Figure 1).

Endocytotic pathways. The figure depicts the multiple endocytotic pathways that may be involved in uptake of oligonucleotides. The black arrows represent well-documented trafficking routes within the cell. The names in red indicate well-known protein markers for various endomembrane compartments; in most cases, antibodies to these marker proteins are commercially available.
Figure 1.

Endocytotic pathways. The figure depicts the multiple endocytotic pathways that may be involved in uptake of oligonucleotides. The black arrows represent well-documented trafficking routes within the cell. The names in red indicate well-known protein markers for various endomembrane compartments; in most cases, antibodies to these marker proteins are commercially available.

The complex flow of endomembrane traffic (48) is regulated by the Rab family of small GTPases and by tethering complexes, while vesicular fusion events are controlled by SNARE proteins (49–51). The SNX (sorting nexin) proteins also are important in sorting and cargo retrieval from endosomes (52). The trafficking of internalized receptor and ligand can often be very complex and dependent on the nature of the receptor, the physiological status of the cell and the cell type. Some receptors can enter cells via multiple pathways, for example, via both clathrin-coated vesicles and caveolae (53). The route of entry can affect the ultimate fate and function of the ligand–receptor complex. For example, at low ligand levels the EGF-receptor is internalized via coated pits and can continue to signal, while at high ligand concentrations the receptor is ubiquitinated, internalized via a CLIC or caveolar pathway, and degraded (54).

There are a number of tools commonly used to probe pathways of internalization. For example, many studies of oligonucleotide uptake have used pharmacological inhibitors that putatively block specific endocytotic pathways. In reality, however, such inhibitors often affect multiple uptake pathways, as well as having many other effects on the cell. For example, although β-cyclodextrin is often used to block uptake involving lipid rafts, it also affects clathrin-mediated endocytosis (55). Likewise, agents that block macropinocytosis by affecting actinomyosin function will also block other endocytotic pathways, as well as having generalized effects on the cytoskeleton (56,57). Thus, results obtained with pharmacological inhibitors need to be interpreted conservatively. Potentially, more powerful approaches include using antibodies to key marker proteins to co-visualize fluorophore-tagged oligonucleotides in specific endomembrane compartments (58), and use of dominant-negative Rab proteins to interfere selectively with trafficking patterns (59).

Although general aspects of the cellular uptake of oligonucleotides have been studied for a long time and extensively reviewed (7,60,61), it is only recently that investigators have probed oligonucleotide internalization in light of current understanding of endosomal trafficking pathways. While not enough work has been done to achieve a broad consensus, some interesting examples of more detailed trafficking studies have recently emerged (62–64).

Ligands for enhancing and targeting delivery

In this section, we will examine ligands that have been used to attain cell-type selective targeting or to enhance the uptake of oligonucleotides. For the sake of simplicity, we have divided the discussion into agents that target specific receptors (CTLs, cell targeting ligands) and agents that enhance transmembrane permeation (primarily cell penetrating peptides, CPPs).

CPPs

The chemistry and actions of CPPs have been extensively reviewed (65–69) and here we touch on only a few key aspects. The prototypical CPPs derived from the Tat and antennepedia transcriptional regulators have been joined by a large number of new moieties. Most of these are relatively short (9–30 amino acids) polycationic peptides rich in argine and lysine, although some include membrane-interactive hydrophobic sequences (Figure 2). CPPs have been linked to proteins by recombinant DNA techniques or chemically coupled to peptides, oligonucleotides or nanocarriers, which then comprise the ‘cargo’ for the CPP. Initially, CPPs were thought to convey their cargo directly across the plasma membrane. However, it is now clear that polycationic CPPs initially bind to cell surface glycosaminoglycans; this is followed by endocytotic uptake (possibly macropinocytosis) (70), and eventual release of cargo from the endosome to the cytosol. Although initial reports emphasized the great promise of CPPs for delivery of macromolecules, recently there has been some controversy as to just how effective they are (71). Certainly, the nature of the cargo (in terms of size, charge and other molecular characteristics) has an important impact on the effectiveness and the toxicity of CPP-mediated delivery (72–75). In a section below, we will discuss how CPPs have been applied to oligonucleotide delivery.

Cell penetrating peptides. This describes some of the more commonly used cell penetrating peptides in terms of molecular weight and net charge.
Figure 2.

Cell penetrating peptides. This describes some of the more commonly used cell penetrating peptides in terms of molecular weight and net charge.

CTLs

A promising strategy is to deliver antisense and siRNA oligonucleotides by use of a CTL that binds with high affinity to a cell surface receptor that is capable of undergoing efficient internalization. A wide variety of potential ligands are available including antibodies (76), polypeptides derived from phage display libraries (77) and small organic molecules. Since various receptors are often preferentially expressed on particular cell types, this approach offers the possibility of improved selectivity for the oligonucleotide reagents. While a rich variety of cell surface receptors are expressed in the human body, work thus far involving delivery of oligonucleotides has primarily focused on lipoprotein receptors (particularly those in the liver) (78), integrins (79,80) and receptor tyrosine kinases (81). Another potentially rich source of targets is the G-protein coupled receptor (GPCR) superfamily, by far the largest family of receptors in the human genome with approximately 850 members (82). GPCRs have long been a major interest of the pharmaceutical industry and thus a vast number of high affinity small organic molecule ligands for GPCRs are available or are emerging via high-throughput screening procedures (83).

CELLULAR DELIVERY OF OLIGONUCLEOTIDE CONJUGATES AND COMPLEXES

In this section, we discuss the use of CPPs and CTLs in the delivery of antisense and siRNA oligonucleotides. Emphasis here is on cellular studies, while in vivo work is discussed in a following section. As a prelude, it is important to recall that oligonucleotides do not permeate intact cell membranes to any significant degree via simple diffusion, primarily because of the hydrophobic nature of the membrane lipid bilayer. This is true for both negatively charged siRNA or antisense moieties as well as for molecules with uncharged backbones such as methylphosphonates, PNAs and morpholinos (84,85).

CPP—oligonucleotide conjugates or complexes

A considerable effort has gone into the preparation and evaluation of conjugates of CPPs and oligonucleotides; however, on the whole this has been only moderately successful (86,87). Our laboratory has reported biological effects of conjugates of CPPs with anionic antisense oligonucleotides (88,89), and others have reported on CPP–siRNA conjugates (90,91). However, the bulk of the literature suggests that CPPs are primarily able to deliver oligonucleotides with uncharged backbones, such as peptide nucleic acids and morpholino compounds (92–96).

CPPs have been studied in connection with both antisense and siRNA, and as chemical conjugates or noncovalent complexes with the oligonucleotide. In addition to the usual assays based on ‘knockdown’ of mRNA by antisense or siRNA, another popular approach has been the use of an assay based on the splice-correcting properties of certain types of antisense oligonucleotides (25). Briefly, an aberrant intron is placed into a reporter gene (luciferase, GFP) cassette and stably expressed in cells. The aberrant intron results in incorrect splicing and failure to produce functional mRNA and protein. However, appropriate splice switching oligonucleotides (SSOs) can correct splicing leading to expression of the reporter gene. Since splicing only takes place in the nucleus, this splice correction assay provides a convenient positive read-out for delivery of the SSO to the nuclear compartment.

Early work from our laboratory explored conjugates of the CPPs Tat and Antennepedia (also know as penetratin) with either standard phosphorothioate oligonucleotides targeting the MDR1 drug resistance gene (89,97) or with SSOs comprised of 2′-O-Me phosphorothioates (88); in both cases, the peptide and oligonucleotide were joined by bio-reversible S–S bridges. In both types of assay, the presence of the CPP enhanced biological effects over those attained with unconjugated ‘free’ antisense oligonucleotide; additionally, a limited amount of microscopy was done to evaluate intracellular distribution and evidence was found for nuclear accumulation of the oligonucleotide. In contrast, a later study from another group examined a number of disulfide bridged conjugates between various CPPs and 2′-O-Me or LNA oligonucleotides complementary to the HIV-1 TAR element (98). In this case, there was little biological effect unless endosome disrupting agents were used; further, the oligonucleotides were observed by fluorescence microscopy to be restricted to cytoplasmic vesicles, with no sign of nuclear localization. The reason for the discrepancies between these two sets of early studies is unclear. One possibility is that, in our early studies, the peptide–oligonucleotide conjugates became aggregated during use, and this actually enhanced their effectiveness. Interestingly, there have been several reports of noncovalent complexes or aggregates between anionic siRNA oligonucleotides and cationic CPPs that seem to have provided moderately effective delivery to cells in culture. In one case, a modified version of the CPP penetratin having endosomolytic properties was superior to other CPPs that bound the siRNA equally well but lacked endosome lytic ability (99). In another case, a designed peptide comprised of both positively charged residues and a fusion peptide sequence was found to complex with siRNA and deliver it to the cytosol (100); this approach has also been followed by other groups (101). Whether conjugates or complexes are likely to provide more effective delivery of anionic oligonucleotides in vivo is an important issue and will be further explored below.

A number of investigators have evaluated conjugates of CPP with oligonucleotides having uncharged backbones. In one study, both stable and bioreversible disulfide linkages were used to produce conjugates between various CPPs and a PNA targeting the HIV-1 Tar motif (92). Certain conjugates, particularly an R6–penetratin version, demonstrated clear-cut biological effects, although little nuclear localization was seen by fluorescence microscopy. Several other CPP–oligonucleotide conjugates, with either stable or disulfide linkages, were able to attain biological effects when chloroquine was used to enhance their release from endosomes. An additional study used an R6–penetratin conjugate of a PNA SSO to activate a luciferase reporter gene; good effects were attained at micromolar concentrations (96). Another study also utilized PNA SSOs coupled via disulfide bridges to various CPPs to activate luciferase (93); here the transportan CPP was found to be particularly effective. Confocal microscopy and use of endosomal markers indicated that the CPP–PNA conjugates were most likely taken up by macropinocytosis, but there was little evidence of nuclear localization despite the observed effects on splicing. Studies from another group examined additional conjugates between PNA SSOs and various CPPs; they also found that a transportan–PNA conjugate linked via a bioreversible disulfide bridge was most effective (94). Recent studies have described a novel CPP termed ‘M918’ (63). When conjugated to PNA SSOs, M918 did not require binding to cell surface glycosaminoglycans for uptake (in contrast to Tat or penetratin). Nonetheless, it entered cells by endocytosis and attained good biological effects. An interesting variation used CPP–PNA conjugates to target chromosomal DNA and cause effects at the transcriptional level (102).

