Tải bản đầy đủ (.pdf) (14 trang)

Báo cáo khoa học: Efficient and targeted delivery of siRNA in vivo pdf

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (4.96 MB, 14 trang )

MINIREVIEW
Efficient and targeted delivery of siRNA in vivo
Min Suk Shim
1
and Young Jik Kwon
1,2,3
1 Department of Chemical Engineering and Materials Science, University of California, Irvine, CA, USA
2 Department of Pharmaceutical Sciences, University of California, Irvine, CA, USA
3 Department of Biomedical Engineering, University of California, Irvine, CA, USA
Introduction
RNA interference (RNAi) is a highly conserved biologi-
cal process among yeasts, worms, insects, plants and
humans [1]. A single strand of exogenously introduced
double-stranded small interfering RNA (siRNA; 20–30
nucleotides) guides an RNA-inducing silencing protein
complex to degrade the mRNA with the matching
sequence; thus, translation into the target proteins is
silenced [2–4]. RNAi has been of great interest not only
as a powerful research tool to suppress the expression of
a target gene, but also as an emerging therapeutic strat-
egy to silence disease genes [5]. Theoretically, siRNA
can interfere with the translation of almost any mRNA,
as long as the mRNA has a distinctive sequence,
whereas the targets of traditional drugs are limited by
types of cellular receptors and enzymes [6].
Cancer, viral infections, autoimmune diseases and
neurodegenerative diseases have been explored as
promising disease targets of RNAi [7,8]. Recent pro-
gress in clinical trials using siRNA to cure age-related
macular degeneration (bevasiranib; Opko Health, Inc.,
Miami, FL, USA; phase III) and respiratory syncytial


virus infection (ALN-RSV01; Alnylam, Cambridge,
MA, USA; phase II) have demonstrated the therapeu-
tic potential of RNAi [9]. Moreover, the first evidence
Keywords
administration routes; barriers in siRNA
delivery; chemically modified RNA; in vivo
disease models; nanoparticles; nonviral
carriers; nucleic acid therapeutics; RNA
interference; targeted delivery in vivo;
viral vectors
Correspondence
Y. J. Kwon, Department of Pharmaceutical
Sciences, 916 Engineering Tower,
University of California, Irvine, CA 92697,
USA
Fax: +1 949 824 2541
Tel: +1 949 824 8714
E-mail:
(Received 7 July 2010, accepted
26 August 2010)
doi:10.1111/j.1742-4658.2010.07904.x
RNA interference (RNAi) has been regarded as a revolutionary tool for
manipulating target biological processes as well as an emerging and prom-
ising therapeutic strategy. In contrast to the tangible and obvious effective-
ness of RNAi in vitro, silencing target gene expression in vivo using small
interfering RNA (siRNA) has been a very challenging task due to
multiscale barriers, including rapid excretion, low stability in blood serum,
nonspecific accumulation in tissues, poor cellular uptake and inefficient
intracellular release. This minireview introduces major challenges in achiev-
ing efficient siRNA delivery in vivo and discusses recent advances in over-

coming them using chemically modified siRNA, viral siRNA vectors and
nonviral siRNA carriers. Enhanced specificity and efficiency of RNAi
in vivo via selective accumulations in desired tissues, specific binding to
target cells and facilitated intracellular trafficking are also commonly
attempted utilizing targeting moieties, cell-penetrating peptides, fusogenic
peptides and stimuli-responsive polymers. Overall, the crucial roles of the
interdisciplinary approaches to optimizing RNAi in vivo, by efficiently and
specifically delivering siRNA to target tissues and cells, are highlighted.
Abbreviations
ApoB, apolipoprotein B; CPP, cell-penetrating peptide; FA, folic acid; GFP, green fluorescent protein; HER-2, human epidermal growth
factor 2; i.p., intraperitoneal; i.t., intratumoral; i.v., intravenous; 9R, nonamer arginine residues; RGD, Arg-Gly-Asp peptide; RNAi,
RNA interference; siRNA, small interfering RNA; VEGF, vascular endothelial growth factor.
4814 FEBS Journal 277 (2010) 4814–4827 ª 2010 The Authors Journal compilation ª 2010 FEBS
of targeted in vivo gene silencing for human cancer
therapy via systemic delivery of siRNA using transfer-
rin-tagged, cyclodextrin-based polymeric nanoparticles
(CALAA-01; Calando Pharmaceuticals, Pasadena, CA,
USA; phase I) has been recently announced [10].
Despite quite efficient and reliable gene silencing
in vitro, only limited RNAi has been achieved in vivo
because of rapid enzymatic degradation in combina-
tion with poor cellular uptake of siRNA [11]. There-
fore, novel delivery systems, which enable prolonged
circulation of siRNA with resistance against enzymatic
degradation, high accessibility to target cells via clini-
cally feasible administration routes and optimized
cytosolic release of siRNA after efficient cellular
uptake, are indispensably required [12]. In this mini-
review, major factors in determining overall RNAi
efficiency in vivo are introduced. Moreover, up-to-date

progress in achieving efficient and targeted siRNA
delivery in vivo, particularly by overcoming multiscale
hurdles using novel siRNA carriers, is discussed.
Challenges in RNAi in vivo
Design and in vivo delivery of siRNA
There are multiple key considerations in order to
achieve efficient RNAi in vivo by delivering exogenous
siRNA. siRNA has to be designed to target hybridiza-
tion-accessible regions within the target mRNA while
avoiding unintended (off-target) effects [13–15], which
is extensively reviewed in this series by Walton et al.
[16]. In addition, siRNA can also induce adverse
effects such as immune responses, as discussed by
Samuel-Abraham & Leonard [17]. siRNA may induce
interferon responses either through the double-
stranded RNA-activated protein kinase PKR [18] or
toll-like receptor 3 [19]. Therefore, a combination of
computer algorithms and experimental validation
should be employed to determine the optimized siRNA
sequences that are complementary to target mRNA
while inducing minimal immune responses [20].
Naked siRNA is relatively unstable in blood in its
native form and is rapidly cleared from the body (i.e.
short half-lives in vivo) via degradion by ribonucleases,
rapid renal excretion and nonspecific uptake by the
reticuloendothelial system [21]. The phosphorothioate
backbone, or various 2¢ positions in the sugar moiety
of siRNA, is conventionally modified to enhance its
stability and activity against nuclease degradation
[22,23], without affecting gene silencing activity [24].

siRNA is an anionic macromolecule and does not
readily enter cells by passive diffusion mechanisms. An
appropriate siRNA delivery system enhances cellular
uptake, protects its payload from enzymatic digestion
and immune recognition, and improves the pharmacoki-
netics by avoiding excretion via the reticuloendothelial
system and renal filtration (i.e. prolonged half-life
in vivo) [25–27]. In addition, targeted delivery systems
localize siRNA in the desired tissue, resulting in a
reduction in the amount of siRNA required for
efficient gene silencing in vivo , as well as minimized
side effects. Therefore, the development of effective
in vivo delivery systems is pivotal in overcoming the
challenges in achieving efficient and targeted siRNA
delivery in vivo. Major hurdles in siRNA delivery
in vivo and various approaches to overcoming them
are illustrated in Fig. 1.
Local versus systemic delivery
The types of target tissues and cells dictate the
optimum administration routes of local versus systemic
delivery. For example, siRNA can be directly applied to
the eye, skin or muscle via local delivery, whereas sys-
temic siRNA delivery is the only way to reach meta-
static and hematological cancer cells. Local delivery
offers several advantages over systemic delivery, such as
low effective doses, simple formulation (e.g. no
targeting moieties), low risk of inducing systemic
side effects and facilitated site-specific delivery [28].
Therefore, if applicable, local delivery is likely to be a
more cost-efficient strategy for siRNA delivery in vivo

than systemic administration. For example, initial clini-
cal trials for RNAi-based treatment of age-related mac-
ular degeneration have exclusively used local injections
of siRNA directly into the eye [10]. Other promising
local routes include intranasal siRNA administration
for pulmonary delivery [10,29–31] and direct injection
into the central nervous system [10,32,33].
Alternatively, systemic delivery via intravenous
(i.v.), intraperitoneal (i.p.) or oral administration is
widely applicable when the target sites are not locally
confined or not readily accessible. Metastatic tumors
are especially amenable for systemic delivery compared
with local administration. For example, human bcl-2
oncogene-targeting siRNA, which was complexed with
cationic liposomes and i.v. injected, effectively inhib-
ited tumor growth in a mouse liver metastasis model
[34]. Another study showed that siRNA encapsulated
in a lipid vesicle was able to impart efficient and per-
sistent antiviral activity after being injected into a
hepatitis B virus mouse model [35]. However, impor-
tantly, systemic siRNA delivery imposes several
additional barriers in comparison with local delivery.
siRNA should remain in its active form during circula-
tion and be able to reach target tissues after passing
M. S. Shim and Y. J. Kwon In vivo siRNA delivery
FEBS Journal 277 (2010) 4814–4827 ª 2010 The Authors Journal compilation ª 2010 FEBS 4815
through multiple barrier organs (e.g. liver, kidney and
lymphoid organs).
Extracellular and intracellular barriers in siRNA
delivery in vivo

