ArticlePDF AvailableLiterature Review

Yi, Y, Hahm, SH and Lee, KH. Retroviral gene therapy: safety issues and possible solutions. Curr Gene Ther 5: 25-35

Authors:

Abstract and Figures

The recent incidents of leukemia development in X-SCID patients after a successful treatment of the disease with retroviral gene therapy raised concerns regarding the safety of the use of retroviral vectors in clinical gene therapy. In this review, we have tried to re-evaluate the safety issues related to the use of retroviral vectors in human clinical trials and to suggest possible appropriate solutions to the issues. As revealed by the X-SCID incident, oncogenesis caused by retroviral insertional activation of host genes is one of the most prominent risks. An ultimate solution to this problem will be in re-engineering retroviral vectors so that the retroviral insertion takes place only at the desired specific sites of the host cell chromosome. This is, however, a technically demanding tasks, and it will take years to develop retroviral vectors with targeted insertion capability. In the mean time, the use of chromatin insulators can reduce chances for retrovirus-mediated oncogenesis by inhibiting non-specific activation of nearby cellular proto-oncogenes. Co-transduction of a suicidal gene under the control of an inducible promoter could also be one of the important safety features, since destruction of transduced cells can be triggered if abnormal growth is observed. Additionally, conditional expression of the transgene only in appropriate target cells via the combination of targeted transduction, cell type-specific expression, and targeted local administration will increase the overall safety of the retroviral systems. Finally, splitting of the viral genome, use of self-inactivating (SIN) retroviral vectors, or complete removal of the coding sequences for gag, pol, and env genes is desirable to virtually eliminate the possibility of generation of replication competent retroviruses (RCR).
Content may be subject to copyright.
Current Gene Therapy, 2005, 5, 25-35 25
1566-5232/05 $50.00+.00 © 2005 Bentham Science Publishers Ltd.
Retroviral Gene Therapy: Safety Issues and Possible Solutions
Youngsuk Yi
1
, Sung Ho Hahm
2
and Kwan Hee Lee
1,3,
*
1
TissueGene Inc., Gaithersburg, MD 20877;
2
Omega Biosciences LLC, Baltimore, MD 21227;
3
Clinical Research
Center, College of Medicine, Inha University, Inchon, South Korea 400-711
Abstract: The recent incidents of leukemia development in X-SCID patients after a successful treatment of the
disease with retroviral gene therapy raised concerns regarding the safety of the use of retroviral vectors in clinical
gene therapy. In this review, we have tried to re-evaluate the safety issues related to the use of retroviral vectors in
human clinical trials and to suggest possible appropriate solutions to the issues. As revealed by the X-SCID
incident, oncogenesis caused by retroviral insertional activation of host genes is one of the most prominent risks.
An ultimate solution to this problem will be in re-engineering retroviral vectors so that the retroviral insertion
takes place only at the desired specific sites of the host cell chromosome. This is, however, a technically demanding
tasks, and it will take years to develop retroviral vectors with targeted insertion capability. In the mean time, the use
of chromatin insulators can reduce chances for retrovirus-mediated oncogenesis by inhibiting non-specific
activation of nearby cellular proto-oncogenes. Co-transduction of a suicidal gene under the control of an inducible
promoter could also be one of the important safety features, since destruction of transduced cells can be triggered if
abnormal growth is observed. Additionally, conditional expression of the transgene only in appropriate target cells
via the combination of targeted transduction, cell type-specific expression, and targeted local administration will
increase the overall safety of the retroviral systems. Finally, splitting of the viral genome, use of self-inactivating
(SIN) retroviral vectors, or complete removal of the coding sequences for gag, pol, and env genes is desirable to
virtually eliminate the possibility of generation of replication competent retroviruses (RCR).
Keywords: Retroviral, gene, therapy, safety, insertional, targeted, vectors, insulator.
INTRODUCTION
Retroviral vectors have been the most preferred gene
transfer systems in clinical gene therapy due to its well-
understood biology and its high efficiency of transduction
[Somia and Verma, 2000; Thomas et al., 2003]. In spite of
the popularity, the possibility of oncogenic transformation of
host cells as a result of non-specific insertion of retroviral
DNA into the host chromosome has always been a concern.
Although it is generally believed that a single insertional
mutation is not sufficient to develop a malignant phenotype
by itself, a leukemia induction has been reported in an
animal model [Li et al., 2003]. The potential risk of
insertional oncogenesis was realized in a human trial for the
X-linked severe combined immunodeficiency (SCID)
[Hacein-Bey-Abina et al., 2003; Kaiser, 2003; Kohn et al.,
2003]. In the trial, infants with X-SCID who would
otherwise have little chance of survival were actually cured
by retrovirus-mediated ex-vivo gene transfer, and the trial
was credited as the first unequivocal success for gene therapy
[Cavazzana-Calvo et al., 2000]. Unfortunately, however, two
out of nine successfully treated patients later developed
leukemia, and it is generally believed that leukemeogenesis
was triggered by unexpected activation of a cellular proto-
oncogene as a result of retroviral integration. Even though no
major adverse effects have been reported in many other
retroviral gene therapy clinical trials, including the ones
*Address correspondence to this author at the TissueGene, Inc. 209 Perry
Parkway, Suite 13, Gaithersburg, MD 20877, USA; Tel: 301-921-6000;
Fax: 301-921-6011; E-mail: ortholee@tissuegene.com
using more committed hematopoietic cells and mature
lymphocytes [Anderson, 2000], it is now time to re-evaluate
all the safety issues concerning the use of retroviral vectors
and to address them fully.
Stable incorporation of retroviral DNA into the host
genome is in itself advantageous, as long-term expression of
the transgene is possible which is often required for
achieving therapeutic efficacy. However, a non-specific
incorporation of viral DNA throughout the host genome can
either cause a disruption of a host gene at the site of
incorporation or cause an abnormal expression of nearby host
genes driven by the enhancer of the inserted viral DNA. If
the insertional interference occurs against a host gene
involved in a critical cellular function such as cell cycle
progression, it can become a major cause of cell
transformation and oncogenesis, especially in the presence of
additional physiological and/or genetic insults.
In addition to the risk of insertional mutagenesis, another
safety issue with the retroviral vector is the possibility of
generating replication competent retroviruses (RCR).
Although new generations of retroviral vectors have been
designed to reduce the production of RCR, additional efforts
are required to ensure the complete elimination of the
problem.