Similar studies have also been done with CPP conjugates of morpholino oligonucleotides. In one very comprehensive investigation, a variety of linkages were formed between a morpholino SSO and several CPPs (103). A peptide containing nine arginines was particularly effective and resulted in splice correction activity at micromolar concentrations; fluorescence microscopy indicated some delivery of the oligonucleotide to the nucleus as well as to intracellular vesicles. More recently, this group has investigated the properties of conjugates comprised of morpholino SSOs linked to CPPs containing 6-aminohexanoic acid residues (104,105), finding that these have superior properties in terms of stability and effects on splicing.

The strategy of using CPP conjugates of SSOs having uncharged backbones has recently been reviewed (106). The overall picture seems to be that conjugates of various CPPs with uncharged backbone oligonucleotides can enter cells and effectively alter RNA splicing processes. Various CPPs differ somewhat in their potency in this regard; however, in most cases biological effects are only attained when the conjugates are used at micromolar concentrations. This may indicate that most of the material taken up by cells remains in endosomal compartments, with only a tiny fraction reaching the nucleus where RNA splicing occurs. One technical issue with many of these studies is their reliance on a single model system involving correction of splicing in a modified HeLa cell line. As discussed earlier, this system allows facile evaluation of whether a splice switching oligonucleotide can reach the nucleus and correct splicing of the aberrant reporter gene present there. However, it seems unwise to rely so heavily on a single cell type.

CTL–oligonucleotide conjugates or complexes

A number of studies have appeared recently using CTLs for the delivery of antisense or siRNA. Some of these studies had in vivo components that will be more fully discussed in a section below. Here, we will focus on the cell targeting and uptake aspects.

An aptamer-siRNA chimera targeting prostate-specific membrane antigen (PSMA) was able to effectively deliver the associated siRNA to LNCaP prostate cancer cells; use of plk-1 siRNA triggered apoptosis and resulted in cell death both in culture and in a prostate tumor model (107). In this case, concentrations in the 2–400 nM range were effective in attaining gene silencing in cultured LNCaP cells, but not in PC-3 cells that lack PSMA. Other interesting approaches to aptamer-mediated siRNA delivery have also been described (108). The conjunction of nucleic acid aptamer technology and siRNA could potentially be a very powerful avenue for developing reagents for cell type selective regulation of gene expression.

Another approach involved a chimeric protein comprising the highly positive peptide protamine and an antibody Fab fragment directed against the HIV-envelop glycoprotein; this proved to be an effective carrier for siRNA that is complexed noncovalently with the protamine moiety. The chimeric protein was able to deliver an HIV gag siRNA to HIV infected CD4+ T cells causing inhbibition of HIV replication (109). A later version of this approach used a conformation-sensitive single chain antibody directed against the LFA-1 integrin to specifically target siRNA to activated leukocytes (110); in this case, the complexed siRNA was directed against the CCR5 chemokine receptor. Effective gene silencing was obtained with amounts of siRNA in the sub-nanomol range, although the exact concentrations used are unclear.

In another example, a small cyclic peptide that binds the IGF1-receptor was able to deliver a PNA antisense moiety to the cytoplasm of cells expressing this receptor (111). A similar approach was also used for delivery of siRNA directed to the signaling protein IRS1 (112). Here, the peptide was conjugated via an NHS linker to a 5′-aminolinked sense strand. Significant silencing of the target gene in MCF7 breast cancer cells was observed using concentrations of the conjugate in the 100 nM range.

In a similar vein, work from our laboratory has recently shown that a bivalent RGD peptide having high affinity for the avb3 integrin can effectively deliver conjugated SSOs to melanoma cells that express this integrin (62). Significant effects on splicing were attained with concentrations of conjugated SSO in the 10 nM range. The uptake process of the conjugates was traced via confocal fluorescence microscopy; this indicated that the RGD-conjugates entered via caveolae and other lipid raft-rich structures and then eventually trafficked to the trans-golgi. While substantial nuclear localization was not seen, the biological effects observed make it clear that some of the SSO reached the nucleus. An important point is that conjugates of this type display very little cytotoxicity, even when used at concentrations far higher than those needed to obtain a biological effect.

Another study used a polymer as a carrier for both siRNA and a targeting ligand. Thus, the polymer was covalently ‘decorated’ with siRNA, polyethylene glycol (PEG) and N-acetylgalactosamine as a ligand to target the hepatic asialo-glycoprotein receptor (113). This approach permitted effective silencing of two endogenous genes in cultured hepatocytes.

A particularly impressive study involved delivery of siRNA to neurons in culture, and to the brain, by complexation with a peptide that comprises a positively charged (Arg9) sequence to bind the oligonucleotide and a sequence that binds with high affinity to the nicotinic acetycholine receptor in neurons (114). The chimeric peptide selectively delivered siRNA to neural cells expressing the acetycholine receptor, but not to other cells, and could silence a GFP reporter gene in the neuronal cells when used at 10 pmol levels.

There is a striking functional contrast between the studies utilizing CTLs for oligonucleotide delivery and those using CPPs. In many cases, biological effects were attained using nanomolar concentrations of oligonucleotides when delivery took place via receptor targeting, whereas delivery using various CPPs attained strong biological effects only at micromolar concentrations. The reason for this differential is unclear and may have little to do with total uptake (although it is not possible to reliably compare this parameter for the different studies). Possibly, a key issue is the intracellular trafficking pathway(s) accompanying the various initial uptake processes.

Nanocarriers for oligonucelotide delivery

A variety of supramolecular nanocarriers including liposomes (115), cationic polymer complexes (116) and various polymeric nanoparticles (117) have been used to deliver antisense and siRNA oligonucleotides, as more fully described in several recent reviews (3,9,38,118–122). There is an enormous literature on use of various nanocarriers to deliver nucleic acids; thus here we can only touch on selected recent examples. Complexation of oligonucleotides with various polycations is a popular approach for intracellular delivery; this includes use of PEGlyated polycations (123), polyethyleneamine (PEI) complexes (124,125), cationic block co-polymers (126) and dendrimers (127–130). Several cationic nanocarriers including PEI and polyamidoamine dendrimers exert a so called ‘proton sponge effect’ that helps to release contents from endosomes (131). Thus, as the nucleic acid–polymer conjugate enters the low pH endosome compartment, secondary amino groups in the polymer are titrated; the necessary proton influx also brings chloride and water into the endosome, causing swelling and increased leakiness. Other widespread approaches include use of polymeric nanoparticles (132), polymer micelles (133), quantum dots (134,135) and lipoplexes (136,137). Lipoplexes comprised of cationic lipids also exert endosome destabilizing effects (138); in this case, the cationic lipids interact with anionic lipids of the endosome membrane, leading to the formation of nonbilayer structures and consequent endosome instability. In some cases, nanoparticle approaches have been coupled with targeting strategies. As one example, a lactosyl–PEG–siRNA conjugate was complexed with polylysine to form nanoparticles; these were effectively delivered to liver tumor cells via interaction with the asialo-glycoprotein receptor (139). In considering the various types of nanocarriers, it is important to keep in mind that the carrier systems themselves can have significant effects on gene expression, and may potentially cause toxicity. This has been emphasized in two excellent recent reviews that comprehensively describe effects of polymers and nanocarriers on gene expression (17,140).

IN VIVO DELIVERY OF OLIGONUCLEOTIDES

In this section, we will consider recent investigations regarding the in vivo behavior of various oligonucleotide conjugates and nanocarrier formulations. We will place particular emphasis on several studies that have demonstrated clear-cut enhanced pharmacological effects of systemically administered siRNA as a result of use of delivery modalities. Some of these reports have been discussed above in terms of results at the cell culture level. Prior to initiating this discussion it is important to realize that there is a dichotomy in the behavior of oligonucleotides when comparing the in vivo situation to cell culture. Thus, almost without exception, effective use of antisense or siRNA in cell culture requires a delivery agent such as a cationic lipid; in contrast, many of the successful in vivo studies with oligonucleotides have used ‘free’ compounds (7,60). There seem to be two possible interpretations of this dichotomy. One version suggests that cells undergo radical changes in organization and gene expression as they go from a 3D tissue environment to a 2D culture environment and that in this process key oligonucleotide transporters are lost. While it is clear that cells do undergo dramatic changes in the transition from 3D to 2D (141), it seems unlikely to us that the same oligonucleotide transport systems would be lost in every single type of cell. Another interpretation is a pharmacodynamic one. Because of experimental constraints, in culture cells are only briefly exposed to the oligonucleotides; in contrast, most in vivo therapeutic experiments involve multiple doses and protracted exposures, thus perhaps allowing gradual intercellular accumulation of oligonucleotides. Surprisingly, this important dichotomy has not been carefully addressed via experimentation.

Biological barriers to in vivo delivery of oligonucleotides

In planning for the effective delivery of oligonucleotides, it is essential to understand key aspects of endocytosis and intracellular trafficking at the cellular level. However, in vivo there are a number of other important parameters to consider as well (Figure 3). Essentially, a series of biological barriers stand between the newly administered oligonucleotide and its ultimate site of action in the cytosol or nucleus of tissue cells (142). For ‘free’ oligonucleotides or small conjugates, an important limitation is rapid excretion via the kidney. Molecules less than 5000 molecular weight are rapidly ultrafiltered in the glomerulus and, in the absence of re-uptake, are accumulated in the urine (143). This picture is somewhat altered with oligonucleotides that bind strongly to plasma proteins thus retarding ultrafiltration (144). The vascular endothelial wall comprises another major barrier, especially for larger carriers. In general, molecules with a diameter of >45 Å (equivalent to about the size of an immunoglobulin) do not readily pass across the capillary endothelium and thus cannot efficiently enter the extracellular fluid that bathes tissue cells (145). In addition, it is not only the size but also the shape of the macromolecule or nanocarrier that affects its ability to traverse the endothelium (146,147). In a few tissues, including liver and spleen, the vascular endothelium is ‘fenestrated’ with gaps that allow the egress of macromolecules and nanoparticles up to about 200 nm diameter (148). In addition, work in xenograft tumors has given rise to the concept that the tumor vasculature is far ‘leakier’ than normal vasculature, thus also allowing egress of relatively large macromolecules and nanoparticles, the so-called ‘EPR effect’ (enhanced permeation and retention) (149). However, it is not clear that spontaneously occurring tumors in animals or humans uniformly display such increased leakiness. Further, tumors often exhibit other properties, such and increased interstitial pressure, that would tend to oppose delivery of nanocarriers to the tumor (150). Although one must be concerned about the issue of vascular permeability, it is also important to realize that the vascular endothelial cells themselves can be portals for therapy. This is especially true in sites of inflammation, where the endothelium upregulates key cell surface proteins including VCAM, ICAM and P-Selectin (151) or in angiogenic endothelium where the avb3 integrin is upregulated (152); thus, in both situations there is enhanced expression of receptors that can be addressed via CTLs linked directly to oligonucleotides or to nanocarriers. There are already many examples in the literature where endothelial receptors, especially avb3, have been targeted by various macromolecular or nanoparticle carriers bearing drugs or imaging agents (153–155).