Regardless of administration routes, the final desti-
nation of siRNA is the cytoplasm of the target cell,
where it incorporates into RNA-inducing silencing
protein complex and encounters target mRNAs. First,
siRNA that survives in the plasma and is transported
close to a target tissue must extravasate through the
tight vascular endothelial junctions. It has been
reported that microvascular transport of macromoel-
cules > 5 nm in diameter is significantly inhibited in
normal tissues [36]. However, transport of macromole-
cules across the tumor endothelium is more efficient
than that of normal endothelium because of its leaky
and discontinuous vascular structures with poor lym-
phatic drainage. Thus, tumor endothelium allows the
penetration of high molecular mass macromolecules
(> 40 kDa), which is also referred to as ‘enhanced
permeation and retention effect’ [37]. siRNA, in its
native form or formulated in a delivery carrier, must
then diffuse through the extracellular matrix, a dense
network of fibrous protein and carbohydrates
surrounding a cell [38], before accessing target cells.
siRNA or its complex adheres preferably to target cells
via receptor-mediated specific binding, followed by
cellular uptake. Even after it is internalized by a cell,
siRNA should be released from the endosome, while
avoiding entrapment and degradation [39,40]. Because
the condition in the endosome ⁄ lysosome is mildly
acidic, facilitated cytosolic release of siRNA using
acid-responsive delivery carriers has been a popular
strategy to overcome this intracellular hurdle [41,42].

Fusogenic peptides which undergo acid-triggered con-
formational changes have also been shown to acceler-
Fig. 1. Interdisciplinary approaches to achieving efficient and targeted RNAi in vivo by overcoming multiscale barriers in systemic siRNA
delivery. Detailed design parameters of an ideal siRNA carrier are depicted in Fig. 2.
In vivo siRNA delivery M. S. Shim and Y. J. Kwon
4816 FEBS Journal 277 (2010) 4814–4827 ª 2010 The Authors Journal compilation ª 2010 FEBS
ate endosomal escape of nucleic acids [43,44]. Finally,
siRNA delivered by a carrier should be decomplexed
in the cytoplasm [45]. A broad range of novel materials
that provide enhanced siRNA release have been devel-
oped (e.g. disulfide-based cationic polymers) [46].
Fig. 1 shows extracellular and intracellular barriers in
siRNA delivery with various approaches to overcom-
ing them.
Chemically modified siRNA for
enhanced RNAi in vivo
Various molecular positions in siRNA have been chem-
ically replaced or modified, mainly to resist enzymatic
hydrolysis. For example, phosphodiester (PO
4
) linkages
were replaced with phosphothioate (PS) at the 3¢-end,
and introducing O-methyl (2¢-O-Me), fluoro (2¢-F)
group or methoxyethyl (2¢-O-MOE) group greatly pro-
longed half-lives in plasma and enhanced RNAi effi-
ciency in cultured cells [47–51]. In addition, efficiency
enhancer molecules were conjugated to either the 5¢-or
3¢-end of the sense strand, without affecting the activity
of the antisense strand [52]. There are some potential
risks that chemically modifying siRNA may compro-

mise RNAi efficiency. For example, boranophospho-
nate modification at the central position of the
antisense strand of siRNA showed improved resistance
to nuclease degradation, but simultaneously reduced
RNAi activity [53]. In addition, non-natural molecules
produced upon the degradation of a chemically modi-
fied siRNA may generate metabolites that might be
unsafe or trigger unwanted effects. To date, cholesterol
and aptamers are the most promising siRNA conjugates
that have demonstrated efficient RNAi in vivo.
Cholesterol–siRNA conjugates
Improved pharmacokinetic and cellular uptake proper-
ties of cholesterol–siRNA conjugates silenced apolipo-
protein B (ApoB) in mice via i.v. administration [22].
By contrast, ApoB siRNA unconjugated with choles-
terol was unable to induce mRNA interference and
was rapidly cleared. The mechanisms of improved dis-
tribution and cellular uptake of siRNA through cho-
lesterol conjugation were demonstrated in a recent
study; cholesterol–siRNA conjugates seem to incorpo-
rate into circulating lipoprotein particles (i.e. improved
distribution in vivo) and are efficiently internalized by
hepatocytes via receptor-mediated processes (i.e. effi-
cient cellular uptake) [54]. Prebinding of cholesterol–
siRNA conjugates to lipoparticles dramatically
improved silencing efficacy in mice, and lipoparticle
types affected cholesterol–siRNA conjugate distribu-
tion in various tissues [54]. Using a transgenic mouse
model for Huntington’s disease, it was also demon-
strated that a single intrastriatal injection of choles-

terol–siRNA conjugates silenced a mutant Huntingtin
gene, attenuating neuronal pathology as well as
delaying the abnormal behavioral phenotype [55].
RNA aptamer–siRNA conjugates
RNA aptamers have been popularly used to selectively
deliver siRNA in vivo to target tissues and cells, such
as prostate cancer cells and tumor vascular endothe-
lium overexpressing prostate-specific membrane anti-
gen [56]. A key advantage of aptamer-mediated
targeted delivery systems is that RNA aptamers can be
facilely obtained by in vitro transcription reaction and,
therefore, avoid contamination by cell or bacterial
products. Promising in vitro and in vivo RNAi was
obtained using siRNA that was directly linked with
prostate-specific membrane antigen aptamers [57]. An
aptamer-based delivery system has also been used to
suppress HIV infection. Anti-gp120 RNA aptamers
were covalently conjugated with a strand of siRNA,
and the other siRNA strand was subsequently
annealed to the aptamer-conjugated strand. These
aptamer–siRNA conjugates were able to access HIV-
infected cells and silence viral replication in vitro [58].
Viral vectors: natural siRNA carriers
Various recombinant viral vectors have been shown to
be efficient in obtaining gene silencing for an extended
period in a wide range of mammalian cells [59]. For
example, an adenoviral vector encoding siRNA against
pituitary tumor transforming gene 1 significantly inhib-
ited the growth of the pituitary tumor transforming
gene 1-overexpressing hepatocellular carcinoma cells

in vitro and in vivo [60]. It was also demonstrated that
the herpes simplex virus type 1-based amplicon vectors
suppressed in vivo tumorigenicity of human polyomavi-
rus BK-transformed cells (pRPc cells) [61]. Recombi-
nant lentiviral vectors have also been frequently used to
achieve in vivo gene silencing. In particular, lentiviral
vectors containing the U6 promoter were found to be
efficient in green fluorescent protein (GFP) silencing
in vitro, resulting in  80% gene silencing at an average
of one integrated vector genome per target cell genome.
In addition, the U6 promoter was shown to be superior
to the H1 promoter in achieving in vivo gene silencing
and led to persistent GFP knockdown in the mouse
brain for at least 9 months [62]. This indicates that
lentivirus-mediated RNAi is a promising strategy for
long-term gene silencing in vitro and in vivo. Other viral
M. S. Shim and Y. J. Kwon In vivo siRNA delivery
FEBS Journal 277 (2010) 4814–4827 ª 2010 The Authors Journal compilation ª 2010 FEBS 4817
siRNA carriers such as retroviral vectors have not been
intensively explored for their use in vivo [63–65].
Although viral vectors provide excellent tissue-specific
tropism and high RNAi efficiency, safety concerns (e.g.
insertion mutagenesis and immunogenicity) and difficul-
ties with large-scale manufacture may limit the use of
viral vectors for siRNA delivery in clinical setting
[66,67]. Therefore, synthetic counterparts (nonviral
vectors) have been more and more intensively explored
as safe and effective alternatives that are easy to be
prepared and can deliver large payloads of siRNA.
Nonviral carriers: Trojan horses for

efficient, biocompatible and versatile
siRNA delivery in vivo
Delivery of siRNA in its unmodified form has several
advantages over using a chemically modified form.
Unmodified siRNA possesses untouched RNAi
capability (maximized RNAi per siRNA molecule) and
does not require potentially inefficient and time ⁄
labor-consuming modification processes (cost-effective
preparation). However, its highly anionic nature and
the macromolecular size of siRNA necessitates using
efficient carriers to overcome multiscale barriers.
Unlike viral vectors, which deliver siRNA in the form
of a viral genome, nonviral carriers deliver native
siRNA, generate low immunogenicity and offer high
structural and functional tunability. An ideally designed
nonviral siRNA carrier with its desirable structural
and functional multicomponents is depicted in Fig. 2.
Liposomes and lipoplexes
One of the most significant advances in RNAi in vivo
is successful knockdown of ApoB in nonhuman
primates by systemically delivered siRNA in stable
nucleic acid–lipid particles [68]. The siRNA–lipid
complexes showed significantly enhanced cellular inter-
nalization and endosomal escape of siRNA. ApoB
siRNA-carrying stable nucleic acid–lipid particles
greatly reduced ApoB expression and serum choles-
terol levels in monkeys when a clinically acceptable
single siRNA dose of 2.5 mgÆkg
)1
was injected i.v.