The recently developed lentiviral gene transfer system
shares many features of the retroviral system. The viral
genome integrates into host chromosomes, and the inserted
gene can be maintained in the cells permanently. In contrast
to conventional retroviral vectors requiring cell division for
26 Current Gene Therapy, 2005, Vol. 5, No. 1 Lee et al.
infection, lentiviral vectors can infect efficiently non-
dividing cells and dividing cells. Therefore, it can be applied
for transgene expression in neuronal cells. Many of the
lentiviral vectors used in gene therapy are based on the
human immunodeficiency virus (HIV). The major limita-
tions of using lentiviral vectors in clinical trial are the safety
concerns related to their HIV origin. Recent trials of
lentiviral vector system will be described in relevant
chapters.
In this review, we will try to discuss about
aforementioned and additional safety issues relating to the
use of retroviral vectors as a tool for therapeutic gene
delivery, and to suggest possible available solutions to these
issues and future directions for an overall increase in the
safety and efficiency of retroviral gene therapy protocols.
Other important parameters in cell-based ex-vivo retroviral
gene therapy protocols are already well covered elsewhere
[Baum et al., 2003], including cell type and numbers,
observation period, proliferation capacity, age of patients,
immunity, and side effects of transgene expression.
TARGETED RETROVIRAL TRANSDUCTION
Current retroviral transfer system lacks specificity for
target cell types. It makes the system unsuitable for
protocols requiring gene
transfer to particular cell types in a
mixed cell population. Therefore, vector targeting
is highly
desirable and will become a prerequisite for certain ex vivo
gene therapy
protocols.
Infection of target cells by retroviruses is initiated by
binding of the viral envelope protein to cell surface receptors
[Hunter and Swanstrom, 1990]. Infusion of viral and cellular
membranes leads to the internalization of the viral core
[White, 1992]. In the past years, various retrovirus receptors,
coreceptors and cofactors have been identified and studied for
their role in viral entry [Overbaugh et al., 2001], and
attempts have been made to engineer viral envelope proteins
and cellular receptors for attaining changes in the viral
tropism [Ting et al., 1998; Chaudry et al., 1999].
In an approach for a targeted transduction, envelope
protein modification by attaching a peptide ligand such as
epidermal growth factor (EGF) receptor binding domain to
the NH
2
-terminus of the envelope glycoprotein (SU) was
attempted [Cosset et al., 1995; Peng et al., 1997; Peng et
al., 1999]. Incorporation of the chimeric envelope protein
into the viral particle allows binding of retrovirus to the
receptor-positive target cells. The subsequent viral entry
steps are blocked, however, because EGF receptors do not
support these processes. A protease cleavable linker was used
in this chimeric protein to join the peptide ligand and SU,
so that an appropriate cellular protease can cleave the attached
ligand. Factor Xa [Nilson et al., 1996; Fielding et al.,
1998], plasmin [Peng et al., 1999], matrix-metalloproteases
(MMPs) [Peng et al., 1997], or intracellular protein conver-
tases [Buchholz et al., 1998] were used for this purpose.
Because overexpression of MMPs is frequently associated
with angiogenesis, inflammation, and cancer invasion,
MMPs are considered to be interesting targets for the
protease-activatable gene delivery systems. Using MMP-
activatable retroviral vectors, selective transduction of MMP-
rich tumor cells was achieved in a heterogeneous cell
population, but with somewhat reduced efficiency of
transduction [Peng et al., 1999; Schneider et al., 2003].
Similarly, targeted infection for the high-molecular-weight
melanoma-associated antigen (HMWMAA) expressing
tumors was achieved by fusing a single chain antibody
recognizing HMWMAA to the amino terminus of the surface
domain of MLV with a matrix metalloprotease-2 (MMP2)
cleavage site linker [Martin et al., 2002].
Matrix targeting is another approach and matrix-targeted
retroviral vectors were found to be more efficient than
unmodified vectors [Gordon et al., 2000; Hall et al., 2000].
A matrix-targeted retroviral vector was constructed by
attaching collagen-binding polypeptide sequence to the
Table 1. Retroviral Vectors for Targeted Infection
Modification Target cell references
EGF-MMP cleavable linker chimeric env Cancer invasion, angiogenesis, inflammation [Peng et al., 1997; Buchholz et al., 1998; Peng et
al., 1999]
IL-2 chimeric env IL-2 R [Maurice et al., 1999]
EGF chimeric env EGFR [Cosset et al., 1995]
SCF-Factor Xa chimeric env Stem cell (Kit) [Fielding et al., 1998]
vWF (collagen binding) chimeric env Cancer (collagen expressing); vascular lesion [Gordon et al., 2000; Hall et al., 2000]
Single-chain variable fragmented antibody
(scFv) for EGFRvIII
Cancer (brain, breast, lung, ovary) [Lorimer and Lavictoire, 2000]
scFv for HMWMAA Cancer [Martin et al., 2002]
scFv from phage display T cell [Engelstadter et al., 2000]
scFv for Carcino embryonic antigen (CEA) Cancer [Khare et al., 2001]
Receptor pseudotyping (CD4 and CXCR4) HIV-1 infected cell [Somia et al., 2000; Bittner et al., 2002]
Retroviral Gene Therapy Current Gene Therapy, 2005, Vol. 5, No. 1 27
amino-terminal region of the amphotropic 4070A envelope
protein. Because tumor development and accompanied
angiogenesis are associated with remodeling of extracellular
matrix components, these vectors accumulate at sites of
tumor development with newly exposed collagens. When the
matrix-targeted retroviral vector expressing dominant mutant
cyclin G1 was administered by portal vein infusion, vector
particles accumulated in the angiogenic tumor vasculature
within 1 hour of infusion. These vectors transduced tumor
cells with high efficiency and reduced the volume of tumor
[Gordon et al., 2000].
Targeting retroviral delivery to quiescent interleukin-2
(IL-2)-dependent cells was also reported [Maurice et al.,
1999]. In this report, chimeric amphotropic MLV envelope
glycoprotein fused with IL-2 was used for a direct binding of
the viral particles to the IL-2 receptors expressed on G0/G1
arrested cells, resulting in a transient stimulation of cell
proliferation. Subsequent viral entry was mediated by
unmodified envelope proteins co-expressed on the same virus
particles. A 34-fold increase in transduction efficiency was
observed with this method. Additionally, targeting efforts
for T cells [Engelstadter et al., 2000], and other cancer cells
[Lorimer and Lavictoire, 2000; Khare et al., 2001] were
reported. For the specific transduction of HIV-envelope
expressing cells, envelope pseudotyping was used to create
hybrid CD4/CXCR4 receptors for MLV retrovirus [Somia et
al., 2000; Bittner et al., 2002] and lentivirus [Somia et al.,
2000; Bittner et al., 2002], In order to improve transduction
efficiency frequently observed to be low for the targeted
retroviral vectors, binding defective but fusion competent
hemagglutinin (HA) protein has also been tried [Lin et al.,
2001].