In vivo barriers. The figure lists key barriers to effective in vivo delivery of oligonucleotides. Rapid excretion is an issue for low molecular weight compounds. Clearance by phagocytes, capillary permeability and slow diffusion in the extracellular matrix apply to larger molecules and nanoparticles. Both small and large delivery agents can be affected by poor cellular uptake and inefficient release from endosomes.
Figure 3.

In vivo barriers. The figure lists key barriers to effective in vivo delivery of oligonucleotides. Rapid excretion is an issue for low molecular weight compounds. Clearance by phagocytes, capillary permeability and slow diffusion in the extracellular matrix apply to larger molecules and nanoparticles. Both small and large delivery agents can be affected by poor cellular uptake and inefficient release from endosomes.

Even if a nanocarrier exits the vasculature, it still needs to diffuse through the extracellular matrix to reach tissue cells; the ECM is comprised of a dense meshwork of proteins and proteoglycans that can hinder nanocarrier diffusion and in some cases may even tightly bind the carrier (142). Another key barrier is presented by the phagocytes of the reticuloendothelial system (RES). These cells monitor the blood and remove foreign materials such bacteria and viruses; unfortunately, they also tend to treat administered macromolecules and nanoparticles as foreign, thus accumulating these materials in hepatic Kupffer cells, splenic macrophages and other sites rich in phagocytes (156,157). This process can be attenuated to some degree by ‘passivating’ or ‘stabilizing’ the surfaces of nanoparticles with hydrophilic polymers such as PEG (158); PEGlyation serves to reduce the adsorbtion of ‘opsonins’, plasma proteins that enhance phagocytosis, but this tactic is never completely effective. PEG-conjugated nanocarriers can remain in the circulation much longer than unmodified versions, but ultimately significant clearance by the RES takes place. In summary, a variety of considerations at the both cellular and whole organism levels are involved in the design of effective in vivo delivery strategies for therapeutic oligonucleotides.

In vivo delivery of oligonucleotide conjugates

At this point, relatively little is known about the in vivo behavior of ligand–oligonucleotide conjugates. A recent study examined the biodistribution of a conjugate between an arginine-rich CPP and a morpholino oligonucleotide and suggested increased uptake in many tissues as compared to free oligonucleotide (159). Studies of cholesterol-linked siRNAs indicated that their association with serum proteins plays an important part in their pharmacokinetics, biodistribution and ultimate effects on gene expression in liver cells (14). Thus, the elimination half-life and tissue accumulation of an apolipoprotein B-targeted siRNA was substantially increased via cholesterol conjugation, leading to enhanced reduction of target mRNA and protein and consequent effects on blood cholesterol levels. A more detailed analysis of the behavior of lipidic conjugates of siRNA indicated that HDL and LDL were the primary carriers for cholesterol-linked siRNA, while conjugates of medium chain fatty acids primarily bound albumin (160). This report further demonstrated the key role of the LDL-and HDL-receptors in tissue uptake of cholesterol siRNA, with uptake via the LDL-receptor predominating in liver. Interestingly, this report also suggests a very novel mechanism for cell entry of siRNA via lipoprotein receptors. Thus, instead of simple receptor-mediated endocytosis of the siRNA-loaded lipoprotein, the authors suggest that the siRNA is passed from the lipoprotein receptor to Sid 1, a multispanning plasma membrane protein whose homolog in C. elegans can potentiate RNAi uptake and effects. Interestingly, other groups, working with mammalian cell cultures, have also linked Sid1 to intracellular uptake of siRNA (161,162). If these observations are generalizable, it would have a profound effect on our understanding of the transport of siRNA and on approaches to delivery and therapeutics.

Cholesterol-conjugated siRNA has also been administered directly into the lung via intratracheal instillation. In this case, the target was p38 MAP kinase mRNA. Significant target reduction was attained with the cholesterol siRNA but not with CPP-conjugated siRNA; indeed, in this case, potentially toxic results were observed (163). An interesting variation on delivery via lipid conjugation is embodied in a recent report on α-tocopherol modified siRNA; this lipidic material also promotes siRNA delivery via binding to serum proteins and lipoproteins, but may involve a different set of binding partners (164).

In vivo delivery of oligonucleotides using nanocarriers

A variety of nanocarriers have been developed to promote the effective in vivo delivery of oligonucleotides, with the emphasis on siRNA. An impressive early study involved complexation of siRNA with cationic cyclodextrin polymers to form nanoparticles, and used transferrin to target the nanocomplex to Ewing's sarcoma tumor cells that express high levels of the transferrin receptor (165). Delivery of siRNA targeting the EWS-FLI1 oncogene product resulted in reduced tumor growth. This same technology has more recently been tested for safety in primates, but with the active material being siRNA targeting the M2 subunit of ribonucleotide reductase (166).

In another study involving in vivo targeted delivery, ‘self assembled nanoparticles’ (a form of lipoplex) were used to deliver siRNA to tumors (167). Anti-EGF-receptor siRNA (as well as carrier DNA) was complexed with protamine and then with lipid. The particles were passivated with PEG and targeted using anisamide as a ligand. The nanoparticles were used to treat mice bearing xenografts of the NCI-H460 tumor, which expresses high levels of a cell surface receptor that binds anisamide; this treatment resulted in partial reduction of tumor growth. In addition, this study examined the pharmacokinetics and biodistribution of the administered siRNA, observing extensive tumor uptake for the targeted nanoparticles but not untargeted controls.

A very recent study also used a liposome-type carrier for targeted in vivo delivery of siRNA (168). Here, uncharged lipids were used to form small unilamellar liposomes. These were covalently linked to hyaluronan to stabilize and passivate the liposomes, which were then further conjugated to an antibody that binds the beta7 integrin subunit. The antibody-targeted liposomes were then ‘loaded’ with a protamine–siRNA complex. This technology was used to selectively deliver cyclin D1 siRNA to beta7-positive leukocyte subsets involved in intestinal inflammation, resulting in amelioration of an experimentally induced colitis.

Stable nucleic acid lipid particles (SNALPs), a type of liposome, have proven very effective for delivery of siRNA to the liver (169). In this case, apoB siRNA was used to treat cynomolgus monkeys resulting in substantial silencing of apoB mRNA expression followed by reduced protein levels and reductions in blood cholesterol. Extensive pharmacokinetic studies were also performed. The SNALPs are not a targeted nanocarrier, but rather rely on stability and long circulation time to attain effective delivery. Another promising approach using liposomes involved the suppression of liver fibrosis via delivery of siRNA targeting a heat-shock protein (HSP47) in hepatic stellate cells. These cells express a receptor for vitamin A and the liposomes were thus complexed with this substance in order to attain targeting both in hepatic cell cultures and in vivo (170). This study was notable for the very extensive validation of the in vivo therapeutic response.

Another interesting approach, discussed previously, involves creating protein chimeras of Fab or scFv antibody components with protamine, followed by complexation with siRNA (109). This ‘nanocomplex’ approach was used to deliver growth inhibitory siRNAs in vivo to B16 melanoma cells engineered to express the HIV envelop glycoprotein, or to breast tumor cells overexpressing ErbB2, using the appropriate antibody component in each case. A modification of this approach (110) used a protamine chimera with a conformation-sensitive single chain antibody directed against the LFA-1 integrin to specifically target siRNA in vivo to lymphoid tumor cells that express the activated form of LFA-1 (110).

Polymer systems have also been used for targeted in vivo delivery. Thus, as mentioned earlier, a polymer nanocarrier linked to N-acetylgalactosamine was used to promote selective uptake by hepatocytes (113). In vivo, this system was able to effectively deliver apoB siRNA to mouse liver, resulting in gene silencing and reduced systemic cholesterol levels. Somewhat similarly, a system involving an RGD-conjugate of PEI was used to deliver siRNA targeting VEGF-receptor to tumor vasculature (125).

Finally, as mentioned earlier, delivery of siRNA to the brain was attained by formation of a nanocomplex with a peptide comprising positively charged sequence to bind the oligonucleotide and a sequence that binds with high affinity to the nicotinic acetycholine receptor in neurons (114). Use of an antiviral siRNA provided protection against a potentially fatal viral encephalitis in mice. It is very surprising that the siRNA nanocomplex was able to cross the blood–brain barrier since even most small molecule drugs fail to do so. However, in this case it is possible that the presence of an active viral infection altered the permeability of the barrier.

In summary, several types of nanoparticle technologies have effectively delivered siRNAs to the liver, to tumors, or to inflammed tissues. It should be noted, however, that the vasculature in these tissues differs substantially from that more generally present in the body, as has been discussed earlier. Thus, it is unclear whether these approaches can be generalized to other tissues. Several of the studies mentioned above have also been reviewed recently in another venue (171).

CONCLUSIONS

As reflected in this article, a great deal of effort is currently focused on the delivery of therapeutic oligonucleotides. Significant strides have been made, but the issue has not been fully resolved. Thus, it seems valuable to compare and contrast in broad terms the various strategies that are being pursued. Perhaps the most significant parameter to consider is the size of the moiety being delivered, contrasting nanoparticles that have diameters in the 50–200 nm range and molecular weights in the millions to monomolecular oligonucleotide conjugates with molecular weights in the thousands. Another key issue is charge since both polyanionic and polycationic molecules or nanocarriers can interact with blood proteins to incite toxicity or affect biodistribution (172,173).