[68]. Importantly, expression of ApoB was silenced for
at least 11 days. With addressing the high toxicity
of the currently available liposomes for siRNA deliv-
ery in vitro and in vivo [69,70], cationic cardiolipin
analog-based liposomes carrying c-raf siRNA inhibited
the growth of breast tumor xenografts in mice [71].
Cationic liposomes formulated with anisamide-
conjugated poly(ethylene glycol) effectively penetrated
the lung metastasis of melanoma tumors in mice and
resulted in 70–80% gene silencing after a single i.v.
injection [72].
Further noticeable progress in siRNA delivery using
liposomes is the use of neutral lipids for systemic
siRNA delivery in order to address the toxicity of cat-
ionic lipids. For example, cyclin D1 (CyD1) siRNA
was efficiently encapsulated in neutral phospholipid-
based liposomes coated with hyaluronan [73]. The
resulting siRNA-carrying liposomes were stable during
circulation in vivo after i.v injection and suppressed
leukocyte proliferation and cytokine secretion by
type 1 T-helper cells. Another neutral dioleoyl phos-
phatidylcholine-based delivery system, which targets
EphA2 [74] and focal adhesion kinase [75], demon-
strated significantly inhibited tumor growth in an
orthotropic ovarian cancer model in mice. The same
type of liposome has also been reported to efficiently
silence neuropilin-2 expression and inhibit the growth
of colorectal cancer xenografts in the mouse liver [76].
Polymers and peptides
Nucleic acids such as siRNA are easily complexed

with synthetic cationic polymers e.g., polyethylenimine
Fig. 2. An ideally designed nonviral siRNA carrier for efficient and
targeted RNAi in vivo.
In vivo siRNA delivery M. S. Shim and Y. J. Kwon
4818 FEBS Journal 277 (2010) 4814–4827 ª 2010 The Authors Journal compilation ª 2010 FEBS
(PEI), biodegradable cationic polysaccharide (e.g.
chitosan) and cationic polypeptides [e.g. atelocollagen,
poly(l-lysine) and protamine], via attractive electro-
static interactions. For example, i.t. injection of siR-
NA– atelocollagen complexes silenced luciferase
expression in germ cell tumor xenografted in mice and
inhibited tumor growth [77]. In another study, vascular
endothelial growth factor (VEGF) siRNA–atelocolla-
gen complexes significantly suppressed tumor angio-
genesis and growth in a prostate tumor model in mice
[78]. Intravenous administration of chitosan–RhoA
siRNA complexes resulted in effective gene silencing in
subcutaneously implanted breast cancer cells in mice
[79]. In addition, intranasally administered chitosan–
siRNA complexes efficiently silenced GFP expression
in bronchiole epithelial cells in GFP-transgenic mice
[29]. Tumor necrosis factor expression in systemic mac-
rophages was silenced in mice after i.p. administration
of chitosan ⁄ siRNA complexes, thus downregulating
systemic and local inflammation [80].
Polyethylenimine is one of the most popularly inves-
tigated synthetic cationic polymers for nucleic acid
delivery in vitro and in vivo. Polyethylenimine is very
potent in transfection with its uniquely high buffering
capability at an endosomal pH (proton sponge effect)

which releases nucleic acid payloads into the cytoplasm
[39]. c-erbB2 ⁄ neu (HER-2) siRNA was delivered to
subcutaneous tumors via i.p. administration of siRNA ⁄
polyethylenimine complexes and resulted in a remark-
able reduction of tumor growth [81]. Pain receptors for
N-methyl-d-aspartate were effectively knocked-down by
intrathecal delivery of polyethylenimine-conjugated
siRNA in rats [82]. Inhibited viral propagation in the
lungs was also observed after deacetylated linear
polyethylenimine ⁄ siRNA complexes targeting influenza
nucleoprotein was retro-orbitally administered [83]. In
another study, polyethylenimine-conjugated siRNA
against secreted growth factor pleiotrophin reduced
tumor growth and cell proliferation with no toxicity or
abnormal animal behaviors after intracerebral adminis-
tration in an orthotopic glioblastoma mouse model [84].
Overall, polyethylenimine seems to be a promising
nonviral carrier for siRNA delivery in vivo, if its high
toxicity and limited biodegradability are appropriately
addressed.
Polypeptides, such as poly(l-lysine) and protamine,
have also commonly been used to deliver siRNA.
A sixth generation of dendritic poly(l-lysine) was
employed to systemically deliver siRNA to silence
ApoB expression without hepatotoxicity in hepatocytes
of apolipoprotein E-deficient mice [85]. Protamine, a
natural arginine-rich cationic polypeptide, condenses
negatively charged nucleic acids and has been used as
an efficient gene-delivery carrier [86]. An in vivo study
demonstrated that complexes of siRNA and low

molecular mass protamine, which possess membrane-
translocating potency, were accumulated in tumors via
i.p. administration and successfully inhibited the
expression of VEGF, thereby suppressing the growth
of hepatocarcinoma tumors in mice [87]. In addition,
no noticeable increase in inflammatory cytokines,
including interferon-a and interleukin-12, in serum was
observed when the low molecular mass protamine ⁄
siRNA complexes were administered, indicating
negligible immunostimulatory effects.
One of the fundamental concerns in using synthetic
polymers for siRNA delivery in vivo is dose-dependent
toxicity upon systemic administration. For example,
polyethylenimine and poly(l-lysine) have been shown
to trigger necrosis and apoptosis in a variety of cell lines
[88,89]. The toxicity can be ameliorated by conjugation
with biocompatible, hydrophilic polymers such as
poly(ethylene glycol) or by removing excess (i.e., uncom-
plexed) cationic polymers. In gneral, natural cationic
polymers (e.g. chitosan and protamine), which are bio-
compatible, biodegradable and nontoxic, are more desir-
able in siRNA delivery in vivo than synthetic polymers.
Targeted siRNA delivery in vivo
In order to achieve RNAi in vivo via systemic delivery,
it is crucial for siRNA to be efficiently located in
desired tissues ⁄ cells. This requires three important pro-
cesses: prolong circulation in the body, high accessibil-
ity to target tissues and specific binding to target cells.
Targeted siRNA delivery maximizes the local concen-
tration in the desired tissue (maximized and localized

silencing effects) and prevents nonspecific siRNA dis-
tribution (minimized unwanted effects in non-target
tissues). For example, recent studies have reported can-
cer-targeted siRNA delivery using nanoparticles that
specifically bind to cancer-specific or cancer-associated
antigens and receptors [90,91].
Folate-conjugated siRNA carriers
One of the most popular target molecules in cancer-
specific gene and drug delivery is the folate receptor
[92]. Folic acid (FA) is needed for rapid cell growth,
and many cancer cells overexpress folate receptors to
which FA and monoclonal antibodies specifically bind
[93]. FA can be easily conjugated onto the surface of
liposomal and polymeric siRNA carriers with or with-
out a poly(ethylene glycol) spacer [92]. For example,
FA-conjugated polyethylenimine showed enhanced
gene silencing via receptor-mediated endocytosis [94].
M. S. Shim and Y. J. Kwon In vivo siRNA delivery
FEBS Journal 277 (2010) 4814–4827 ª 2010 The Authors Journal compilation ª 2010 FEBS 4819
Chimeric survivin siRNA incorporated with bacterio-
phage phi29-encoded RNA and when further
conjugated with FA suppressed the growth of naso-
pharyngeal carcinoma in mice, whereas control
FA-free siRNA–phi29-encoded RNA hybrid did not
affect tumor development [95]. As described earlier,
RNA aptamer-mediated targeted siRNA delivery by
direct conjugation with siRNA or tethering onto
carriers has been a frequently adapted strategy.
Arg–Gly–Asp peptide-conjugated siRNA carriers
Arg–Gly–Asp (RGD) peptide targets tumor vascula-

ture expressing a
v
b
3
integrin. Poly(ethylene glycol)
ylated poly(ethylenimine) conjugated with RGD
peptides was developed to selectively deliver VEGF
siRNA to tumors [96]. In this study, i.v. injected poly-
ethylenimine-poly(ethylene glycol)-RGD ⁄ siRNA com-
plexes inhibited tumor angiogenesis and the growth of
integrin-expressing murine neuroblastoma tumors in
mice [96]. Systemic delivery (i.v. injection) and local
delivery of poly(ethylene glycol)-polyethylenimine-RGD
complexing VEGF siRNA also showed a significant
inhibitory effect on virus-induced angiogenesis as well
as the development of herpetic stromal keratitis lesions
[97].
Antibody-conjugated siRNA carriers
Many studies have suggested that antibodies are good
targeting modalities for targeted siRNA delivery
in vivo, when careful selection of target antigen is
made. Ideal antigens should be exclusively expressed
or substantially overexpressed on target cells. Exam-
ples of antigens that have been used for cancer-tar-
geted drug and gene delivery include HER-2 [98] and
epidermal growth factor receptor [99]. For example,
HER-2 siRNA-carrying liposomes decorated with
transferrin receptor-specific antibody fragments (i.e.
nanoimmunoliposome) silenced the HER-2 gene in
xenograft tumors in mice, significantly inhibiting