LOCAL DELIVERY
If local delivery of retroviral vectors is available for an
effective treatment of a disease, it will be generally safer than
systemic delivery in terms of toxicology and long term side
effects. A number of studies have shown the efficacy and
safety of locally delivered retroviral vectors. A retroviral
vector expressing antiproliferative dominant negative mutant
cyclin G1 (dnG1) was successfully used for the prevention of
eximer laser-induced corneal haze [Behrens et al., 2002].
Biodistribution study after the treatment of surgically
induced rabbits with eye drops containing dnG1 retroviral
vectors showed no evidence of vector dissemination in non-
target organs. Localized delivery of lentiviral vectors into the
substantia nigra of adult rats has also been tried [Deglon et
al., 2000]. In a phase I clinical trial for direct intratumoral
injection of interferon-γ retroviral vectors in advanced
melanoma patients, viral injection was well tolerated and no
toxicity was reported [Nemunaitis et al., 1999]. This
suggests that the direct injection approach is feasible for
treating solid tumors with retroviral vectors.
In terms of potential problems associated with
concomitant transduction of surrounding non-target cells, ex
vivo cell-based gene therapy with local delivery can be a
better choice, if extra time and expenses are tolerated. As an
example, an ex vivo cell-mediated gene therapy trial has been
performed successfully for the treatment of artificially
induced hyaline cartilage damage in animals, by injecting
TGF-β1-retrovirus transduced fibroblasts into the knee joints
[Lee et al., 2001].
INTEGRATION OF RETROVIRUS INTO THE HOST
CHROMOSOME
In the ex-vivo gene therapy trials to treat the rare immune
deficiency disorder X-SCID, patients were treated with
autologous hematopoietic stem cells transduced with a
recombinant retrovirus expressing the common gamma chain
(γ
c
) of interleukin receptor [Cavazzana-Calvo et al., 2000].
Although, nine out of eleven treated children showed
dramatic improvements with almost fully restored immune
systems, two of the nine cured patients developed leukemia
three years after the treatment. A copy of the vector DNA
was found in the first intron of the growth-promoting LMO2
gene of the leukemic clones in the first patient, and
approximately 3 kb upstream of the first exon of the same
gene in the second [Hacein-Bey-Abina et al., 2003; Kaiser,
2003; Kohn et al., 2003]. LMO2 is a LIM domain
transcription regulator involved in hematopoiesis [Yamada et
al., 1998] and is reported to be activated in T cell leukemia
by chromosomal translocation [Royer-Pokora et al., 1991].
Deregulation of LMO2 expression can affect hematopoiesis
and lymphocyte development, and consequently can be the
cause of leukemic transformation. However, it is not yet
clear if deregulation of LMO2 alone was solely responsible
for the oncogenic transformation. Other contributing factors
such as immune deficiency shared by the two X-SCID
patients or the expression of the therapeutic transgene
(interleukin receptor γ
c
) itself, by mediating proliferative
cellular signaling events, might also have contributed in the
transformation process.
The U.S. Food and Drug Administration (FDA) put a
hold on U.S. trials using retroviruses to insert genes into
blood stem cells, and the U.S. National Institutes of Health
(NIH) Recombinant DNA Advisory Committee (RAC) urged
about 90 other retroviral trials targeting blood cells to
consider a pause. Subsequently, a public meeting was held
on February 10, 2003, and the NIH RAC made the
following tentative recommendations [Friedmann, 2003a].
1. Pending further data or extenuating circumstances
reviewed on a case-by-case basis, retroviral gene transfer
studies for X-linked SCID should be limited to patients
who have failed identical or haploidentical stem-cell
transplantation or for whom no suitable stem cell donor
can be identified.
2. There are not sufficient data or reports of adverse events
directly attributable to the use of retroviral vectors at
this time to warrant cessation of other retroviral human
gene transfer studies, including studies for non-X-linked
SCID. Such studies may be justified contingent upon
appropriate risk/benefit analysis accompanied by
implementation of appropriate informed consent and
monitoring plans.
The general consensus of the experts of the review boards
in the US and other countries is that the benefits might
outweigh the risks in most SCID-related retroviral gene
therapy trials. FDA has since announced its willingness to
remove the clinical hold on most retroviral gene therapy
28 Current Gene Therapy, 2005, Vol. 5, No. 1 Lee et al.
trials involving hematopoietic stem cells [Friedmann,
2003a]. These trials will likely be allowed to continue with
several necessary precautions as advised by the NIH RAC
meeting held on June 18-19, 2003 [Friedmann, 2003b].
These include a complete integration analysis, modification
of vector design to become self-inactivating (SIN),
conditional expression of the transgenes, incorporation of
suicidal elements or silencers, longer-term studies in large
animals, and close monitoring of clonally expanding cells.
Although FDA has not yet announced specific
recommendations for other non-SCID retroviral gene therapy
trials, Gene Therapy Advisory Committee (GTAC) and
Committee on Safety of Medicines (CSM) Working Party
on Retroviruses in United Kingdom excluded other retroviral
gene therapy trials from the intensified watch list.
TARGETING RETROVIRAL INTEGRATION
After entering the host cell, a single-stranded retroviral
RNA genome is released into the cytoplasm and converted
into a double-stranded DNA by virus-encoded reverse
transcriptase. The viral DNA then forms a large
nucleoprotein structure, termed pre-integration complex,
containing proteins necessary for nuclear localization and
insertion of viral DNA into the host genome. Although the
protein components and the exact mechanism of action of the
complex is still not completely understood, it has been
demonstrated that viral integrase (IN) catalyzes
the key DNA
cutting and joining reactions for inserting viral
DNA into the
host genome [Katz and Skalka, 1994; Wei et al., 1997].
Recent experimental evidence suggests that retroviral
integration is not a completely random process but favors
sites of active gene expression [Schroder et al., 2002; Wu et
al., 2003]. Murine leukemia virus (MLV) can only infect
actively dividing cells in which proto-oncogenes play major
roles in cell-cycle progression or other important cellular
processes. This means that the risk of retrovirus insertion-
mediated abnormal regulation of cellular proto-oncogenes can
be much higher than anticipated. In addition, while MLV
prefers integration near the start of transcriptional units, HIV-
1 integrates anywhere in the transcriptional units [Schroder et
al., 2002; Wu et al., 2003]. The difference of integration
preferences indicates that there are possible fundamental
differences between MLV and HIV for the mechanism of
integration. Elucidating the mechanisms of integration and
establishing the database for preferred integration sites could
permit a better prediction of the integration sites of retroviral
vectors. This may eventually lead to the development of
retroviral vectors capable of integrating to the specific sites
of the host cell chromosome, providing the ultimate solution
to the problems of insertional mutagenesis.