Nanoparticles of various types offer many advantages as delivery agents. They can carry a large ‘payload’ comprising hundred or thousands of copies of the siRNA or antisense oligonucleotide. They can be decorated with multiple copies of targeting ligands, thus providing high-avidity interaction with the target cells. Nanoparticles can be designed to release their contents at prescribed rates and can also be engineered to assist in the release of their contents from endosomes. Novel approaches for producing extremely uniform nanoparticles with controllable drug release characteristics are becoming available (174). Technologies for producing nanoparticles are reasonably mature thus permitting relatively facile scale-up for clinical studies (although reliable large-scale formulation of nanoparticles under GMP conditions is not without problems). Additionally, regulatory agencies are familiar with nanoparticles as delivery agents, based in part on the several liposomal drugs now on the market (175). Many of these useful aspects of nanoparticle-mediated oligonucleotide delivery are implicit in the studies reviewed earlier. However, the many positive features of nanoparticles are counterbalanced by some important negative ones. First, despite advances in using PEG or other hydrophilic polymers for stabilization, a large fraction of the injected dose of nanoparticles will accumulate in the liver, and much of that will be taken up by hepatic phagocytes. Thus, a significant portion of the dose of oligonucleotide will wind up at the wrong site, where toxic effects could potentially occur. Second, because of the vascular endothelial barrier, nanoparticles can only reach certain tissues, such as liver and spleen, where gaps in endothelium occur. This is one of the reasons that the various siRNA and antisense companies have focused so strongly on diseases that involve the liver. Nanoparticles that have long circulation times can also accumulate in some types of tumors due to the EPR effect; inclusion of a ligand that binds a receptor on the tumor cells can then promote uptake and intracellular delivery (176). However, what nanoparticles cannot do is to access parenchymal cells in most normal tissues; they are simply excluded by the endothelial barrier. Thus, many potential disease targets will not be addressable by oligonucleotide-based therapeutic platforms that involve nanoparticles.

In contrast, oligonucleotide conjugates are usually far smaller than the pores in normal vascular endothelium; thus, in principle, they should be able to access virtually all tissues, just as conventional drugs do (with the exception of the central nervous system). It should be noted that in some cases the oligonucleotide conjugate may be rapidly excreted via the kidney. Clearly, the detailed physical and chemical properties of individual oligonucleotide conjugates will influence their interactions with plasma proteins and their overall biodistribution; however, to a first approximation, the relatively small size of oligonucleotide conjugates implies a fundamental difference in their in vivo behavior as compared to nanoparticles. Nonetheless, there are some liabilities associated with this approach. First, each conjugate requires a separate synthesis, whereas a particular type of nanoparticle can potentially accomodate a variety of different oligonucleotides. Second, since only a single ligand is conjugated to the oligonucleotide this implies a lower-affinity interaction with target receptors than is the case for multivalent nanoparticles. Another issue concerns release from endosomes subsequent to cell uptake. It is hard to visualize building both a targeting moiety and an endosome escape moiety into a monomolecular oligonucleotide conjugate; at minimum, the chemistry will be quite difficult.

Thus, both nanocarriers and molecular conjugates exhibit pluses and minuses as delivery agents. Ultimately, the most attractive delivery system may turn out to be neither a relatively small monomolecular oligonucleotide conjugate nor a very large nanoparticle or nanocomplex. Rather it may be an intermediate-sized moiety, perhaps comprised of oligonucleotides and targeting agents covalently linked to a small polymer (113) or protein that is large enough to avoid rapid excretion but yet small enough to be able to pass the vascular endothelial barrier. This approach may offer some of the high payload and high-affinity targeting aspects of nanoparticles without the constraints due to relatively large particle size. This is certainly an appealing approach and one that our group is actively pursuing.

An issue that applies to both conjugates and nanocarriers is the choice of accessory ligands. As mentioned earlier, there have been a number of attempts to improve both cellular uptake and endosomal release of oligonucleotides using CPPs. Surprisingly, however, this has not proven very effective, at least for monomolecular oligonucleotide conjugates. On the other hand, studies of polymolecular complexes of CPPs with oligonucleotides have seemed more promising; perhaps it requires multiple copies of CPPs to attain strong endosome destabilizing effects. As discussed earlier, use of cell targeting ligands that bind to specific receptors seems a more productive strategy. However, in most of the examples cited, while the targeting ligand can certainly enhance uptake, it is not clear that endosomal release is also enhanced. Perhaps one aspect of the receptor targeting strategy may entail differential opportunities for release from endosomes as the internalized oligonucleotide trafficks through different endomembrane compartments. Thus, it would be interesting to see if the same oligonucleotide, taken up via two different receptors, had the same or different ultimate biological effect. Certainly, it will be important to use current cell and molecular biological approaches to learn more about the details of intracellular trafficking of oligonucleotides and conjugates.

Another issue concerns the use of conjugates versus complexes. Some of the most exciting in vivo observations to date have involved noncovalent complexes between an oligonucleotide and a delivery agent (109,114). In most cases, however, the stoichiometry and physical characteristics of these complexes are essentially unknown. This raises issues concerning scale-up, reproducibility and the pharmaceutical acceptability of these approaches. Conjugates have the advantage of being well-defined molecular entities that may be easier to move along the path toward large-scale production, commercialization and clinical utilization.

In summary, while much progress has been made in the area of delivery of antisense and siRNA, much remains to be done. It will be important to pursue basic studies concerning both subcellular trafficking of olignucleotides and their detailed biodistribution in animals. A measured approach may, in the long run, serve the field better than hasty attempts to ‘hit a home run’ by bringing poorly characterized delivery strategies prematurely into the clinic.

ACKNOWLEDGEMENTS

This work was supported by NIH grant P01 GM059299 to R.L.J. The authors thank Betsy Clarke for expert editorial assistance. Funding to pay the Open Access publication charges for this article was provided by the NIH.

Conflict of interest statement. None declared.