tumor growth [100]. An antibody fragment against an
HIV gp160 has also been used for targeted siRNA
delivery in vivo. siRNA linked to a protamine–anti-
body fusion protein, called F105-P, showed inhibited
HIV replication in infected primary T cells [101].
Moreover, i.t. or i.v. injection of F105-P ⁄ siRNA com-
plexes into mice successfully targeted gp160-expressing
B16 melanoma cells. A synthetic chimeric peptide,
which consists of nonamer arginine residues (9R)
added to the C-terminus of a rabies virus glycoprotein
peptide (29 amino acids) (RVG-9R), was able to spe-
cifically deliver siRNA to acetylcholine receptor-
expressing neuronal cells after i.v. administration [102].
In addition, treating mice with Japanese encephalitis
virus siRNA complexed with RVG-9R showed robust
protection of the animals from lethal infection.
Intracellular siRNA delivery
In many aspects, siRNA delivery is similar to that of
delivering other types of nucleic acids such as plasmid
DNA, because they share most extracellular and intra-
cellular barriers. However, several unique challenges in
siRNA delivery make achieving efficient RNAi difficult
compared with plasmid DNA delivery. First, the final
target destination of siRNA is the cytoplasm, whereas
plasmid DNA must be transported into the nucleus.
This implies that siRNA should be rapidly released
from its carrier upon endosomal escape. Second, overall
RNAi efficiency is proportional to the number of
siRNAs complexed with RNA-inducing silencing
protein complex, whereas a successfully delivered single

copy of plasmid DNA might be sufficient to express new
transgene proteins. In other words, the maximum possi-
ble number of siRNA needs to be delivered in the cyto-
plasm in order to achieve the desired biological effects.
Third, siRNA acts only once, whereas plasmid DNA
can be replicated or even can be incorporated into the
host chromosome [103] (short vs. permanent effects).
Cell-penetrating peptide-mediated siRNA delivery
Cell-penetrating peptides (CPPs), short cationic poly-
peptides with a maximum of 30 amino acids, have been
extensively used to obtain enhanced intracellular deliv-
ery of a wide range of macromolecules [104]. CPPs have
been shown to bind the anionic cell surface through elec-
trostatic interactions and rapidly induce cellular inter-
nalization through relatively unclear mechanisms,
although recent evidence shows that CPP-mediated
internalization might be an endocytosis-mediated pro-
cess [105,106]. Various CPPs, including TAT and MPG
proteins from HIV-1 [107–110], as well as penetratin and
polyarginine [111,112], have been employed for intracel-
lular delivery of various proteins and nucleic acids.
Oligoarginine (e.g. 9 arginine, 9R), the simplest
CPP, conjugated with cholesterol was shown to effi-
ciently deliver siRNA to a transplanted tumor in mice
[113]. It was also reported that HER-2 siRNA com-
plexed with short arginine peptide was localized in
perinuclear regions of the cytoplasm in vitro, further
significantly inhibiting tumor growth of ovarian cancer
xenografts [114]. Polyamidoamine dendrimer-TAT
conjugated with bacterial magnetic nanoparticles was

also used to deliver epidermal growth factor receptor
In vivo siRNA delivery M. S. Shim and Y. J. Kwon
4820 FEBS Journal 277 (2010) 4814–4827 ª 2010 The Authors Journal compilation ª 2010 FEBS
siRNA to human glioblastoma cells in vitro as well as
xenografts [115]. Another type of CPP, MPG-8, was
also used to complex cyclin B1 siRNA, and the result-
ing complexes were further decorated with cholesterol
for i.v. injection to the mice bearing human prostate
carcinoma and human lung cancer xenografts [116].
The results showed efficient siRNA delivery in vivo at
a low effective dose (0.5 mgÆkg
)1
), indicated by
inhibited tumor growth.
CPP-mediated cellular internalization via endocyto-
sis requires additional molecules for facilitated cyto-
solic release of siRNA. For example, it was found that
TAT–siRNA conjugates resulted in no gene silencing
because they were entrapped in the endosomes even
after efficiently entering cells [117]. Photostimulating
fluorescently labeled TAT efficiently released TAT–
siRNA conjugates from the endosome, resulting in
enhanced gene silencing efficiency. Chloroquine and
Table 1. siRNA delivery systems for RNAi in vivo. BCL-2, B-cell lymphoma 2; Cyb1, cyclin B1; CyD1, cyclin D1; DOPC, 1,2-dioleoyl-sn-glyce-
ro-3-phosphatidylcholine; DOPE, dioleoyl phosphatidylethanolamine; DOTAP, (N-[1-(2,3-dioleoyloxy)]-N-N-N-trimethyl ammonium propane);
DPPE, dipalmitoyl phosphatidylethanolamine; DSPE, distearoyl phosphatidylethanolamine; FAK, focal adhesion kinase; HST-1 ⁄ FGF-4,
fibroblast growth factor; i.c.v., intracerebroventricular; i.p., intraperitoneal; i.v., intravenous; MMP-2, matrix metalloproteinase-2; NMDA,
N-methyl-
D-aspartate; NR2B, NMDA-R2B receptor subunit protein receptors; PAMAM, polyamidoamine dendrimer; PLK-1, polo-like kinase 1;
PTTG1, pituitary tumor transforming gene 1; RVG, rabies virus glycoprotein; SNALP, stable nucleic acid-lipid particles; TNF-a, tumor necrosis

factor-a; VEGF, vascular endothelial growth factor.
Delivery system Target gene In vivo model
a
Delivery route Ref.
Cholesterol–siRNA ApoB ApoB transgenic mice i.v. 22
RNA aptamer–siRNA PLK-1, BCL-2 Prostate tumor xenograft i.t. 57
Adenoviral vector PTTG1 Hepatoma tumor xenograft i.t. 60
Lentiviral vector GFP GFP transgenic brain Stereotactic 62
Stable nucleic acid lipid particles (SNALP) ApoB Monkeys i.v. 68
Cardiolipin analog-based liposome c-Raf Breast tumor xenograft i.v. 71
DSPE–poly(ethylene glycol)
–DOTAP–cholesterol liposome
Luciferase B16F10 melanoma tumors i.v. 72
Hyaluronan–DPPE liposome CyD1 Gut inflammation i.v. 73
Neutral DOPC liposome EphA2 Ovarian cancer i.v. 74
Neutral DOPC liposome FAK Ovarian cancer i.p. 75
Neutral DOPC liposome Neuropilin-2 Colorectal tumor xenograft i.p. 76
Atelocollagen HST-1 ⁄ FGF-4 Luciferase Germ cell xenograft i.t. 77
Atelocollagen VEGF Prostate tumors xenograft i.t. 78
Chitosan EGFP Transgenic EGFP mice Intranasal 29
Chitosan RhoA Breast tumors xenograft i.v. 79
Chitosan TNF-a Mice i.p. 80
Polyethylenimine HER-2 Ovarian tumor xenograft i.p. 81
Polyethylenimine NR2B Nociception in rats Intrathecal 82
Polyethylenimine Influenza nucleoprotein Influenza virus infected-lung Retro-orbital 83
Polyethylenimine Pleiotrophin (PTN) Glioblastoma xenograft Intracerebral, i.p. 84
Poly(
L-lysine) ApoB Mice i.v. 85
Protamine VEGF Hepatocarcinoma xenograft i.p. 87
RGD–poly(ethylene glycol)–poly(ethylenimine) VEGF Neuroblastoma xenograft i.v. 96

RGD–poly(ethylene glycol)–poly(ethylenimine) VEGF Corneal neovascularization Subconjunctival, i.v. 97
HER-2-liposomes with histidine–lysine peptide HER-2 Pancreatic tumor xenograft i.v. 100
HIV antibody–protamine c-myc, MDM2, VEGF B16 melanoma cells expressing i.t. 101
HIV envelop i.v.
Arginine RVG Neuronal cells i.v. 102
Oligoarginine (9R) conjugated-water-soluble
lipopolymer (WSLP)
VEGF Colon adenocarcinoma xenograft i.t. 113
Oligoalginine (15R) HER-2 Ovarian tumor xenograft i.t. 114
TAT-PAMAM EGF receptor Glioblastoma xenograft i.t. 115
Cholesterol-MPG-8 CyB1 Prostate tumor xenograft i.t. 116
Lung tumor xenograft i.v.
DOPE-Cationic Lipid Luciferase Mouse brain i.c.v. 123
GALA peptide–poly(ethylene glycol)
–MMP-2 cleavable peptide-DOPE
Luciferase Fibrosarcoma xenograft i.t. 128
a
All the listed in vivo models involved a mouse model except Zimmermann et al. [68] and Tan et al. [82].
M. S. Shim and Y. J. Kwon In vivo siRNA delivery
FEBS Journal 277 (2010) 4814–4827 ª 2010 The Authors Journal compilation ª 2010 FEBS 4821
influenza virus-derived hemagglutinin peptide have also
been frequently used to destabilize the endosomal
membrane and enhance the cytosolic release of CPP-
conjugated macromolecules [118–120].
Fusogenic or pH-responsive intracellular delivery
of siRNA
Fusogenic peptides and lipids and pH-responsive lipo-
plexes and polyplexes have been used to ensure facili-
tated siRNA into the cytoplasm from the endosomes.
For example, the incorporation of polypeptides derived