There have been attempts to target retroviral integration
to pre-selected locations of the host genome by fusing viral
integrase with sequence-specific DNA binding domains
obtained from phage lambda repressor, bacterial LexA, or a
zinc finger protein zif268 [Goulaouic and Chow, 1996; Katz
et al., 1996; Bushman and Miller, 1997]. However, these
trials show only a limited success, as the specificity of
integration was only partially altered. In a separate
experiment, Bovine Leukemia Virus (BLV) integrase was
used for site-specific integration of naked DNA to the pre-
integrated integrase recognition sequence of mouse genome
[Tanaka et al., 1998]. Similarly, site-specific integration of
naked DNA into human chromosome 8 has been attempted
with limited success using modified phage φC-31 integrase
[Sclimenti et al., 2001]. In this study, enhanced sequence
specificity and increased integrase efficiency was achieved
through a directed evolution strategy. It is clear that
concentrated efforts are required in defining the precise
mechanism of action of the retroviral pre-integration complex
and in designing modified integrases with sequence-specific
integration capability. The latter may be accomplished either
by rational modification of the protein or by using the
directed evolution approach [Chen, 2001]. One example of
rational modification is fusing integrase with synthetic zinc
finger motifs with defined sequence specificities [Bushman
and Miller, 1997; Kim and Pabo, 1998; Jamieson et al.,
2003]. Directed evolution utilizes error-prone PCR-driven
mutagenesis, recombination, or DNA shuffling, combined
with a high throughput screening for the selection of
modified proteins with significantly improved function. The
newly developed integrases should also maintain the ability
to form a pre-integration complex with a high-level of
infection capability.
INSULATORS TO PREVENT POSITIONAL
EFFECTS AND INSERTIONAL ONCOGENE ACTI-
VATION
Retroviruses are often susceptible to positional effects
and transcriptional silencing depending on the site of
integration in the chromosome [Pannell and Ellis, 2001]. In
order to overcome positional silencing effect, chromatin
insulators have been used in retroviral vectors. Chromatin
insulators are believed to form expression boundaries [Sun
and Elgin, 1999; Burgess-Beusse et al., 2002] and can block
positive and negative positional effects at the site of
integration when they flank a transgene [Kellum and Schedl,
1991; Chung et al., 1993; Chung et al., 1997]. They
prevent interferences between promoters and enhancers of
adjacent genes [Labrador and Corces, 2002]. As an example,
when a 1.2 kb chromatin insulator obtained from the chicken
β-globin locus control region hypersensitive site 4 (cHS4)
was inserted in the retrovirus 3’ LTR, protection of the
positional effects was observed from transduced cultured
cells and from the mice transplanted with transduced marrow
cells [Emery et al., 2000]. Similarly, cHS4 insulator used
with gamma-globin expression cassette increased the
likelihood of stable gamma-globin expression nearly 10-fold,
allowing for the expression at the therapeutic range for
treating sickle cell anemia and beta thalassemia in mouse
bone marrow transplantation models [Emery et al., 2002].
Because sequences in the 3’ LTR of retroviral vectors are
copied to the 5’ LTR during the processing of viral genome
into the provirus, if an insulator is inserted in the 3’ LTR of
the recombinant vector, a barrier of insulators will be formed
surrounding the transgene (Fig. (1A)). In addition, insulators
are inserted in the place of the U3 region of the 3’ LTR,
which is the viral enhancer region. These vectors will
become self-inactivating (SIN) in the proviral form, as there
is no viral enhancer required to produce replication
competent retrovirus. Therefore, insulator containing
retroviral vectors will be less prone to silencing of the
Retroviral Gene Therapy Current Gene Therapy, 2005, Vol. 5, No. 1 29
transgene expression as a result of chromosome positional
effect [Emery et al., 2000], and at the same time, will have
less chance of causing aberrant induction of host genes near
the site of incorporation as it has no viral enhancer.
Boundaries formed by insulators will also prevent the
influence of an internal heterologous enhancer used to drive
transgene expression on the transcription of nearby host
genes, although it has yet to be experimentally proven.
Although retroviral insertion can cause either a disruption
or an abnormal activation of host genes, the latter is a
primary concern in terms of oncogenesis. This is because, in
most cases, retroviral insertion will occur in only one allele
of the host genome leaving the other locus intact, and
insertional disruption of the host gene will become
problematic only in rare cases where haplo-insufficiency is
phenotypically relevant for oncogenic transformation. Thus,
although chromatin insulators cannot prevent host gene
disruption by retroviral insertion, the benefits associated
with the use of insulators preventing unwanted activation of
host genes will be rather significant.
The efficiency of insulator function is, however,
dependent on several factors including topological
constraints, cell types, and the state of cell differentiation
[Rivella et al., 2000; Yannaki et al., 2002]. Also the size
limitations of retroviral vectors should be considered. A 265
bp sea urchin insulator termed sns (silencing nucleoprotein
structure) was found to be effective for insulator function in
human cells, and thus may be useful in retroviral vectors [Di
Simone et al., 2001]. In a recent report, anti-repressor
elements were identified by screening a library of human
genomic DNA fragments between 500 and 2,000 bp, based
on their ability to relieve LexA-dependent transcription
repression [Kwaks et al., 2003]. These elements can confer
high and stable transgene expression in mammalian cells
when they were used to flank the transgene, suggesting that
they play a similar role as insulators. Additional experiments
are required to further improve the function of these and
other insulator elements for more efficient inhibition of
positional repression of the transgene expression and
inhibition of insertional activation of host genes, at the same
time.
Scaffold (or matrix) Attachment Region (SAR) is another
DNA sequence element that is believed to play an important
role in defining boundaries of independent chromatin
domains [Bode et al., 1992; Namciu et al., 1998]. SARs
bind to the nuclear scaffold or nuclear matrix with high
affinity and are proposed to form chromosomal loops [Bode
et al., 2000]. SARs have been used in retroviral vectors with
an enhancement of transgene expression in several different
cell types [Agarwal et al., 1998; Kurre et al., 2003]. A
recent report shows that a high-level transgene expression can
be achieved from a self-inactivating (SIN) lentiviral vector
containing both the human interferon-beta scaffold
attachment region and the chicken beta-globin insulator
[Ramezani and Hawley, 2003]. The proviral form of this
vector does not contain HIV-1 U3 region transcriptional
regulatory elements and is flanked by the enhancer-blocking
beta-globin insulators. These observations indicate that the
usage of SARs in addition to insulators could significantly
improve transgene expression and lower the risk of
uncontrolled activation of cellular proto-oncogenes at or near
the site of incorporation.