REFERENCES

1
Nahta
R
Esteva
FJ
Bcl-2 antisense oligonucleotides: a potential novel strategy for the treatment of breast cancer
Semin. Oncol
2003
, vol. 
30
 (pg. 
143
-
149
)
2
Dean
NM
Bennett
CF
Antisense oligonucleotide-based therapeutics for cancer
Oncogene
2003
, vol. 
22
 (pg. 
9087
-
9096
)
3
Chan
JH
Lim
S
Wong
WS
Antisense oligonucleotides: from design to therapeutic application
Clin. Exp. Pharmacol. Physiol
2006
, vol. 
33
 (pg. 
533
-
540
)
4
Coppelli
FM
Grandis
JR
Oligonucleotides as anticancer agents: from the benchside to the clinic and beyond
Curr. Pharm. Des
2005
, vol. 
11
 (pg. 
2825
-
2840
)
5
Manoharan
M
Oligonucleotide conjugates as potential antisense drugs with improved uptake, biodistribution, targeted delivery, and mechanism of action
Antisense Nucleic Acid Drug Dev
2002
, vol. 
12
 (pg. 
103
-
128
)
6
Kurreck
J
Antisense technologies. Improvement through novel chemical modifications
Eur. J. Biochem
2003
, vol. 
270
 (pg. 
1628
-
1644
)
7
Crooke
ST
Progress in antisense technology
Annu. Rev. Med
2004
, vol. 
55
 (pg. 
61
-
95
)
8
McManus
MT
Sharp
PA
Gene silencing in mammals by small interfering RNAs
Nat. Rev. Genet
2002
, vol. 
3
 (pg. 
737
-
747
)
9
Inoue
A
Sawata
SY
Taira
K
Molecular design and delivery of siRNA
J. Drug Target
2006
, vol. 
14
 (pg. 
448
-
455
)
10
Kim
DH
Rossi
JJ
Strategies for silencing human disease using RNA interference
Nat. Rev. Genet
2007
, vol. 
8
 (pg. 
173
-
184
)
11
Esau
CC
Monia
BP
Therapeutic potential for microRNAs
Adv. Drug Deliv. Rev
2007
, vol. 
59
 (pg. 
101
-
114
)
12
Grimm
D
Kay
MA
Therapeutic application of RNAi: is mRNA targeting finally ready for prime time?
J. Clin. Invest
2007
, vol. 
117
 (pg. 
3633
-
3641
)
13
de Fougerolles
A
Vornlocher
HP
Maraganore
J
Lieberman
J
Interfering with disease: a progress report on siRNA-based therapeutics
Nat. Rev. Drug Discov
2007
, vol. 
6
 (pg. 
443
-
453
)
14
Soutschek
J
Akinc
A
Bramlage
B
Charisse
K
Constien
R
Donoghue
M
Elbashir
S
Geick
A
Hadwiger
P
Harborth
J
, et al. 
Therapeutic silencing of an endogenous gene by systemic administration of modified siRNAs
Nature
2004
, vol. 
432
 (pg. 
173
-
178
)
15
Behlke
MA
Progress towards in vivo use of siRNAs
Mol. Ther
2006
, vol. 
13
 (pg. 
644
-
670
)
16
Lv
H
Zhang
S
Wang
B
Cui
S
Yan
J
Toxicity of cationic lipids and cationic polymers in gene delivery
J. Control Release
2006
, vol. 
114
 (pg. 
100
-
109
)
17
Akhtar
S
Benter
I
Toxicogenomics of non-viral drug delivery systems for RNAi: potential impact on siRNA-mediated gene silencing activity and specificity
Adv. Drug Deliv. Rev
2007
, vol. 
59
 (pg. 
164
-
182
)
18
Akhtar
S
Benter
IF
Nonviral delivery of synthetic siRNAs in vivo
J. Clin. Invest
2007
, vol. 
117
 (pg. 
3623
-
3632
)
19
Mescalchin
A
Detzer
A
Wecke
M
Overhoff
M
Wunsche
W
Sczakiel
G
Cellular uptake and intracellular release are major obstacles to the therapeutic application of siRNA: novel options by phosphorothioate-stimulated delivery
Expert Opin. Biol. Ther
2007
, vol. 
7
 (pg. 
1531
-
1538
)
20
Russ
V
Wagner
E
Cell and tissue targeting of nucleic acids for cancer gene therapy
Pharm. Res
2007
, vol. 
24
 (pg. 
1047
-
1057
)
21
Dorsett
Y
Tuschl
T
siRNAs: applications in functional genomics and potential as therapeutics
Nat. Rev. Drug Discov
2004
, vol. 
3
 (pg. 
318
-
329
)
22
Carette
JE
Overmeer
RM
Schagen
FH
Alemany
R
Barski
OA
Gerritsen
WR
Van Beusechem
VW
Conditionally replicating adenoviruses expressing short hairpin RNAs silence the expression of a target gene in cancer cells
Cancer Res
2004
, vol. 
64
 (pg. 
2663
-
2667
)
23
Grimm
D
Streetz
KL
Jopling
CL
Storm
TA
Pandey
K
Davis
CR
Marion
P
Salazar
F
Kay
MA
Fatality in mice due to oversaturation of cellular microRNA/short hairpin RNA pathways
Nature
2006
, vol. 
441
 (pg. 
537
-
541
)
24
Xu
D
McCarty
D
Fernandes
A
Fisher
M
Samulski
RJ
Juliano
RL
Delivery of MDR1 small interfering RNA by self-complementary recombinant adeno-associated virus vector
Mol. Ther
2005
, vol. 
11
 (pg. 
523
-
530
)
25
Sazani
P
Kole
R
Therapeutic potential of antisense oligonucleotides as modulators of alternative splicing
J. Clin. Invest
2003
, vol. 
112
 (pg. 
481
-
486
)
26
Valencia-Sanchez
MA
Liu
J
Hannon
GJ
Parker
R
Control of translation and mRNA degradation by miRNAs and siRNAs
Genes Dev
2006
, vol. 
20
 (pg. 
515
-
524
)
27
Hannon
GJ
RNA interference
Nature
2002
, vol. 
418
 (pg. 
244
-
251
)
28
Paddison
PJ
Hannon
GJ
RNA interference: the new somatic cell genetics?
Cancer Cell
2002
, vol. 
2
 (pg. 
17
-
23
)
29
Hammond
SM
Dicing and slicing: the core machinery of the RNA interference pathway
FEBS Lett
2005
, vol. 
579
 (pg. 
5822
-
5829
)
30
Ameres
SL
Martinez
J
Schroeder
R
Molecular basis for target RNA recognition and cleavage by human RISC
Cell
2007
, vol. 
130
 (pg. 
101
-
112
)
31
Eulalio
A
Behm-Ansmant
I
Izaurralde
E
P bodies: at the crossroads of post-transcriptional pathways
Nat. Rev. Mol. Cell Biol
2007
, vol. 
8
 (pg. 
9
-
22
)
32
Buhler
M
Moazed
D
Transcription and RNAi in heterochromatic gene silencing
Nat. Struct. Mol. Biol
2007
, vol. 
14
 (pg. 
1041
-
1048
)
33
Czech
MP
MicroRNAs as therapeutic targets
N. Engl. J. Med
2006
, vol. 
354
 (pg. 
1194
-
1195
)
34
Krutzfeldt
J
Kuwajima
S
Braich
R
Rajeev
KG
Pena
J
Tuschl
T
Manoharan
M
Stoffel
M
Specificity, duplex degradation and subcellular localization of antagomirs
Nucleic Acids Res
2007
, vol. 
35
 (pg. 
2885
-
2892
)
35
Krutzfeldt
J
Rajewsky
N
Braich
R
Rajeev
KG
Tuschl
T
Manoharan
M
Stoffel
M
Silencing of microRNAs in vivo with ‘antagomirs’
Nature
2005
, vol. 
438
 (pg. 
685
-
689
)
36
Elbashir
SM
Harborth
J
Lendeckel
W
Yalcin
A
Weber
K
Tuschl
T
Duplexes of 21-nucleotide RNAs mediate RNA interference in cultured mammalian cells
Nature
2001
, vol. 
411
 (pg. 
494
-
498
)
37
Corey
DR
RNA learns from antisense
Nat. Chem. Biol
2007
, vol. 
3
 (pg. 
8
-
11
)
38
De Paula
D
Bentley
MV
Mahato
RI
Hydrophobization and bioconjugation for enhanced siRNA delivery and targeting
RNA
2007
, vol. 
13
 (pg. 
431
-
456
)
39
Fisher
M
Abramov
M
Van Aerschot
A
Xu
D
Juliano
RL
Herdewijn
P
Inhibition of MDR1 expression with altritol-modified siRNAs
Nucleic Acids Res
2007
, vol. 
35
 (pg. 
1064
-
1074
)
40
Manoharan
M
RNA interference and chemically modified small interfering RNAs
Curr. Opin. Chem. Biol
2004
, vol. 
8
 (pg. 
570
-
579
)
41
Corey
DR
Chemical modification: the key to clinical application of RNA interference?
J. Clin. Invest
2007
, vol. 
117
 (pg. 
3615
-
3622
)
42
Reynolds
A
Leake
D
Boese
Q
Scaringe
S
Marshall
WS
Khvorova
A
Rational siRNA design for RNA interference
Nat. Biotechnol
2004
, vol. 
22
 (pg. 
326
-
330
)
43
Kleinman
ME
Yamada
K
Takeda
A
Chandrasekaran
V
Nozaki
M
Baffi
JZ
Albuquerque
RJ
Yamasaki
S
Itaya
M
Pan
Y
, et al. 
Sequence- and target-independent angiogenesis suppression by siRNA via TLR3
Nature
2008
, vol. 
452
 (pg. 
591
-
597
)
44
Kirkham
M
Parton
RG
Clathrin-independent endocytosis: new insights into caveolae and non-caveolar lipid raft carriers
Biochim. Biophys. Acta
2005
, vol. 
1745
 (pg. 
273
-
286
)
45
Perret
E
Lakkaraju
A
Deborde
S
Schreiner
R
Rodriguez-Boulan
E
Evolving endosomes: how many varieties and why?
Curr. Opin. Cell Biol
2005
, vol. 
17
 (pg. 
423
-
434
)
46
Echarri
A
Muriel
O
Del Pozo
MA
Intracellular trafficking of raft/caveolae domains: insights from integrin signaling
Semin. Cell Dev. Biol
2007
, vol. 
18
 (pg. 
627
-
637
)
47
Parton
RG
Simons
K
The multiple faces of caveolae
Nat. Rev. Mol. Cell Biol
2007
, vol. 
8
 (pg. 
185
-
194
)
48
Pfeffer
SR
Unsolved mysteries in membrane traffic
Annu. Rev. Biochem
2007
, vol. 
76
 (pg. 
629
-
645
)
49
Cai
H
Reinisch
K
Ferro-Novick
S
Coats, tethers, Rabs, and SNAREs work together to mediate the intracellular destination of a transport vesicle
Dev. Cell
2007
, vol. 
12
 (pg. 
671
-
682
)
50
van der Goot
FG
Gruenberg
J
Intra-endosomal membrane traffic
Trends Cell Biol
2006
, vol. 
16
 (pg. 
514
-
521
)
51
Grosshans
B
Ortiz
D
Novick
P
Rabs and their effectors: achieving specificity in membrane traffic
Proc. Natl Acad. Sci. USA
2006
, vol. 
103
 (pg. 
11821
-
11827
)
52
Carlton
J
Bujny
M
Rutherford
A
Cullen
P
Sorting nexins—unifying trends and new perspectives
Traffic
2005
, vol. 
6
 (pg. 
75
-
82
)
53
Di Guglielmo
GM
Le Roy
C
Goodfellow
AF
Wrana
JL
Distinct endocytic pathways regulate TGF-beta receptor signalling and turnover
Nat. Cell Biol
2003
, vol. 
5
 (pg. 
410
-
421
)
54
Chen
H
De Camilli
P
The association of epsin with ubiquitinated cargo along the endocytic pathway is negatively regulated by its interaction with clathrin
Proc. Natl Acad. Sci. USA
2005
, vol. 
102
 (pg. 
2766
-
2771
)
55
Subtil
A
Gaidarov
I
Kobylarz
K
Lampson
MA
Keen
JH
McGraw
TE
Acute cholesterol depletion inhibits clathrin-coated pit budding
Proc. Natl Acad. Sci. USA
1999
, vol. 
96
 (pg. 
6775
-
6780
)
56
Pelkmans
L
Puntener
D
Helenius
A
Local actin polymerization and dynamin recruitment in SV40-induced internalization of caveolae
Science
2002
, vol. 
296
 (pg. 
535
-
539
)
57
Aplin
AE
Juliano
RL
Integrin and cytoskeletal regulation of growth factor signaling to the MAP kinase pathway
J. Cell Sci
1999
, vol. 
112
 