from the endodomain of the HIV-1 envelope (HGP) or
influenza virus fusogenic peptide (diINF-7) signifi-
cantly promoted the liposomal fusion with the endoso-
mal membrane, enhancing siRNA escape into the
cytoplasm [40,121]. Similarly, equipping lipoplexes
with fusogenic lipids, such as dioleoyl phosphatidyl-
ethanolamine (DOPE), was shown to facilitate the
endosomal release of siRNA payload [122,123].
Stimuli-triggered macromolecule release from the
mildly acidic endosome (e.g. pH 5.0–6.0) has been
popularly investigated using a number of novel acid-
responsive polymers [124–126]. For example, poly(eth-
ylene glycol) shielding the surface of a highly fusogenic
phosphatidylethanolamine lipid vesicles was cleaved
upon acid hydrolysis of the vinyl ether bond, triggering
fusion with the endosomal membrane [127]. A matrix
metalloproteinase-cleavable and pH-sensitive GALA
peptide was also used to link poly(ethylene glycol) and
dioleoyl phosphatidylethanolamine (DOPE) lipid to
obtain enhanced siRNA delivery specifically into cancer
cells [128]. Highly efficient siRNA-mediated knockdown
of luciferase expression was achieved in human fibrosar-
coma cells in vitro and xenografted tumors using this
method. Acid-degradable ketalized linear polyethyl-
enimine significantly increased gene silencing efficiency
via efficient cytosolic release with high resistance to
serum and low cytotoxicity [129]. It was demonstrated
that ketalized linear polyethylenimine ⁄ siRNA poly-
plexes were efficiently released into the cytoplasm upon
acid-hydrolysis of ketal branches in the endosomes, fol-

lowed by enhanced siRNA disassembly from ketalized
linear polyethylenimine in the cytoplasm [129].
Conclusion
RNAi is an emerging therapeutic strategy and has
been widely investigated. Despite a few promising
clinical trials, effectively delivering siRNA in vivo
remains a pivotal challenge in translating RNAi in the
clinic as a conventional treatment option. A number of
delivery systems and strategies have been developed to
overcome multiscale extracellular and intracellular
barriers to siRNA delivery in vivo, as summarized in
Table 1. Chemically modified siRNA is stable against
enzymatic degradation but can be cleared easily, gener-
ating potentially hazardous metabolites. Viral siRNA
delivery raises several safety and preparation concerns
such as immune responses and limited large-scale pro-
duction. Nonviral siRNA carriers are efficient, safe
and versatile in tackling the barriers in siRNA circula-
tion, permeation into desired tissues, specific binding
to target cells and optimized intracellular trafficking.
Recent advances clearly indicate that interdisciplinary
approaches using biology, chemistry and engineering
play crucial roles in achieving efficient and targeted
siRNA delivery in vivo.
Acknowledgement
This work was supported by NSF CAREER Award
(DMR-0956091) and a Council on Research Comput-
ing and Libraries Research Grant (UC Irvine).
References
1 Scherr M, Morgan MA & Eder M (2003) Gene silenc-

ing mediated by small interfering RNAs in mammalian
cells. Curr Med Chem 10, 245–256.
2 Elbashir SM, Harborth J, Lendeckel W, Yalcin A,
Weber K & Tuschl T (2001) Duplexes of 21-nucleotide
RNAs mediate RNA interference in cultured mamma-
lian cells. Nature 411, 494–498.
3 Hannon GJ & Rossi JJ (2004) Unlocking the potential
of the human genome with RNA interference. Nature
431, 371–378.
4 Novina C & Sharp P (2004) The RNAi revolution.
Nature 430, 161–164.
5 Iorns E, Lord CJ, Turner N & Ashworth A (2007)
Utilizing RNA interference to enhance cancer drug
discovery. Nat Rev Drug Discov 6, 556–568.
6 de Fougerolles A, Vornlocher H-P, Maraganore J &
Lieberman J (2007) Interfering with disease: a progress
report on siRNA-based therapeutics. Nat Rev Drug
Discov 6, 443–453.
7 Castanotto D & Rossi JJ (2009) The promises and
pitfalls of RNA interference-based therapeutics. Nature
457, 426–433.
8 Manjunath N, Kumar P, Lee SK & Shankar P (2006)
Interfering antiviral immunity: application, subversion,
hope? Trends Immunol 27, 328–335.
9 Melnikova I (2007) RNA-based therapies. Nat Rev
Drug Discov 6, 863–864.
10 Oh Y-K & Park TG (2009) siRNA delivery systems
for cancer treatment. Adv Drug Deliver Rev 61, 850–
862.
In vivo siRNA delivery M. S. Shim and Y. J. Kwon

4822 FEBS Journal 277 (2010) 4814–4827 ª 2010 The Authors Journal compilation ª 2010 FEBS
11 Kirchhoff F (2008) Silencing HIV-1 in vivo. Cell 134,
566–568.
12 Grimm D & Kay MA (2007) Therapeutic application
of RNAi: is mRNA targeting finally ready for prime
time? J Clin Invest 117, 3633–3641.
13 Jackson AL, Bartz SR, Schelter J, Kobayashi SV,
Burchard J, Mao M, Li B, Cavet G & Linsley PS
(2003) Expression profiling reveals off-target gene
regulation by RNAi. Nat Biotechnol 21, 635–637.
14 Fedorov Y, Anderson EM, Birmingham A, Reynolds
A, Karpilow J, Robinson K, Leake D, Marshall WS &
Khvorova A (2006) Off-target effects by siRNA can
induce toxic phenotype. RNA 12, 1188–1196.
15 Birmingham A, Anderson EM, Reynolds A, Ilsley-
Tyree D, Leake D, Fedorov Y, Baskerville S,
Maksimova E, Robinson K, Karpilow J et al. (2006)
3¢ UTR seed matches, but not overall identity, are
associated with RNAi off-targets. Nat Methods 3,
199–204.
16 Walton SP, Wu M, Gredell JA & Chan C (2010)
Designing highly active siRNAs for therapeutic
applications. FEBS J 277, 4806–4813.
17 Samuel-Abraham S & Leonard JN (2010) Staying on
message: design principles for controlling nonspecific
responses to siRNA. FEBS J 277, 4828–4836.
18 Katze MG, Wambach M, Wong ML, Garfinkel M,
Meurs E, Chong K, Williams BRG, Hovanessian
AG & Barber GN (1991) Functional expression and
RNA binding analysis of the interferon-induced,

double-stranded RNA-activated, 68,000-M
r
protein
kinase in a cell-free system. Mol Cell Biol 11, 5497–
5505.
19 Alexopoulou L, Holt AC, Medzhitov R & Flavell RA
(2001) Recognition of double-stranded RNA and acti-
vation of NF-jB by Toll-like receptor 3. Nature 413,
732–738.
20 Amarzguioui M & Prydz H (2004) An algorithm for
selection of functional siRNA sequences. Biochem
Biophys Res Commun 316, 1050–1058.
21 Whitehead KA, Langer R & Anderson DG (2009)
Knocking down barriers: advances in siRNA delivery.
Nat Rev Drug Discov 8, 129–138.
22 Soutschek J, Akinc A, Bramlage B, Charisse K,
Constien R, Donoghue M, Elbashir S, Geick A,
Hadwiger P, Harborth J et al. (2004) Therapeutic silenc-
ing of an endogenous gene by systemic administration
of modified siRNAs. Nature 432, 173–178.
23 Morrissey D, Blanchard K, Shaw L, Jensen K,
Lockridge J, Dickinson B, McSwiggen J, Vargeese C,
Bowman K, Shaffer C et al. (2005) Activity of stabilized
short interfering RNA in a mouse model of hepatitis B
virus replication. Hepatology 41, 1349–1356.
24 Bumcrot D, Manoharan M, Koteliansky V & Sah DW
(2006) RNAi therapeutics: a potential new class of
pharmaceutical drugs. Nat Chem Biol 2, 711–719.
25 Sioud M & Sorensen D (2003) Cationic liposome-
mediated delivery of siRNAs in adult mice. Biochem