Activation of proto-oncogene may also arise due to the
retroviral RNA processing. A strong internal splice acceptor
(SA) is recommended after the splice donor (SD) of retroviral
vector to reduce the positional effects of inserted retroviral
RNA processing on the expression of the transgene as well
as the disrupted host gene. Combination of strong splice
acceptor, deleting cryptic SD in the transgene [Knipper et.
al., 2001], using an improved polyadenylation signal
[Isamail et al., 2001], the removal of LTR promoter, and the
insulator will largely prevent the interactions of the retroviral
splice donor with downstream chromosome sequences.
TRANSCRIPTIONAL TARGETING
Transcriptional targeting using cell type-specific
promoters and enhancers can be applied either alone or in
combination with targeted transduction to minimize the
expression of transduced genes in non-target cells and
thereby reducing potential side effects. (Table 2) summarizes
examples of cell type-specific promoters and enhancers used
for transcriptional targeting in retroviral gene therapy. As an
example, promoters of oncogenes overexpressed in the tumor
cells can be the target for tumor specific promoters (e.g. c-
erbB2 and c-myc). For the treatment of malignant
melanomas, tyrosinase promoter was used for the expression
of HSV-tk or IL-2 [Vile et al., 1994]. Tyrosinase is rate-
limiting enzyme for melanin production, which is highly
expressed in melanomas. In addition to cell type-specific
promoters, inducible or regulatable expression systems can
also be used for increased safety and efficacy. In the case for
mammary tumor and prostate cancer, the growth of tumor is
hormone dependent. Thus, use of combined steroid
hormone-responsive or cell type specific promoters is an
attractive approach for retroviral gene therapy of these cancers
[Sparmann et al., 1994]. Additionally, genes that are
induced by cancer therapy (e.g. by Gamma-irradiation and
chemotherapy) can also be the targets for the regulatory
elements [Walther et al., 1997; Walther et al., 2000].
Although cell type-specific promoters are used
successfully in many different trials, the overall efficiency of
transcription achieved from these promoters is relatively
weak compared with the generally used viral promoters. In a
recent report, hypoxic and cytokine-inducible enhancers, both
of which are active in a tumor environment, are combined
with endothelial cell-specific E-selectin and VEGF receptor 2
promoters [Modlich et al., 2000] to achieve a maximum
possible tumor endothelium-specific transcri-ption. In
another report, human α-fetoprotein (AFP) enhancer was
combined with a house keeping gene phosphoglycerate
kinase-1 (PGK-1) promoter, to augment the activity of the
weak tumor-selective AFP promoter [Cao et al., 2001].
Rat alpha-fetoprotein promoter was used as a cell type-
specific promoter for a lentivirally transduced expression in
human hepatocarcinoma cells [Uch et al., 2003]. Replace-
ment of U3 region of the lentiviral LTR with an upstream
enhancer (HS2) of the erythroid-specific GATA-1 gene and
HIV-1 promoter showed a high level of transgene
expression specifically in mature erythroblasts [Lotti and
Mavilio, 2003].
30 Current Gene Therapy, 2005, Vol. 5, No. 1 Lee et al.
Finally, in many cases, the size constraints of retroviral
vector limit the use of enhancers, which are generally long in
size. Therefore, construction of minimum enhancer/promoter
cassettes with strategic combinations of different sequence
elements will be required to facilitate the efficacy of gene
therapy trials.
CO-EXPRESSION OF A SUICIDAL GENE
Herpes Simplex Virus thymidine kinase (HSV-tk) has
been used for selective destruction of cells in several different
settings. When anti-viral prodrug nucleobase analogue
ganciclovir (GCV) is applied to HSV-tk expressing cells,
GCV is efficiently converted into monophosphate form by
HSV-tk, and then into cytotoxic triphosphate derivatives by
cellular kinases. Actively dividing cells will be killed as
they incorporate the nucleotide derivatives into their genome.
In allogeneic bone marrow transplantation (BMT), donor T
cells are able to mediate anti-leukemic effects but they can
also induce Graft-vs-Host Disease (GvHD), which is often
fatal. In an attempt to reduce GvHD while maintaining anti-
leukemic effect, scientists have retrovirally transduced HSV-
tk to donor T-cells before being used in animal
myeloablative BMT trials [Drobyski et al., 2003]. At first,
the donor T-cells helped to eliminate residual malignant
leukemic cells, but when signs of GvHD development were
noticed, proliferating donor T-cells were rapidly destroyed by
treating the animals with ganciclovir. When this strategy was
used in a human trial, three out of eight patients treated with
donor lymphocytes transduced with HSV-TK gene could be
effectively controlled by ganciclovir-induced elimination of
the transduced cells when they developed GvHD 12 months
after transduction [Bonini et al., 1997].
Similarly, retroviral vectors can be designed to co-express
HSV-tk suicide gene to be used as a safety switch, in
addition to a therapeutic gene. In this case, only if abnormal
growth of transduced cells, such as the leukemic T-cell
clones shown in the X-SCID case, is observed, treating with
ganciclovir can theoretically eliminate all the transduced
cells. However, constitutive expression of HSV-tk can also
induce the death of neighboring uninfected cells (bystander
effects), when ganciclovir is administered. In order to
minimize unwanted side effects due to bystander effects, the
use of cell type-specific or inducible promoter for the
expression of HSV-tk or the use of other pro-apoptotic genes
with a minimum bystander effect may be advantageous. As
an example, lentivirally transduced expression unit
containing the rat alpha-fetoprotein promoter was used to
restrict the HSV-tk induced GCV sensitivity to human
hepatocarcinoma cells [Uch et al., 2003].
On the other hand, retroviral vectors expressing HSV-tk
have been used in antitumor treatment trials [Greco et al.,
2002; Orchard et al., 2002; Barzon et al., 2003]. In this
case, maximum “bystander effect” is required to kill
neighboring uninfected tumor cells as well as infected cells.