Pt 5
(pg. 
695
-
706
)
58
Marbet
P
Rahner
C
Stieger
B
Landmann
L
Quantitative microscopy reveals 3D organization and kinetics of endocytosis in rat hepatocytes
Microsc. Res. Tech
2006
, vol. 
69
 (pg. 
693
-
707
)
59
Zerial
M
McBride
H
Rab proteins as membrane organizers
Nat. Rev. Mol. Cell Biol
2001
, vol. 
2
 (pg. 
107
-
117
)
60
Juliano
RL
Alahari
S
Yoo
H
Kole
R
Cho
M
Antisense pharmacodynamics: critical issues in the transport and delivery of antisense oligonucleotides
Pharm. Res
1999
, vol. 
16
 (pg. 
494
-
502
)
61
Wang
L
Prakash
RK
Stein
CA
Koehn
RK
Ruffner
DE
Progress in the delivery of therapeutic oligonucleotides: organ/cellular distribution and targeted delivery of oligonucleotides in vivo
Antisense Nucleic Acid Drug Dev
2003
, vol. 
13
 (pg. 
169
-
189
)
62
Alam
MR
Dixit
V
Kang
H
Li
Z-B
Chen
X
Trejo
J
Fisher
M
Juliano
RL
Intracellular delivery of an anionic antisense oligonucleotide via receptor mediated endocytosis
Nucleic Acids Res
2008
, vol. 
36
 (pg. 
2764
-
2776
)
63
El-Andaloussi
S
Johansson
HJ
Holm
T
Langel
U
A novel cell-penetrating peptide, M918, for efficient delivery of proteins and peptide nucleic acids
Mol. Ther
2007
, vol. 
15
 (pg. 
1820
-
1826
)
64
Overhoff
M
Sczakiel
G
Phosphorothioate-stimulated uptake of short interfering RNA by human cells
EMBO Rep
2005
, vol. 
6
 (pg. 
1176
-
1181
)
65
Dietz
GP
Bahr
M
Delivery of bioactive molecules into the cell: the Trojan horse approach
Mol. Cell Neurosci
2004
, vol. 
27
 (pg. 
85
-
131
)
66
Patel
LN
Zaro
JL
Shen
WC
Cell penetrating peptides: intracellular pathways and pharmaceutical perspectives
Pharm. Res
2007
, vol. 
24
 (pg. 
1977
-
1992
)
67
Gupta
B
Levchenko
TS
Torchilin
VP
Intracellular delivery of large molecules and small particles by cell-penetrating proteins and peptides
Adv. Drug Deliv. Rev
2005
, vol. 
57
 (pg. 
637
-
651
)
68
Howl
J
Nicholl
ID
Jones
S
The many futures for cell-penetrating peptides: how soon is now?
Biochem. Soc. Trans
2007
, vol. 
35
 (pg. 
767
-
769
)
69
Meade
BR
Dowdy
SF
Exogenous siRNA delivery using peptide transduction domains/cell penetrating peptides
Adv. Drug Deliv. Rev
2007
, vol. 
59
 (pg. 
134
-
140
)
70
Duchardt
F
Fotin-Mleczek
M
Schwarz
H
Fischer
R
Brock
R
A comprehensive model for the cellular uptake of cationic cell-penetrating peptides
Traffic
2007
, vol. 
8
 (pg. 
848
-
866
)
71
Trehin
R
Merkle
HP
Chances and pitfalls of cell penetrating peptides for cellular drug delivery
Eur. J. Pharm. Biopharm
2004
, vol. 
58
 (pg. 
209
-
223
)
72
Tunnemann
G
Martin
RM
Haupt
S
Patsch
C
Edenhofer
F
Cardoso
MC
Cargo-dependent mode of uptake and bioavailability of TAT-containing proteins and peptides in living cells
FASEB J
2006
, vol. 
20
 (pg. 
1775
-
1784
)
73
Barany-Wallje
E
Gaur
J
Lundberg
P
Langel
U
Graslund
A
Differential membrane perturbation caused by the cell penetrating peptide Tp10 depending on attached cargo
FEBS Lett
2007
, vol. 
581
 (pg. 
2389
-
2393
)
74
El-Andaloussi
S
Jarver
P
Johansson
HJ
Langel
U
Cargo-dependent cytotoxicity and delivery efficacy of cell-penetrating peptides: a comparative study
Biochem. J
2007
, vol. 
407
 (pg. 
285
-
292
)
75
Vives
E
Richard
JP
Rispal
C
Lebleu
B
TAT peptide internalization: seeking the mechanism of entry
Curr. Protein Pept. Sci
2003
, vol. 
4
 (pg. 
125
-
132
)
76
Holliger
P
Hudson
PJ
Engineered antibody fragments and the rise of single domains
Nat. Biotechnol
2005
, vol. 
23
 (pg. 
1126
-
1136
)
77
Ruoslahti
E
Vascular zip codes in angiogenesis and metastasis
Biochem. Soc. Trans
2004
, vol. 
32
 (pg. 
397
-
402
)
78
Brown
MS
Goldstein
JL
A receptor-mediated pathway for cholesterol homeostasis
Science
1986
, vol. 
232
 (pg. 
34
-
47
)
79
Hynes
RO
Integrins: bidirectional, allosteric signaling machines
Cell
2002
, vol. 
110
 (pg. 
673
-
687
)
80
Juliano
RL
Signal transduction by cell adhesion receptors and the cytoskeleton: functions of integrins, cadherins, selectins, and immunoglobulin-superfamily members
Annu. Rev. Pharmacol. Toxicol
2002
, vol. 
42
 (pg. 
283
-
323
)
81
Schlessinger
J
Cell signaling by receptor tyrosine kinases
Cell
2000
, vol. 
103
 (pg. 
211
-
225
)
82
Armbruster
BN
Roth
BL
Mining the receptorome
J. Biol. Chem
2005
, vol. 
280
 (pg. 
5129
-
5132
)
83
Lundstrom
K
Latest development in drug discovery on G protein-coupled receptors
Curr. Protein Pept. Sci
2006
, vol. 
7
 (pg. 
465
-
470
)
84
Akhtar
S
Basu
S
Wickstrom
E
Juliano
RL
Interactions of antisense DNA oligonucleotide analogs with phospholipid membranes (liposomes)
Nucleic Acids Res
1991
, vol. 
19
 (pg. 
5551
-
5559
)
85
Wittung
P
Kajanus
J
Edwards
K
Haaima
G
Nielsen
PE
Norden
B
Malmstrom
BG
Phospholipid membrane permeability of peptide nucleic acid
FEBS Lett
1995
, vol. 
375
 (pg. 
27
-
29
)
86
Juliano
RL
Peptide-oligonucleotide conjugates for the delivery of antisense and siRNA
Curr. Opin. Mol. Ther
2005
, vol. 
7
 (pg. 
132
-
136
)
87
Abes
S
Moulton
H
Turner
J
Clair
P
Richard
JP
Iversen
P
Gait
MJ
Lebleu
B
Peptide-based delivery of nucleic acids: design, mechanism of uptake and applications to splice-correcting oligonucleotides
Biochem. Soc. Trans
2007
, vol. 
35
 (pg. 
53
-
55
)
88
Astriab-Fisher
A
Sergueev
D
Fisher
M
Shaw
BR
Juliano
RL
Conjugates of antisense oligonucleotides with the Tat and antennapedia cell-penetrating peptides: effects on cellular uptake, binding to target sequences, and biologic actions
Pharm. Res
2002
, vol. 
19
 (pg. 
744
-
754
)
89
Astriab-Fisher
A
Sergueev
DS
Fisher
M
Shaw
BR
Juliano
RL
Antisense inhibition of P-glycoprotein expression using peptide-oligonucleotide conjugates
Biochem. Pharmacol
2000
, vol. 
60
 (pg. 
83
-
90
)
90
Muratovska
A
Eccles
MR
Conjugate for efficient delivery of short interfering RNA (siRNA) into mammalian cells
FEBS Lett
2004
, vol. 
558
 (pg. 
63
-
68
)
91
Chiu
YL
Ali
A
Chu
CY
Cao
H
Rana
TM
Visualizing a correlation between siRNA localization, cellular uptake, and RNAi in living cells
Chem. Biol
2004
, vol. 
11
 (pg. 
1165
-
1175
)
92
Turner
JJ
Ivanova
GD
Verbeure
B
Williams
D
Arzumanov
AA
Abes
S
Lebleu
B
Gait
MJ
Cell-penetrating peptide conjugates of peptide nucleic acids (PNA) as inhibitors of HIV-1 Tat-dependent trans-activation in cells
Nucleic Acids Res
2005
, vol. 
33
 (pg. 
6837
-
6849
)
93
El-Andaloussi
S
Johansson
HJ
Lundberg
P
Langel
U
Induction of splice correction by cell-penetrating peptide nucleic acids
J. Gene Med
2006
, vol. 
8
 (pg. 
1262
-
1273
)
94
Bendifallah
N
Rasmussen
FW
Zachar
V
Ebbesen
P
Nielsen
PE
Koppelhus
U
Evaluation of cell-penetrating peptides (CPPs) as vehicles for intracellular delivery of antisense peptide nucleic acid (PNA)
Bioconjug. Chem
2006
, vol. 
17
 (pg. 
750
-
758
)
95
Moulton
HM
Hase
MC
Smith
KM
Iversen
PL
HIV Tat peptide enhances cellular delivery of antisense morpholino oligomers
Antisense Nucleic Acid Drug Dev
2003
, vol. 
13
 (pg. 
31
-
43
)
96
Abes
S
Turner
JJ
Ivanova
GD
Owen
D
Williams
D
Arzumanov
A
Clair
P
Gait
MJ
Lebleu
B
Efficient splicing correction by PNA conjugation to an R6-Penetratin delivery peptide
Nucleic Acids Res
2007
, vol. 
35
 (pg. 
4495
-
4502
)
97
Fisher
AA
Ye
D
Sergueev
DS
Fisher
MH
Shaw
BR
Juliano
RL
Evaluating the specificity of antisense oligonucleotide conjugates. A DNA array analysis
J. Biol. Chem
2002
, vol. 
277
 (pg. 
22980
-
22984
)
98
Turner
JJ
Arzumanov
AA
Gait
MJ
Synthesis, cellular uptake and HIV-1 Tat-dependent trans-activation inhibition activity of oligonucleotide analogues disulphide-conjugated to cell-penetrating peptides
Nucleic Acids Res
2005
, vol. 
33
 (pg. 
27
-
42
)
99
Lundberg
P
El-Andaloussi
S
Sutlu
T
Johansson
H
Langel
U
Delivery of short interfering RNA using endosomolytic cell-penetrating peptides
FASEB J
2007
, vol. 
21
 (pg. 
2664
-
2671
)
100
Simeoni
F
Morris
MC
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
, vol. 
31
 (pg. 
2717
-
2724
)
101
Veldhoen
S
Laufer
SD
Trampe
A
Restle
T
Cellular delivery of small interfering RNA by a non-covalently attached cell-penetrating peptide: quantitative analysis of uptake and biological effect
Nucleic Acids Res
2006
, vol. 
34
 (pg. 
6561
-
6573
)
102
Hu
J
Corey
DR
Inhibiting gene expression with peptide nucleic acid (PNA)—peptide conjugates that target chromosomal DNA
Biochemistry
2007
, vol. 
46
 (pg. 
7581
-
7589
)
103
Moulton
HM
Nelson
MH
Hatlevig
SA
Reddy
MT
Iversen
PL
Cellular uptake of antisense morpholino oligomers conjugated to arginine-rich peptides
Bioconjug. Chem
2004
, vol. 
15
 (pg. 
290
-
299
)
104
Wu
RP
Youngblood
DS
Hassinger
JN
Lovejoy
CE
Nelson
MH
Iversen
PL
Moulton
HM
Cell-penetrating peptides as transporters for morpholino oligomers: effects of amino acid composition on intracellular delivery and cytotoxicity
Nucleic Acids Res
2007
, vol. 
35
 (pg. 
5182
-
5191
)
105
Youngblood
DS
Hatlevig
SA
Hassinger
JN
Iversen
PL
Moulton
HM
Stability of cell-penetrating peptide-morpholino oligomer conjugates in human serum and in cells
Bioconjug. Chem
2007
, vol. 
18
 (pg. 
50
-
60
)
106
Abes
R
Arzumanov
AA
Moulton
HM
Abes
S
Ivanova
GD
Iversen
PL
Gait
MJ
Lebleu
B
Cell-penetrating-peptide-based delivery of oligonucleotides: an overview
Biochem. Soc. Trans
2007
, vol. 
35
 (pg. 
775
-
779
)
107
McNamara
JOII
Andrechek
ER
Wang
Y
Viles
KD
Rempel
RE
Gilboa
E
Sullenger
BA
Giangrande
PH
Cell type-specific delivery of siRNAs with aptamer-siRNA chimeras
Nat. Biotechnol
2006
, vol. 
24
 (pg. 
1005
-
1015
)
108
Chu
TC
Twu
KY
Ellington
AD
Levy
M
Aptamer mediated siRNA delivery
Nucleic Acids Res
2006
, vol. 
34
 pg. 
e73
 