Biophys Res Commun 312, 1220–1225.
26 Sorensen D, Leirdal M & Sioud M (2003) Gene
silencing by systemic delivery of synthetic siRNAs in
adult mice. J Mol Biol 327, 761–766.
27 Lorenz C, Hadwiger P, John M, Vornlocher H-P &
Unverzagt C (2004) Steroid and lipid conjugates of
siRNAs to enhance cellular uptake and gene silencing
in liver cells. Bioorg Med Chem Lett 14, 4975–4977.
28 Dykxhoorn DM, Palliser D & Lieberman J (2006) The
silent treatment: siRNAs as small molecule drugs.
Gene Ther 13, 541–552.
29 Howard KA, Rahbek UL, Liu X, Damgaard CK,
Glud SZ, Andersen MØ, Hovgaard MB, Schmitz A,
Nyengaard JR, Besenbacher F et al. (2006) RNA
interference in vitro and in vivo using a chitosan ⁄
siRNA nanoparticle system. Mol Ther 14, 476–484.
30 Bitko V, Musiyenko A, Shulyayeva O & Barik S
(2005) Inhibition of respiratory viruses by nasally
administered siRNA. Nat Med 11, 50–55.
31 Zhang X, Shan P, Jiang D, Noble P, Abraham N,
Kappas A & Lee P (2004) Small interfering RNA
targeting hemeoxygenase-1 enhances ischemia-reper-
fusion-induced lung apoptosis. J Biol Chem 279, 10677–
10684.
32 Dorn G, Patel S, Wotherspoon G, Hemmings-Mieszc-
zak M, Barclay J, Natt F, Martin P, Bevan S, Fox A,
Ganju P et al. (2004) siRNA relieves chronic neuro-
pathic pain. Nucleic Acids Res 32, e4.
33 Querbes W, Ge P, Zhang W, Fan Y, Costigan J,
Charisse K, Maier M, Nechev L, Manoharan M,

Kotelianski V et al. (2009) Direct CNS delivery of
siRNA mediates robust silencing in oligodendrocytes.
Oligonucleotides 19, 23–29.
34 Yano J, Hirabayashi K, Nakagawa S-I, Yamaguchi T,
Nogawa M, Kashimori I, Naito H, Kitagawa H, Ishiy-
ama K, Ohgi T et al. (2004) Antitumor activity of small
interfering RNA ⁄ cationic liposome complex in mouse
models of cancer. Clin Cancer Res 10, 7721–7726.
35 Morrissey DV, Lockridge JA, Shaw L, Blanchard K,
Jensen K, Breen W, Hartsough K, Machemer L,
Radka S, Jadhav V et al. (2005) Potent and persistent
in vivo anti-HBV activity of chemically modified
siRNAs. Nat Biotechnol 23, 1002–1007.
36 Juliano R, Bauman J, Kang H & Ming X (2009)
Biological barriers to therapy with antisense and
siRNA oligonucleotides. Mol Pharm 6, 686–695.
37 Jang SH, Wientjes MG, Lu D & Au JL-S (2003) Drug
delivery and transport to solid tumors. Pharm Res 20 ,
1337–1350.
38 Za
´
mec
ˇ

´
k J, Vargova
´
L, Homola A, Kodet R &
Sykova
´

E (2003) Extracellular matrix glycoproteins
and diffusion barriers in human astrocytic tumours.
Neuropathol Appl Neurobiol 30, 338–350.
M. S. Shim and Y. J. Kwon In vivo siRNA delivery
FEBS Journal 277 (2010) 4814–4827 ª 2010 The Authors Journal compilation ª 2010 FEBS 4823
39 Boussif O, Lezoualc’h F, Zanta MA, Mergny MD,
Scherman D, Demeneix B & Behr J-P (1995) A versa-
tile vector for gene and oligonucleotide transfer into
cells in culture and in vivo: polyethylenimine. Proc Natl
Acad Sci USA 92, 7297–7301.
40 Oliveira S, van Rooy I, Kranenburg O, Storm G &
Schiffelers RM (2007) Fusogenic peptides enhance
endosomal escape improving siRNA-induced silencing
of oncogenes. Int J Pharm 331, 211–214.
41 Medina-Kauwe LK, Xie J & Hamm-Alvarez S (2005)
Intracellular trafficking of nonviral vectors. Gene Ther
12, 1734–1751.
42 Cho YW, Kim J-D & Park K (2003) Polycation gene
delivery systems: escape from endosomes to cytosol.
J Pharm Pharmacol 55, 721–734.
43 Plank C, Oberhauser B, Mechtler K, Koch C &
Wagner E (1994) The influence of endosome-disruptive
peptides on gene transfer using synthetic virus-like
gene transfer systems. J Biol Chem 269, 12918–12924.
44 Plank C, Zauner W & Wagner E (1998) Application of
membrane-active peptides for drug and gene delivery
across cellular membranes. Adv Drug Deliver Rev 34,
21–35.
45 Zeng Y & Cullen BR (2002) RNA interference in
human cells is restricted to the cytoplasm. RNA 8,

855–860.
46 Breunig M, Hozsa C, Lungwitz U, Watanabe K,
Umeda I, Kato H & Goepferich A (2008) Mechanistic
investigation of poly(ethyleneimine)-based siRNA
delivery: disulfide bonds boost intracellular release of
the cargo. J Control Release 130, 57–63.
47 Chiu YL & Rana TM (2003) siRNA function in
RNAi: a chemical modification analysis. RNA 9,
1034–1048.
48 Harborth J, Elbashir SM, Vandenburgh K,
Manninga H, Scaringe SA, Weber K & Tuschl T
(2003) Sequence, chemical, and structural variation of
small interfering RNAs and short hairpin RNAs and
the effect on mammalian gene silencing. Antisense
Nucleic Acid Drug Dev 13, 83–105.
49 Czauderna F, Fechtner M, Dames S, Aygu
¨
nH,
Klippel A, Pronk GJ, Giese K & Kaufmann J (2003)
Structural variations and stabilising modifications of
synthetic siRNAs in mammalian cells. Nucleic Acids
Res 31, 2705–2716.
50 Braasch DA, Jensen S, Liu Y, Kaur K, Arar K,
White MA & Corey DR (2003) RNA interference in
mammalian cells by chemically modified RNA.
Biochemistry 42, 7967–7975.
51 Layzer JM, McCaffrey AP, Tanner AK, Huang Z,
Kay MA & Sullenger BA (2004) In vivo activity of
nuclease-resistant siRNAs. RNA 10, 766–771.
52 Jeong JH, Mok H, Oh Y-K & Park TG (2009)

siRNA conjugate delivery systems. Bioconjug Chem
20, 5–14.
53 Hall AHS, Wan J, Shaughnessy EE, Ramsay Shaw B
& Alexander KA (2004) RNA interference using
boranophosphate siRNAs: structure–activity
relationships. Nucleic Acids Res 32, 5991–6000.
54 Wolfrum C, Shi S, Jayaprakash KN, Jayaraman M,
Wang G, Pandey RK, Rajeev KG, Nakayama T,
Charrise K, Ndungo EM et al. (2007) Mechanisms
and optimization of in vivo delivery of lipophilic
siRNAs. Nat Biotechnol 25, 1149–1157.
55 DiFiglia M, Sena-Esteves M, Chase K, Sapp E,
Pfister E, Sass M, Yoder J, Reeves P, Pandey RK,
Rajeev KG et al. (2007) Therapeutic silencing of
mutant huntingtin with siRNA attenuates striatal and
cortical neuropathology and behavioral deficits. Proc
Natl Acad Sci USA 104, 17204–17209.
56 Lupold SE, Hicke BJ, Lin Y & Coffey DS (2002)
Identification and characterization of nuclease-
stabilized RNA molecules that bind human prostate
cancer cells via the prostate-specific membrane antigen.
Cancer Res 62, 4029–4033.
57 McNamara JO, Andrechek ER, Wang Y, Viles KD,
Rempel RE, Gilboa E, Sullenger BA & Giangrande PH
(2006) Cell type-specific delivery of siRNAs with apt-
amer–siRNA chimeras. Nat Biotechnol 24, 1005–1015.
58 Zhou J, Li H, Li S, Zaia J & Rossi JJ (2008) Novel
dual inhibitory function aptamer–siRNA delivery
system for HIV-1 therapy. Mol Ther 16, 1481–1489.
59 Brummelkamp TR, Bernards R & Agami R (2002)

Stable suppression of tumorigenicity by virus-mediated
RNA interference. Cancer Cell 2, 243–247.
60 Jung C-R, Yoo J, Jang YJ, Kim S, Chu I-S, Yeom YI,
Choi JY & Im D-S (2006) Adenovirus-mediated trans-
fer of siRNA against PTTG1 inhibits liver cancer cell
growth in vitro and in vivo . Hepatology 43, 1042–1052.
61 Sabbioni S, Callegari E, Manservigi M, Argnani R,
Corallini A, Negrini M & Manservigi R (2007) Use of
herpes simplex virus type 1-based amplicon vector for
delivery of small interfering RNA. Gene Ther 14, 459–
464.
62 Ma
¨
kinen PI, Koponen JK, Ka
¨
rkka
¨
inen AM, Malm
TM, Pulkkinen KH, Koistinaho J, Turunen MP &
Yla
¨
-Herttuala S (2006) Stable RNA interference:
comparison of U6 and H1 promoters in endothelial
cells and in mouse brain. J Gene Med 8, 433–441.
63 Wadhwa R, Kaul SC, Miyagishi M & Taira K (2004)
Vectors for RNA interference. Curr Opin Mol Ther 6,
367–372.
64 Chen C, Buck AK, Liu X, Winkler M & Reske SN
(2003) Gene silencing by adenovirus-delivered siRNA.
FEBS Lett 539, 111–114.