The results of tumor treatment with HSV-tk expressing
Table 2. Cell Type-specific Promoters and Enhancers for Transcriptional Targeting in Retroviral Gene Therapy
Promoter Target cell/tissue Transgene References
PEPCK promoter Hepatocyte Neo, bovine growth hormone [Hatzoglou et al., 1990]
hAAT promoter Hepatocyte Alpha I antitrypsin [Hafenrichter et al., 1994]
MMTV-LTR Mammary gland TNF-α [Sparmann et al., 1994]
MCK promoter Muscle β-galactosidase, dystrophin
minigene
[Ferrari et al., 1995]
AFP promoter Cancer: Hepatocellular carcinomas HSV-tk, VZV-tk [Ido et al., 1995]
Tyrosine promoter Cancer: Melanomas HSV-tk, IL-2 [Vile et al., 1994]
Col1a1 promoter Bone β-geo (β-gal, neo fusion) [Stover et al., 2001]
HSP70 promoter Cancer Dominant negative IGF-IR [Romano et al., 2001]
WAP promoter Cancer: Mammary β-galactosidase [Ozturk-Winder et al., 2002]
ppET1 promoter Cancer: Endothelium β-galactosidase [Mavria et al., 2000]
AFP enhancer; PGK promoter Cancer: Hepatocellular carcinomas HSV-tk [Cao et al., 2001]
HRE, PGK-1 enhancer; E-selectin,
KDR promoter
Cancer: Endothelium TNF-α, luciferase [Modlich et al., 2000]
HRE enhancer; AFP promoter Cancer: Hepatocellular carcinomas HSV-tk, luciferase [Ido et al., 2001]
Rat alpha-fetoprotein Human hepatocarcinoma cell HSV-tk, luciferase [Uch et al., 2003]
HS2 of erythroid-specific GATA-1
gene; HIV-1 promoter
Mature erythroblasts GFP [Lotti and Mavilio, 2003]
Retroviral Gene Therapy Current Gene Therapy, 2005, Vol. 5, No. 1 31
retroviral vectors were, however, not fully successful due to
low infection efficiency and low bystander effects.
One potential obstacle for co-expressing HSV-tk suicide
gene as a safety switch in addition to the therapeutic gene is
the limited insert size constraint of the retroviral vector. In
cell-based ex-vivo gene therapy and when the use of clonally-
derived cells is allowed, selecting for single clones
transduced with both the HSV-tk gene and the therapeutic
gene expressed from separate retroviral vectors can be a
solution for the size problem.
AVOIDING REPLICATION COMPETENT RETRO-
VIRUS (RCR)
Generation of RCR remains a potential safety issue in
retroviral gene therapy. Retroviral vectors transfected into the
packaging cell line can produce RCR by recombination
processes between homologous sequences of the retroviral
vector DNA and the gag, pol, and env coding sequences in
the packaging systems. In order to lower the chances for
recombination, both minimizing the homologous sequences
and physically separating genes for gag, pol and env into
two different expression cassettes have become standard
practices. However, residual gag, pol, and env coding
sequences are frequently included in these vectors in an
attempt to increase transduction efficiency and viral titer,
raising a concern for RCR generation. In a recent report, a
complete removal of residual coding sequences for gag, pol,
and env genes was shown without any detrimental effect on
viral transduction efficiency, and thus these vectors further
reduced the chance for RCR generation [Yu et al., 2000].
Same concept for the safer vector by reducing the chance of
recombination was also tried in the lentiviral vector. To
avoid the production of Replication Competent Lentivirus
(RCL), components required for the production of lentivirus
were divided into at least three parts. Vector plasmid
contains gene of interest and the minimal cis-acting element
of HIV. Packaging plasmid has all HIV viral genes except
the env gene. Envelope proteins were provided from a
plasmid containing the envelope gene via co-transfection.
Typically, the Glycoprotein from Vesicular Stomatitis Virus
(VSV-G) is used as an envelope gene [Burns et al., 1993].
Safety concerns specific to HIV virus are recombination of
the vector sequences with the endogenous HIV sequences in
the HIV positive patients. Even though most endogenous
retroviral sequences in the human genome have large
deletions, the possibility of recombination and mobilization
cannot be overlooked. Because the probability of
recombination during the packaging process through
transient transfection is expected to be much higher than that
occurring when stable producer cell lines are used, stable
lentivirus producer cell line has been tried [Farson et al.,
2001; Xu et al., 2001]. Due to the major concern of using
HIV in humans, researchers are developing nonhuman
lentivirus vector systems. Simian Immunodeficiency Virus
(SIV), Feline Immunodeficiency Virus (FIV), and Equine
Immunodeficiency Virus (EIAV) have been used [Poeschla et
al., 1998; Pandya et al., 2001; Bienemann et al., 2003].
However, the safety features of non-primate lentiviruses in
humans have yet to be determined [Connolly, 2002].
Self-inactivating (SIN) retroviral vector was developed by
introducing a deletion in the U3 region of the 3’ LTR which
contains all the enhancer and promoter activities of the viral
vector (Fig. (1B)) [Yu et al., 1986]. Because the 5’LTR of
the provirus carrying the same deletion will be incapable of
inducing transcription for the production of packagable
RNAs, no active viral particle can be made from this vector
even after successful recombination processes to acquire gag,
pol, and env sequences. Additional advantage of the SIN
vector is that because of the lack of promoter activities of the
viral LTRs, the possibility of LTR-mediated activation of
cellular proto-oncogenes near the site of incorporation is also
minimized. SIN vector approach has been tested more
extensively in lentiviruses [Zufferey et al., 1998]. Although
SIN vector was considered as an ideal vector, a very low
level of SIN vector mobilization was detected [Xu et al.,
2001]. Additionally, one of the drawbacks of the SIN
retroviral vectors is the internal promoter’s low
transcriptional activity compared to the viral LTR in a
number of different cell types. Improvements in the design
of the internal promoter/enhancer will be required to
overcome this obstacle.
Split-intron retroviral vector has been shown to enhance
expression of an inserted gene and improve its safety [Ismail
et al., 2000]. A strong synthetic splice donor site and a
splice acceptor site were inserted between U3 and R of the 3’
LTR and downstream of the packaging signal, respectively.
During the reverse transcription process, the strong synthetic
splice donor site introduced in the 3’ LTR is copied into
5’LTR, and theoretically all the transcripts made in the
transduced cells are spliced and the packaging signal is
removed (Fig. (1C)). Therefore, the possibility of producing
RCR will be greatly reduced.
Finally, activation of proto-oncogene may arise due to
aberrant retroviral RNA processing. A strong internal splice
acceptor (SA) is recommended after the splice donor (SD)
site of retroviral vector to prevent a generation of aberrant
readthrough transcripts containing both the transgene and a
portion of the disrupted host gene. Combination of strong
SA, deleting cryptic SD in the transgene [Knipper et al.,
2001], using an improved polyadenylation signal [Ismail et
al., 2001], the removal of LTR promoter, and the presence
of an insulator will largely prevent the interactions of the
retroviral splice donor with downstream chromosome
sequences.