109
Song
E
Zhu
P
Lee
SK
Chowdhury
D
Kussman
S
Dykxhoorn
DM
Feng
Y
Palliser
D
Weiner
DB
Shankar
P
, et al. 
Antibody mediated in vivo delivery of small interfering RNAs via cell-surface receptors
Nat. Biotechnol
2005
, vol. 
23
 (pg. 
709
-
717
)
110
Peer
D
Zhu
P
Carman
CV
Lieberman
J
Shimaoka
M
Selective gene silencing in activated leukocytes by targeting siRNAs to the integrin lymphocyte function-associated antigen-1
Proc. Natl Acad. Sci. USA
2007
, vol. 
104
 (pg. 
4095
-
4100
)
111
Tian
X
Aruva
MR
Qin
W
Zhu
W
Sauter
ER
Thakur
ML
Wickstrom
E
Noninvasive molecular imaging of MYC mRNA expression in human breast cancer xenografts with a [99mTc]peptide-peptide nucleic acid-peptide chimera
Bioconjug. Chem
2005
, vol. 
16
 (pg. 
70
-
79
)
112
Cesarone
G
Edupuganti
OP
Chen
CP
Wickstrom
E
Insulin receptor substrate 1 knockdown in human MCF7 ER+ breast cancer cells by nuclease-resistant IRS1 siRNA conjugated to a disulfide-bridged D-peptide analogue of insulin-like growth factor 1
Bioconjug. Chem
2007
, vol. 
18
 (pg. 
1831
-
1840
)
113
Rozema
DB
Lewis
DL
Wakefield
DH
Wong
SC
Klein
JJ
Roesch
PL
Bertin
SL
Reppen
TW
Chu
Q
Blokhin
AV
, et al. 
Dynamic PolyConjugates for targeted in vivo delivery of siRNA to hepatocytes
Proc. Natl Acad. Sci. USA
2007
, vol. 
104
 (pg. 
12982
-
12987
)
114
Kumar
P
Wu
H
McBride
JL
Jung
KE
Kim
MH
Davidson
BL
Lee
SK
Shankar
P
Manjunath
N
Transvascular delivery of small interfering RNA to the central nervous system
Nature
2007
, vol. 
448
 (pg. 
39
-
43
)
115
Drummond
DC
Meyer
O
Hong
K
Kirpotin
DB
Papahadjopoulos
D
Optimizing liposomes for delivery of chemotherapeutic agents to solid tumors
Pharmacol. Rev
1999
, vol. 
51
 (pg. 
691
-
743
)
116
Tiera
MJ
Winnik
FO
Fernandes
JC
Synthetic and natural polycations for gene therapy: state of the art and new perspectives
Curr. Gene Ther
2006
, vol. 
6
 (pg. 
59
-
71
)
117
Pridgen
EM
Langer
R
Farokhzad
OC
Biodegradable, polymeric nanoparticle delivery systems for cancer therapy
Nanomed
2007
, vol. 
2
 (pg. 
669
-
680
)
118
Gilmore
IR
Fox
SP
Hollins
AJ
Akhtar
S
Delivery strategies for siRNA-mediated gene silencing
Curr. Drug Deliv
2006
, vol. 
3
 (pg. 
147
-
145
)
119
Li
W
Szoka
FC
Jr
Lipid-based nanoparticles for nucleic acid delivery
Pharm. Res
2007
, vol. 
24
 (pg. 
438
-
449
)
120
Fattal
E
Couvreur
P
Dubernet
C
“Smart” delivery of antisense oligonucleotides by anionic pH-sensitive liposomes
Adv. Drug Deliv. Rev
2004
, vol. 
56
 (pg. 
931
-
946
)
121
Vinogradov
SV
Batrakova
EV
Kabanov
AV
Nanogels for oligonucleotide delivery to the brain
Bioconjug. Chem
2004
, vol. 
15
 (pg. 
50
-
60
)
122
Meyer
M
Wagner
E
Recent developments in the application of plasmid DNA-based vectors and small interfering RNA therapeutics for cancer
Hum. Gene Ther
2006
, vol. 
17
 (pg. 
1062
-
1076
)
123
Meyer
M
Philipp
A
Oskuee
R
Schmidt
C
Wagner
E
Breathing life into polycations: functionalization with ph-responsive endosomolytic peptides and polyethylene glycol enables siRNA delivery
J. Am. Chem. Soc
2008
, vol. 
130
 (pg. 
3272
-
3273
)
124
Grzelinski
M
Urban-Klein
B
Martens
T
Lamszus
K
Bakowsky
U
Hobel
S
Czubayko
F
Aigner
A
RNA interference-mediated gene silencing of pleiotrophin through polyethylenimine-complexed small interfering RNAs in vivo exerts antitumoral effects in glioblastoma xenografts
Hum. Gene Ther
2006
, vol. 
17
 (pg. 
751
-
766
)
125
Schiffelers
RM
Ansari
A
Xu
J
Zhou
Q
Tang
Q
Storm
G
Molema
G
Lu
PY
Scaria
PV
Woodle
MC
Cancer siRNA therapy by tumor selective delivery with ligand-targeted sterically stabilized nanoparticle
Nucleic Acids Res
2004
, vol. 
32
 pg. 
e149
 