65 Choy E-Y, Kok K-H, Tsao SW & Jin D-Y (2008)
Utility of Epstein-Barr virus-encoded small RNA
promoters for driving the expression of fusion tran-
scripts harboring short hairpin RNAs. Gene Ther 15,
191–202.
In vivo siRNA delivery M. S. Shim and Y. J. Kwon
4824 FEBS Journal 277 (2010) 4814–4827 ª 2010 The Authors Journal compilation ª 2010 FEBS
66 Tomanin R & Scarpa M (2004) Why do we need new
gene therapy viral vectors? Characteristics, limitations
and future perspectives of viral vector transduction
Curr Gene Ther 4, 357–372.
67 Devroe E & Silver PA (2004) Therapeutic potential of
retroviral RNAi vectors. Expert Opin Biol Ther 4,
319–327.
68 Zimmermann TS, Lee AC, Akinc A, Bramlage B,
Bumcrot D, Fedoruk MN, Harborth J, Heyes JA,
Jeffs LB, John M et al. (2006) RNAi-mediated gene
silencing in non-human primates. Nature 441, 111–114.
69 Gilmore IR, Fox SP, Hollins AJ, Sohail M & Akhtar
S (2004) The design and exogenous delivery of siRNA
for post-transcriptional gene silencing. J Drug Target
12, 315–340.
70 Akhtar S & Benter I (2007) Toxicogenomics of
non-viral drug delivery systems for RNAi: potential
impact on siRNA-mediated gene silencing activity and
specificity. Adv Drug Deliver Rev 59, 164–182.
71 Chien P-Y, Wang J, Carbonaro D, Lei S, Miller B,
Sheikh S, Ali SM, Ahmad MU & Ahmad I (2005)
Novel cationic cardiolipin analogue-based liposome for
efficient DNA and small interfering RNA delivery

in vitro and in vivo. Cancer Gene Ther 12, 321–328.
72 Lia S-D, Chono S & Huang L (2005) Efficient gene
silencing in metastatic tumor by siRNA formulated in
surface-modified nanoparticles. J Control Release 126,
77–84.
73 Peer D, Park EJ, Morishita Y, Carman CV &
Shimaoka M (2008) Systemic leukocyte-directed
siRNA delivery revealing cyclin D1 as an anti-
inflammatory target. Science 319, 627–630.
74 Landen CN Jr, Chavez-Reyes A, Bucana C, Schmandt
R, Deavers MT, Lopez-Berestein G & Sood AK
(2005) Therapeutic EphA2 gene targeting in vivo using
neutral liposomal small interfering RNA delivery.
Cancer Res 65, 6910–6918.
75 Halder J, Kamat AA, Landen CN Jr, Han LY,
Lutgendorf SK, Lin YG, Merritt WM, Jennings NB,
Chavez-Reyes A, Coleman RL et al. (2006) Focal
adhesion kinase targeting using in vivo short interfering
RNA delivery in neutral liposomes for ovarian
carcinoma therapy. Clin Cancer Res 12, 4916–4924.
76 Gray MJ, Van BG, Dallas NA, Xia L, Wang X,
Yang AD, Somcio RJ, Lin YG, Lim S, Fan F et al.
(2008) Therapeutic targeting of neuropilin-2 on
colorectal carcinoma cells implanted in the murine
liver. J Natl Cancer Inst 100, 109–120.
77 Minakuchi Y, Takeshita F, Kosaka N, Sasaki H,
Yamamoto Y, Kouno M, Honma K, Nagahara S,
Hanai K, Sano A et al. (2004) Atelocollagen-mediated
synthetic small interfering RNA delivery for effective
gene silencing in vitro and in vivo. Nucleic Acids Res

32, e109.
78 Takei Y, Kadomatsu K, Yuzawa Y, Matsuo S &
Muramatsu T (2004) A small interfering RNA
targeting vascular endothelial growth factor as cancer
therapeutics. Cancer Res 64 , 3365–3370.
79 Pille
´
J-Y, Li H, Blot E, Bertrand J-R, Pritchard L-L,
Opolon P, Maksimenko A, Lu H, Vannier J-P, Soria
J et al. (2006) Intravenous delivery of anti-RhoA
small interfering RNA loaded in nanoparticles of
chitosan in mice: safety and efficacy in xenografted
aggressive breast cancer. Hum Gene Ther 17,
1019–1026.
80 Howard KA, Paludan SR, Behlke MA, Besenbacher
F, Deleuran B & Kjems J (2009) Chitosan ⁄ siRNA
nanoparticle-mediated TNF- a knockdown in perito-
neal macrophages for anti-inflammatory treatment in a
murine arthritis model. Mol Ther 17, 162–168.
81 Urban-Klein B, Werth S, Abuharbeid S, Czubayko F
& Aigner A (2005) RNAi-mediated gene-targeting
through systemic application of polyethylenimine
(PEI)-complexed siRNA in vivo. Gene Ther 12, 461–
466.
82 Tan P-H, Yang L-C, Shih H-C, Lan K-C & Cheng
J-T (2005) Gene knockdown with intrathecal siRNA
of NMDA receptor NR2B subunit reduces formalin-
induced nociception in the rat. Gene Ther 12, 59–66.
83 Thomas M, Lu JJ, Ge Q, Zhang C, Chen J &
Klibanov AM (2005) Full deacylation of polyethyleni-

mine dramatically boosts its gene delivery efficiency
and specificity to mouse lung. Proc Natl Acad Sci
USA 102, 5679–5684.
84 Grzelinski M, Urban-Klein B, Martens T, Lamszus K,
Bakowsky U, Ho
¨
bel S, Czubayko F & Aigner A
(2006) RNA interference-mediated gene silencing of
pleiotrophin through polyethylenimine-complexed
small interfering RNAs in vivo exerts antitumoral
effects in glioblastoma xenografts. Hum Gene Ther 17,
751–766.
85 Watanabe K, Harada-Shiba M, Suzuki A, Gokuden
R, Kurihara R, Sugao Y, Mori T, Katayama Y &
Niidome T (2009) In vivo siRNA delivery with den-
dritic poly(l-lysine) for the treatment of
hypercholesterolemia. Mol BioSyst 5, 1306–1310.
86 Vangasseri DP, Han SJ & Huang L (2005) Lipid–pro-
tamine–DNA-mediated antigen delivery. Curr Drug
Deliv 2, 401–406.
87 Choi Y-S, Lee JY, Suh JS, Kwon Y-M, Lee S-J,
Chung J-K, Lee D-S, Yang VC, Chung C-P & Park
Y-J (2010) The systemic delivery of siRNAs by a cell
penetrating peptide, low molecular weight protamine.
Biomaterials 31, 1429–1443.
88 Moghimi SM, Symonds P, Murray JC, Hunter AC,
Debska G & Szewczyk A (2005) A two-stage
poly(ethylenimine)-mediated cytotoxicity: implications
for gene transfer ⁄ therapy. Mol Ther 11, 990–995.
M. S. Shim and Y. J. Kwon In vivo siRNA delivery

FEBS Journal 277 (2010) 4814–4827 ª 2010 The Authors Journal compilation ª 2010 FEBS 4825
89 Symonds P, Murray JC, Hunter AC, Debska G,
Szewczyk A & Moghimi SM (2005) Low and high
molecular weight poly(l-lysine)s ⁄ poly(l-lysine)–DNA
complexes initiate mitochondrial-mediated apoptosis
differently. FEBS Lett 579, 6191–6198.
90 Dubey PK, Mishra V, Jain S, Mahor S & Vyas SP
(2004) Liposomes modified with cyclic RGD peptide
for tumor targeting. J Drug Target 12, 257–264.
91 Lu PY, Xie FY & Woodle MC (2005) Modulation of
angiogenesis with siRNA inhibitors for novel
therapeutics. Trends Mol Med 11, 104–113.
92 Gosselin MA & Lee RJ (2002) Folate receptor-
targeted liposomes as vectors for therapeutic agents.
Biotechnol Annu Rev 8, 103–131.
93 Weitman SD, Weinberg AG, Coney LR, Zurawski
VR, Jennings DS & Kamen BA (1992) Cellular
localization of the folate receptor: potential role in
drug toxicity and folate homeostasis. Cancer Res 52,
6708–6711.
94 Kim SH, Mok H, Jeong JH, Kim SW & Park TG
(2006) Comparative evaluation of target-specific GFP
gene silencing efficiencies for antisense ODN, synthetic
siRNA, and siRNA plasmid complexed with PEI–
PEG–FOL conjugate. Bioconjug Chem 17, 241–244.
95 Guo S, Huang F & Guo P (2006) Construction of
folate-conjugated pRNA of bacteriophage phi29 DNA
packaging motor for delivery of chimeric siRNA to
nasopharyngeal carcinoma cells. Gene Ther 13,
814–820.