CONCLUDING REMARKS
We have reviewed safety features for the use of retroviral
vector systems in clinical gene therapy. Targeted infection,
local delivery, targeted retroviral insertion, insulators,
transcriptional targeting, co-transduction with a suicidal
gene, and SIN vectors were suggested as possible solutions
for the potential risks of retroviral gene therapy. Some of
these precautions can be applied to gene therapy protocols
using other viral and non-viral vector systems. One major
immediate concern in terms of retroviral gene therapy, as
revealed by the X-SCID case, is insertional oncogenesis.
Several possible approaches for minimizing, if not
completely eliminating, insertional oncogenesis were
considered in depth. In addition to the more widely used
32 Current Gene Therapy, 2005, Vol. 5, No. 1 Lee et al.
retroviral systems that were discussed above, foamy viruses
may be used as a safe and efficient means of targeting non-
dividing cells [Mergia et al., 2001]. Foamy viruses are
known to have a broad host range, without causing any
disease and persist in infected humans [Schweizer et al.,
1997; Heneine et al., 1998; Callahan et al., 1999].
In the case of ex vivo cell-based gene therapy where the
selection of clonally-derived cells is permissible, transduced
cells can be prescreened to select for clones with the insertion
of the transgene only at a desirable site of the chromosome,
minimizing the chances for insertional oncogenesis.
Applications of gene therapy protocols have been
continuously expanding to include a wide variety of acquired
and inherited diseases, such as cancer, SCID, and other life
threatening diseases. Retroviral gene therapy approaches for
the treatment of each of these diseases will have different
treatment requirements and thus will meet variable
challenges for addressing safety issues. Treatment of
orthopedic indications may be one of the most promising
areas of gene therapy research and trials, despite the general
perception that they are the non-lethal, acquired conditions
[Evans et al., 2000; Evans and Scully, 2000; Lee et al.,
2001; Lee et al., 2002; Lieberman et al., 2002; Palmer et
al., 2002; Nussenbaum et al., 2003; Robbins et al., 2003].
Treatment of these conditions will require only a short-term
expression of the transduced genes with more readily
available options for the delivery. Combined with
aforementioned safety precautions such as incorporation of
suicidal genes or chromatin insulators, cell-based ex vivo
retroviral gene therapy has a unique window of opportunity
to bring about a rewarding outcome.
ACKNOWLEDGEMENTS
We thank Lillian Yun, Vivian Yip, and Jonathan Yun for
proofreading this manuscript.
REFERENCES
Agarwal, M., Austin, T. W., Morel, F., Chen, J., Bohnlein, E. and Plavec, I.
(1998) Scaffold attachment region-mediated enhancement of
retroviral vector expression in primary T cells. J Virol. 72: 3720-8.
Anderson, W. F. (2000) Gene therapy. The best of times, the worst of
times. Science. 288: 627-9.
Barzon, L., Bonaguro, R., Castagliuolo, I., Chilosi, M., Franchin, E., Del
Vecchio, C., Giaretta, I., Boscaro, M. and Palu, G. (2003) Gene
therapy of thyroid cancer via retrovirally-driven combined expression
of human interleukin-2 and herpes simplex virus thymidine kinase.
Eur. J. Endocrinol., 148: 73-80.
Baum, C., Dullmann, J., Li, Z., Fehse, B., Meyer, J., Williams, D. A. and
von Kalle, C. (2003) Side effects of retroviral gene transfer into
hematopoietic stem cells. Blood. 101: 2099-114.
Behrens, A., Gordon, E. M., Li, L., Liu, P. X., Chen, Z., Peng, H., La Bree,
L., Anderson, W. F., Hall, F. L. and McDonnell, P. J. (2002) Retroviral
gene therapy vectors for prevention of excimer laser-induced corneal
haze. Invest. Ophthalmol. Vis. Sci., 43: 968-77.
Bienemann, A. S., Martin-Rendon, E., Cosgrave, A. S., Glover, C. P.,
Wong, L. F., Kingsman, S. M., Mitrophanous, K. A., Mazarakis, N. D.
and Uney, J. B. (2003) Long-term replacement of a mutated
nonfunctional CNS gene: reversal of hypothalamic diabetes insipidus
using an EIAV-based lentiviral vector expressing arginine
vasopressin. Mol. Ther., 7: 588-96.
Fig. (1). Designs of the improved retroviral vectors. (A) Insulator located in U3 of 3’LTR is copied to 5’LTR in target cell. Insulators
located in the both ends of vector DNA blocks the cross activation between retroviral DNA and the chromosomal DNA. (B) Due to the
duplication of U3-deleted LTR of the SIN vector in target cell, the proviral form does not contain viral enhancer any more, and thus
require an internal promoter for the expression of the transgene. Arrows show the transcription start. P, internal promoter; U3,
deletion of U3 sequence. (C) Comparison of splicing in conventional and split-intron vectors. Arrows show transcripts from the
vector DNA. SD, Splice Donor site; SA, Splice Acceptor site.
Retroviral Gene Therapy Current Gene Therapy, 2005, Vol. 5, No. 1 33
Bittner, A., Mitnacht-Kraus, R. and Schnierle, B. S. (2002) Specific
transduction of HIV-1 envelope expressing cells by retroviral vectors
pseudotyped with hybrid CD4/CXCR4 receptors. J. Virol. Methods.
104: 83-92.
Bode, J., Benham, C., Knopp, A. and Mielke, C. (2000) Transcriptional
augmentation: modulation of gene expression by scaffold/matrix-
attached regions (S/MAR elements). Crit. Rev. Eukaryot. Gene Expr.,
10: 73-90.
Bode, J., Kohwi, Y., Dickinson, L., Joh, T., Klehr, D., Mielke, C. and
Kohwi-Shigematsu, T. (1992) Biological significance of unwinding
capability of nuclear matrix-associating DNAs. Science, 255: 195-7.
Bonini, C., Ferrari, G., Verzeletti, S., Servida, P., Zappone, E., Ruggieri, L.,
Ponzoni, M., Rossini, S., Mavilio, F., Traversari, C. and Bordignon, C.
(1997) HSV-TK gene transfer into donor lymphocytes for control of
allogeneic graft-versus-leukemia. Science, 276: 1719-24.
Buchholz, C. J., Peng, K. W., Morling, F. J., Zhang, J., Cosset, F. L. and
Russell, S. J. (1998) In vivo selection of protease cleavage sites from
retrovirus display libraries. Nat. Biotechnol., 16: 951-4.