126
DeRouchey
J
Schmidt
C
Walker
GF
Koch
C
Plank
C
Wagner
E
Radler
JO
Monomolecular assembly of siRNA and poly(ethylene glycol)-peptide copolymers
Biomacromolecules
2008
, vol. 
9
 (pg. 
724
-
732
)
127
Kang
H
DeLong
R
Fisher
MH
Juliano
RL
Tat-conjugated PAMAM dendrimers as delivery agents for antisense and siRNA oligonucleotides
Pharm. Res
2005
, vol. 
22
 (pg. 
2099
-
2106
)
128
Tsutsumi
T
Hirayama
F
Uekama
K
Arima
H
Evaluation of polyamidoamine dendrimer/alpha-cyclodextrin conjugate (generation 3, G3) as a novel carrier for small interfering RNA (siRNA)
J. Control Release
2007
, vol. 
119
 (pg. 
349
-
359
)
129
Yoo
H
Juliano
RL
Enhanced delivery of antisense oligonucleotides with fluorophore-conjugated PAMAM dendrimers
Nucleic Acids Res
2000
, vol. 
28
 (pg. 
4225
-
4231
)
130
Zhou
J
Wu
J
Hafdi
N
Behr
JP
Erbacher
P
Peng
L
PAMAM dendrimers for efficient siRNA delivery and potent gene silencing
Chem. Commun
2006
(pg. 
2362
-
2364
)
131
Boussif
O
Lezoualc'h
F
Zanta
MA
Mergny
MD
Scherman
D
Demeneix
B
Behr
JP
A versatile vector for gene and oligonucleotide transfer into cells in culture and in vivo: polyethylenimine
Proc. Natl Acad. Sci. USA
1995
, vol. 
92
 (pg. 
7297
-
7301
)
132
Toub
N
Bertrand
JR
Tamaddon
A
Elhamess
H
Hillaireau
H
Maksimenko
A
Maccario
J
Malvy
C
Fattal
E
Couvreur
P
Efficacy of siRNA nanocapsules targeted against the EWS-Fli1 oncogene in Ewing sarcoma
Pharm. Res
2006
, vol. 
23
 (pg. 
892
-
900
)
133
Kataoka
K
Itaka
K
Nishiyama
N
Yamasaki
Y
Oishi
M
Nagasaki
Y
Smart polymeric micelles as nanocarriers for oligonucleotides and siRNA delivery
Nucleic Acids Symp. Ser
2005
, vol. 
49
 (pg. 
17
-
18
)
134
Derfus
AM
Chen
AA
Min
DH
Ruoslahti
E
Bhatia
SN
Targeted quantum dot conjugates for siRNA delivery
Bioconjug. Chem
2007
, vol. 
18
 (pg. 
1391
-
1396
)
135
Tan
WB
Jiang
S
Zhang
Y
Quantum-dot based nanoparticles for targeted silencing of HER2/neu gene via RNA interference
Biomaterials
2007
, vol. 
28
 (pg. 
1565
-
1571
)
136
Santel
A
Aleku
M
Keil
O
Endruschat
J
Esche
V
Durieux
B
Loffler
K
Fechtner
M
Rohl
T
Fisch
G
, et al. 
RNA interference in the mouse vascular endothelium by systemic administration of siRNA-lipoplexes for cancer therapy
Gene Ther
2006
, vol. 
13
 (pg. 
1360
-
1370
)
137
Santel
A
Aleku
M
Keil
O
Endruschat
J
Esche
V
Fisch
G
Dames
S
Loffler
K
Fechtner
M
Arnold
W
, et al. 
A novel siRNA-lipoplex technology for RNA interference in the mouse vascular endothelium
Gene Ther
2006
, vol. 
13
 (pg. 
1222
-
1234
)
138
Zelphati
O
Szoka
FC
Jr
Mechanism of oligonucleotide release from cationic liposomes
Proc. Natl Acad. Sci. USA
1996
, vol. 
93
 (pg. 
11493
-
11498
)
139
Oishi
M
Nagasaki
Y
Itaka
K
Nishiyama
N
Kataoka
K
Lactosylated poly(ethylene glycol)-siRNA conjugate through acid-labile beta-thiopropionate linkage to construct pH-sensitive polyion complex micelles achieving enhanced gene silencing in hepatoma cells
J. Am. Chem. Soc
2005
, vol. 
127
 (pg. 
1624
-
1625
)
140
Kabanov
AV
Polymer genomics: an insight into pharmacology and toxicology of nanomedicines
Adv. Drug Deliv. Rev
2006
, vol. 
58
 (pg. 
1597
-
1621
)
141
Nelson
CM
Bissell
MJ
Of extracellular matrix, scaffolds, and signaling: tissue architecture regulates development, homeostasis, and cancer
Annu. Rev. Cell Dev. Biol
2006
, vol. 
22
 (pg. 
287
-
309
)
142
Juliano
RL
Niemeyer
CM
Mirkin
CA
Biological Barriers to Nanocarrier-Mediated Delivery of Therapeutic and Imaging Agents
Nanobiotechnology II.
2007
Weinheim, Germany
Wiley-VCH
(pg. 
263
-
278
)
143
Brenner
BM
Deen
WM
Robertson
CR
Determinants of glomerular filtration rate
Annu. Rev. Physiol
1976
, vol. 
38
 (pg. 
11
-
19
)
144
Geary
RS
Watanabe
TA
Truong
L
Freier
S
Lesnik
EA
Sioufi
NB
Sasmor
H
Manoharan
M
Levin
AA
Pharmacokinetic properties of 2′-O-(2-methoxyethyl)-modified oligonucleotide analogs in rats
J. Pharmacol. Exp. Ther
2001
, vol. 
296
 (pg. 
890
-
897
)
145
Rippe
B
Rosengren
BI
Carlsson
O
Venturoli
D
Transendothelial transport: the vesicle controversy
J. Vasc. Res
2002
, vol. 
39
 (pg. 
375
-
390
)
146
Lee
CC
MacKay
JA
Frechet
JM
Szoka
FC
Designing dendrimers for biological applications
Nat. Biotechnol
2005
, vol. 
23
 (pg. 
1517
-
1526
)
147
Uzgiris
E
The role of molecular conformation on tumor uptake of polymeric contrast agents
Invest. Radiol
2004
, vol. 
39
 (pg. 
131
-
137
)
148
Scherphof
GL
Juliano
RL
In vivo behavior of liposomes: interaction with themononuclear phagocyte system and implications for drug targeting
Targeted Drug Delivery.
1991
Berlin
Springer
(pg. 
285
-
313
)
149
Greish
K
Enhanced permeability and retention of macromolecular drugs in solid tumors: a royal gate for targeted anticancer nanomedicines
J. Drug Target
2007
, vol. 
15
 (pg. 
457
-
464
)
150
Jain
RK
Transport of molecules, particles, and cells in solid tumors
Annu. Rev. Biomed. Eng
1999
, vol. 
1
 (pg. 
241
-
263
)
151
Kobayashi
H
Boelte
KC
Lin
PC
Endothelial cell adhesion molecules and cancer progression
Curr. Med. Chem
2007
, vol. 
14
 (pg. 
377
-
386
)
152
Serini
G
Napione
L
Bussolino
F
Integrins team up with tyrosine kinase receptors and plexins to control angiogenesis
Curr. Opin. Hematol
2008
, vol. 
15
 (pg. 
235
-
242
)
153
Cai
W
Shin
DW
Chen
K
Gheysens
O
Cao
Q
Wang
SX
Gambhir
SS
Chen
X
Peptide-labeled near-infrared quantum dots for imaging tumor vasculature in living subjects
Nano Lett
2006
, vol. 
6
 (pg. 
669
-
676
)
154
Li
ZB
Cai
W
Cao
Q
Chen
K
Wu
Z
He
L
Chen
X
(64)Cu-labeled tetrameric and octameric RGD peptides for small-animal PET of tumor alpha(v)beta(3) integrin expression
J. Nucl. Med
2007
, vol. 
48
 (pg. 
1162
-
1171
)
155
Temming
K
Lacombe
M
van der Hoeven
P
Prakash
J
Gonzalo
T
Dijkers
EC
Orfi
L
Keri
G
Poelstra
K
Molema
G
, et al. 
Delivery of the p38 MAPkinase inhibitor SB202190 to angiogenic endothelial cells: development of novel RGD-equipped and PEGylated drug-albumin conjugates using platinum(II)-based drug linker technology
Bioconjug. Chem
2006
, vol. 
17
 (pg. 
1246
-
1255
)
156
Aderem
A
Underhill
DM
Mechanisms of phagocytosis in macrophages
Annu. Rev. Immunol
1999
, vol. 
17
 (pg. 
593
-
623
)
157
Underhill
DM
Ozinsky
A
Phagocytosis of microbes: complexity in action
Annu. Rev. Immunol
2002
, vol. 
20
 (pg. 
825
-
852
)
158
van Vlerken
LE
Vyas
TK
Amiji
MM
Poly(ethylene glycol)-modified nanocarriers for tumor-targeted and intracellular delivery
Pharm. Res
2007
, vol. 
24
 (pg. 
1405
-
1414
)
159
Amantana
A
Moulton
HM
Cate
ML
Reddy
MT
Whitehead
T
Hassinger
JN
Youngblood
DS
Iversen
PL
Pharmacokinetics, biodistribution, stability and toxicity of a cell-penetrating peptide-morpholino oligomer conjugate
Bioconjug. Chem
2007
, vol. 
18
 (pg. 
1325
-
1331
)
160
Wolfrum
C
Shi
S
Jayaprakash
KN
Jayaraman
M
Wang
G
Pandey
RK
Rajeev
KG
Nakayama
T
Charrise
K
Ndungo
EM
, et al. 
Mechanisms and optimization of in vivo delivery of lipophilic siRNAs
Nat. Biotechnol
2007
, vol. 
25
 (pg. 
1149
-
1157
)
161
Duxbury
MS
Ashley
SW
Whang
EE
RNA interference: a mammalian SID-1 homologue enhances siRNA uptake and gene silencing efficacy in human cells
Biochem. Biophys. Res. Commun
2005
, vol. 
331
 (pg. 
459
-
463
)
162
Tsang
SY
Moore
JC
Huizen
RV
Chan
CW
Li
RA
Ectopic expression of systemic RNA interference defective protein in embryonic stem cells
Biochem. Biophys. Res. Commun
2007
, vol. 
357
 (pg. 
480
-
486
)
163
Moschos
SA
Jones
SW
Perry
MM
Williams
AE
Erjefalt
JS
Turner
JJ
Barnes
PJ
Sproat
BS
Gait
MJ
Lindsay
MA
Lung delivery studies using siRNA conjugated to TAT(48-60) and penetratin reveal peptide induced reduction in gene expression and induction of innate immunity
Bioconjug. Chem
2007
, vol. 
18
 (pg. 
1450
-
1459
)
164
Nishina
K
Unno
T
Uno
Y
Kubodera
T
Kanouchi
T
Mizusawa
H
Yokota
T
Efficient in vivo delivery of siRNA to the liver by conjugation of alpha-tocopherol
Mol. Ther
2008
, vol. 
16
 (pg. 
734
-
740
)
165
Hu-Lieskovan
S
Heidel
JD
Bartlett
DW
Davis
ME
Triche
TJ
Sequence-specific knockdown of EWS-FLI1 by targeted, nonviral delivery of small interfering RNA inhibits tumor growth in a murine model of metastatic Ewing's sarcoma
Cancer Res
2005
, vol. 
65
 (pg. 
8984
-
8992
)
166
Heidel
JD
Yu
Z
Liu
JY
Rele
SM
Liang
Y
Zeidan
RK
Kornbrust
DJ
Davis
ME
Administration in non-human primates of escalating intravenous doses of targeted nanoparticles containing ribonucleotide reductase subunit M2 siRNA
Proc. Natl Acad. Sci. USA
2007
, vol. 
104
 (pg. 
5715
-
5721
)
167
Li
SD
Chen
YC
Hackett
MJ
Huang
L
Tumor-targeted delivery of siRNA by self-assembled nanoparticles
Mol. Ther
2008
, vol. 
16
 (pg. 
163
-
169
)
168
Peer
D
Park
EJ
Morishita
Y
Carman
CV
Shimaoka
M
Systemic leukocyte-directed siRNA delivery revealing cyclin D1 as an anti-inflammatory target
Science
2008
, vol. 
319
 (pg. 
627
-
630
)
169
Zimmermann
TS
Lee
AC
Akinc
A
Bramlage
B
Bumcrot
D
Fedoruk
MN
Harborth
J
Heyes
JA
Jeffs
LB
John
M
, et al. 
RNAi-mediated gene silencing in non-human primates
Nature
2006
, vol. 
441
 (pg. 
111
-
114
)
170
Sato
Y
Murase
K
Kato
J
Kobune
M
Sato
T
Kawano
Y
Takimoto
R
Takada
K
Miyanishi
K
Matsunaga
T
, et al. 
Resolution of liver cirrhosis using vitamin A-coupled liposomes to deliver siRNA against a collagen-specific chaperone
Nat. Biotechnol
2008
, vol. 
26
 (pg. 
431
-
442
)
171
Szoka
F
Molecular biology. The art of assembly
Science
2008
, vol. 
319
 (pg. 
578
-
579
)
172
Luten
J
van Nostrum
CF
De Smedt
SC
Hennink
WE
Biodegradable polymers as non-viral carriers for plasmid DNA delivery
J. Control Release
2008
, vol. 
126
 (pg. 
97
-
110
)
173
Satija
J
Gupta
U
Jain
NK
Pharmaceutical and biomedical potential of surface engineered dendrimers
Crit. Rev. Ther. Drug Carrier Syst
2007
, vol. 
24
 (pg. 
257
-
306
)
174
Petros
RA
Ropp
PA
DeSimone
JM
Reductively labile PRINT particles for the delivery of doxorubicin to HeLa cells
J. Am. Chem. Soc
2008
, vol. 
130
 (pg. 
5008
-
5009
)
175
Allen
TM
Martin
FJ
Advantages of liposomal delivery systems for anthracyclines
Semin. Oncol
2004
, vol. 
31
 (pg. 
5
-
15
)
176
Kirpotin
DB
Drummond
DC
Shao
Y
Shalaby
MR
Hong
K
Nielsen
UB
Marks
JD
Benz
CC
Park
JW
Antibody targeting of long-circulating lipidic nanoparticles does not increase tumor localization but does increase internalization in animal models
Cancer Res
2006
, vol. 
66
 (pg. 
6732
-
6740
)
This is an Open Access article distributed under the terms of the Creative Commons Attribution Non-Commercial License (http://creativecommons.org/licenses/by-nc/2.0/uk/) which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

Comments

0 Comments
Submit a comment
You have entered an invalid code
Thank you for submitting a comment on this article. Your comment will be reviewed and published at the journal's discretion. Please check for further notifications by email.