96 Schiffelers RM, Ansari A, Xu J, Zhou Q, Tang Q,
Storm G, Molema G, Lu PY, Scaria PV & Woodle
MC (2004) Cancer siRNA therapy by tumor selective
delivery with ligand-targeted sterically stabilized
nanoparticle. Nucleic Acids Res 32, e149.
97 Kim B, Tang Q, Biswas PS, Xu J, Schiffelers RM, Xie
FY, Ansari AM, Scaria PV, Woodle MC, Lu P et al.
(2004) Inhibition of ocular angiogenesis by siRNA
targeting vascular endothelial growth factor pathway
genes: therapeutic strategy for herpetic stromal
keratitis. Am J Pathol 165, 2177–2185.
98 Park JW, Hong K, Kirpotin DB, Colbern G, Shalaby
R, Baselga J, Shao Y, Nielsen UB, Marks JD, Moore
D et al. (2002) Anti-HER2 immunoliposomes:
enhanced efficacy attributable to targeted delivery. Clin
Cancer Res 8, 1172–1181.
99 Mamot C, Drummond DC, Noble CO, Kallab V, Guo
Z, Hong K, Kirpotin DB & Park JW (2005) Epidermal
growth factor receptor-targeted immunoliposomes
significantly enhance the efficacy of multiple anticancer
drugs in vivo . Cancer Res 65, 11631–11638.
100 Pirollo KF, Rait A, Zhou Q, Hwang SH, Dagata JA,
Zon G, Hogrefe RI, Palchik G & Chang EH (2007)
Materializing the potential of small interfering RNA
via a tumor-targeting nanodelivery system. Cancer Res
67, 2938–2943.
101 Song E, Zhu P, Lee SK, Chowdhury D, Kussman S,
Dykxhoorn DM, Feng Y, Palliser D, Weiner DB,
Shankar P et al. (2005) Antibody mediated in vivo
delivery of small interfering RNAs via cell-surface

receptors. Nat Biotechnol 23, 709–717.
102 Kumar P, Wu H, McBride JL, Jung K-E, Kim MH,
Davisdon BL, Lee SK, Shankar P & Manjunath N
(2007) Transvascular delivery of small interfering RNA
to the central nervous system. Nature 448, 39–43.
103 Petricciani JC & Patterson RM (1974) Incorporation
of exogenous DNA into mammalian chromosomes.
Nature 249, 649–650.
104 Meade BR & Dowdy SF (2007) Exogenous siRNA
delivery using peptide transduction domains ⁄ cell pene-
trating peptides. Adv Drug Deliver Rev 59, 134–140.
105 Richard JP, Melikov K, Vives E, Ramos C, Verbeure
B, Gait MJ, Chernomordik LV & Lebleu B (2003)
Cell-penetrating peptides: a reevaluation of the
mechanism of cellular uptake. J Biol Chem 278, 585–
590.
106 Lecifert JA, Harkins S & Whitton JL (2002) Full-length
proteins attached to the HIV tat protein transduction
domain are neither transduced between cells, nor exhibit
enhanced immunogenicity. Gene Ther 9, 1422–1428.
107 Schwarze SR, Ho A, Vocero-Akbani A & Dowdy SF
(1999) In vivo protein transduction: delivery of a bio-
logically active protein into the mouse. Science 285,
1569–1572.
108 Torchilin VP, Rammohan R, Weissig V & Levchenko
TS (2001) TAT peptide on the surface of liposomes
affords their efficient intracellular delivery even at low
temperature and in the presence of metabolic inhibi-
tors. Proc Natl Acad Sci USA 98, 8786–8791.
109 Vives E, Brodin P & Lebleu B (1997) A truncated

HIV-1 Tat protein basic domain rapidly translocates
through the plasma membrane and accumulates in the
cell nucleus. J Biol Chem 272, 16010–16017.
110 Morris MC, Vidal P, Chaloin L, Heitz F & Divita G
(1997) A new peptide vector for efficient delivery of
oligonucleotides into mammalian cells. Nucleic Acids
Res 25, 2730–2736.
111 Derossi D, Joliot AH, Chassaing G & Prochiantz A
(1994) The third helix of the antennapedia homeodo-
main translocates through biological membranes.
J Biol Chem 269, 10444–10450.
112 Futaki S, Suzuki T, Ohashi W, Yagami T, Tanaka S,
Ueda K & Sugiura Y (2001) Arginine-rich peptides.
An abundant source of membrane-permeable peptides
having potential as carriers for intracellular protein
delivery. J Biol Chem 276, 5836–5840.
113 Kim WJ, Christensen LV, Jo S, Yockman JW, Jeong
JH, Kim YH & Kim SW (2006) Cholesteryl oligoargi-
nine delivering vascular endothelial growth factor
siRNA effectively inhibits tumor growth in colon
adenocarcinoma. Mol Ther 14, 343–350.
In vivo siRNA delivery M. S. Shim and Y. J. Kwon
4826 FEBS Journal 277 (2010) 4814–4827 ª 2010 The Authors Journal compilation ª 2010 FEBS
114 Kim SW, Kim NY, Choi YB, Park SH, Yang JM &
Shin S (2010) RNA interference in vitro and in vivo
using an arginine peptide ⁄ siRNA complex system.
J Control Release 143, 335–343.
115 Han L, Zhang A, Wang H, Pu P, Jiang X, Kang C &
Chang J (2010) Tat–BMPs–PAMAM conjugates
enhance therapeutic effect of small interference RNA

on U251 glioma cells in vitro and in vivo. Hum Gene
Ther 21, 417–426.
116 Crombez L, Morris MC, Dufort S, Aldrian-Herrada
G, Nguyen Q, McMaster G, Coll J-L, Heitz F &
Divita G (2009) Targeting cyclin B1 through peptide-
based delivery of siRNA prevents tumour growth.
Nucleic Acids Res 37, 4559–4569.
117 Endoh T, Sisido M & Ohtsuki T (2008) Cellular
siRNA delivery mediated by a cell-permeant RNA-
binding protein and photoinduced RNA interference.
Bioconjug Chem 19, 1017–1024.
118 Turner JJ, Ivanova GD, Verbeure B, Williams D, Ar-
zumanov AA, Abes S, Lebleu B & Gait MJ (2005) Cell-
penetrating peptide conjugates of peptide nucleic acids
(PNA) as inhibitors of HIV-1 Tat-dependent trans-acti-
vation in cells. Nucleic Acids Res 33, 6837–6849.
119 Wadia JS, Stan RV & Dowdy SF (2004) Transducible
TAT–HA fusogenic peptide enhances escape of TAT-
fusion proteins after lipid raft macropinocytosis. Nat
Med 10, 310–315.
120 Michiue H, Tomizawa K, Wei F-Y, Matsushita M, Lu
Y-F, Ichikawa T, Tamiya T, Date I & Matsui H
(2005) The NH
2
terminus of influenza virus hemagglu-
tinin-2 subunit peptides enhances the antitumor
potency of polyarginine-mediated p53 protein trans-
duction. J Biol Chem 280, 8285–8289.
121 Kwon EJ, Bergen JM & Pun SH (2008) Application of
an HIV gp41-derived peptide for enhanced intracellu-

lar trafficking of synthetic gene and siRNA delivery
vehicles. Bioconjug Chem 19, 920–927.
122 Heyes J, Palmer L, Bremner K & MacLachlan I
(2005) Cationic lipid saturation influences intracellular
delivery of encapsulated nucleic acids. J Control
Release 107, 276–287.
123 Hassani Z, Lemkine GF, Erbacher P, Palmier K,
Alfama G, Giovannangeli C, Behr J-P & Demeneix
BA (2005) Lipid-mediated siRNA delivery down-regu-
lates exogenous gene expression in the mouse brain at
picomolar levels. J Gene Med 7, 198–207.
124 Xu P, Li S-Y, Li Q, Van Kirk EA, Ren J, Murdoch
WJ, Zhang Z, Radosz M & Shen Y (2008) Virion-
mimicking nanocapsules from pH-controlled hierarchi-
cal self-assembly for gene delivery. Angew Chem Int Ed
47, 1260–1264.
125 Oishi M, Nagasaki Y, Itaka K, Nishiyama N &
Kataoka K (2005) Lactosylated poly(ethylene glycol)–
siRNA conjugate through acid-labile b-thiopropionate
linkage to construct pH-sensitive polyion complex
micelles achieving enhanced gene silencing in
hepatoma cells. J Am Chem Soc 127, 1624–1625.
126 Knorr V, Russ V, Allmendinger L, Ogris M & Wagner
E (2008) Acetal linked oligoethylenimines for use as pH-
sensitive gene carriers. Bioconjug Chem 19, 1625–1634.
127 Shin J, Shum P & Thompson DH (2003) Acid-trig-
gered release via dePEGylation of DOPE liposomes
containing acid-labile vinyl ether PEG-lipids. J Control
Release 91, 187–200.
128 Hatakeyama H, Ito E, Akita H, Oishi M, Nagasaki Y,

Futaki S & Harashima H (2009) A pH-sensitive
fusogenic peptide facilitates endosomal escape and
greatly enhances the gene silencing of siRNA-contain-
ing nanoparticles in vitro and in vivo. J Control Release
139, 127–132.
129 Shim MS & Kwon YJ (2009) Acid-responsive
linear polyethylenimine for efficient, specific, and
biocompatible siRNA delivery. Bioconjug Chem 20,
488–499.
M. S. Shim and Y. J. Kwon In vivo siRNA delivery
FEBS Journal 277 (2010) 4814–4827 ª 2010 The Authors Journal compilation ª 2010 FEBS 4827

×