Burgess-Beusse, B., Farrell, C., Gaszner, M., Litt, M., Mutskov, V.,
Recillas-Targa, F., Simpson, M., West, A. and Felsenfeld, G. (2002)
The insulation of genes from external enhancers and silencing
chromatin. Proc. Natl. Acad. Sci. USA, 99 Suppl 4: 16433-7.
Burns, J. C., Friedmann, T., Driever, W., Burrascano, M. and Yee, J. K.
(1993) Vesicular stomatitis virus G glycoprotein pseudotyped
retroviral vectors: concentration to very high titer and efficient gene
transfer into mammalian and nonmammalian cells. Proc. Natl. Acad.
Sci. USA, 90: 8033-7.
Bushman, F. D. and Miller, M. D. (1997) Tethering human
immunodeficiency virus type 1 preintegration complexes to target
DNA promotes integration at nearby sites. J. Virol., 71: 458-64.
Callahan, M. E., Switzer, W. M., Matthews, A. L., Roberts, B. D., Heneine,
W., Folks, T. M. and Sandstrom, P. A. (1999) Persistent zoonotic
infection of a human with simian foamy virus in the absence of an
intact orf-2 accessory gene. J. Virol., 73: 9619-24.
Cao, G., Kuriyama, S., Gao, J., Nakatani, T., Chen, Q., Yoshiji, H., Zhao,
L., Kojima, H., Dong, Y., Fukui, H. and Hou, J. (2001) Gene therapy
for hepatocellular carcinoma based on tumour-selective suicide gene
expression using the alpha-fetoprotein (AFP) enhancer and a
housekeeping gene promoter. Eur. J. Cancer, 37: 140-7.
Cavazzana-Calvo, M., Hacein-Bey, S., de Saint Basile, G., Gross, F.,
Yvon, E., Nusbaum, P., Selz, F., Hue, C., Certain, S., Casanova, J. L.,
Bousso, P., Deist, F. L. and Fischer, A. (2000) Gene therapy of human
severe combined immunodeficiency (SCID)-X1 disease. Science,
288: 669-72.
Chaudry, G. J., Farrell, K. B., Ting, Y. T., Schmitz, C., Lie, S. Y.,
Petropoulos, C. J. and Eiden, M. V. (1999) Gibbon ape leukemia virus
receptor functions of type III phosphate transporters from CHOK1
cells are disrupted by two distinct mechanisms. J. Virol., 73: 2916-20.
Chen, R. (2001) Enzyme engineering: rational redesign versus directed
evolution. Trends Biotechnol., 19: 13-4.
Chung, J. H., Bell, A. C. and Felsenfeld, G. (1997) Characterization of the
chicken beta-globin insulator. Proc. Natl. Acad. Sci. USA, 94: 575-80.
Chung, J. H., Whiteley, M. and Felsenfeld, G. (1993) A 5' element of the
chicken beta-globin domain serves as an insulator in human erythroid
cells and protects against position effect in Drosophila. Cell, 74: 505-
14.
Connolly, J. B. (2002) Lentiviruses in gene therapy clinical research. Gene
Ther., 9: 1730-4.
Cosset, F. L., Morling, F. J., Takeuchi, Y., Weiss, R. A., Collins, M. K. and
Russell, S. J. (1995) Retroviral retargeting by envelopes expressing an
N-terminal binding domain. J. Virol., 69: 6314-22.
Di Simone, P., Di Leonardo, A., Costanzo, G., Melfi, R. and Spinelli, G.
(2001) The sea urchin sns insulator blocks CMV enhancer following
integration in human cells. Biochem. Biophys. Res. Commun., 284:
987-92.
Drobyski, W. R., Gendelman, M., Vodanovic-Jankovic, S. and Gorski, J.
(2003) Elimination of leukemia in the absence of lethal graft-versus-
host disease after allogenic bone marrow transplantation. J. Immunol.,
170: 3046-53.
Emery, D. W., Yannaki, E., Tubb, J., Nishino, T., Li, Q. and Stamatoy-
annopoulos, G. (2002) Development of virus vectors for gene therapy
of beta chain hemoglobinopathies: flanking with a chromatin insulator
reduces gamma-globin gene silencing in vivo. Blood, 100: 2012-9.
Emery, D. W., Yannaki, E., Tubb, J. and Stamatoyannopoulos, G. (2000) A
chromatin insulator protects retrovirus vectors from chromosomal
position effects. Proc. Natl. Acad. Sci. USA, 97: 9150-5.
Engelstadter, M., Bobkova, M., Baier, M., Stitz, J., Holtkamp, N., Chu, T.
H., Kurth, R., Dornburg, R., Buchholz, C. J. and Cichutek, K. (2000)
Targeting human T cells by retroviral vectors displaying antibody
domains selected from a phage display library. Hum. Gene Ther., 11:
293-303.
Evans, C. H., Ghivizzani, S. C., Herndon, J. H., Wasko, M. C., Reinecke,
J., Wehling, P. and Robbins, P. D. (2000) Clinical trials in the gene
therapy of arthritis. Clin. Orthop., S300-7.
Evans, C. H. and Scully, S. P. (2000) Orthopaedic gene therapy. Clin
Orthop. S2.
Farson, D., Witt, R., McGuinness, R., Dull, T., Kelly, M., Song, J., Radeke,
R., Bukovsky, A., Consiglio, A. and Naldini, L. (2001) A new-
generation stable inducible packaging cell line for lentiviral vectors.
Hum. Gene Ther., 12: 981-97.
Ferrari, G., Salvatori, G., Rossi, C., Cossu, G. and Mavilio, F. (1995) A
retroviral vector containing a muscle-specific enhancer drives gene
expression only in differentiated muscle fibers. Hum. Gene Ther., 6:
733-42.
Fielding, A. K., Maurice, M., Morling, F. J., Cosset, F. L. and Russell, S. J.
(1998) Inverse targeting of retroviral vectors: selective gene transfer
in a mixed population of hematopoietic and nonhematopoietic cells.
Blood, 91: 1802-9.
Friedmann, T. (2003a) Gene therapy's new era: a balance of unequivocal
benefit and unequivocal harm. Mol. Ther., 8: 5-7.
Friedmann, T. (2003b) Slide presentation of Review of Selected American
Society of Gene Therapy (ASGT) Annual Meeting Sessions Related to
Retroviral Vectors. NIH RAC meeting held on June 18-19 2003
Website: http://www.webconferences.com/nihoba/18-19june2003.html].
Gordon, E. M., Liu, P. X., Chen, Z. H., Liu, L., Whitley, M. D., Gee, C.,
Groshen, S., Hinton, D. R., Beart, R. W. and Hall, F. L. (2000)
Inhibition of metastatic tumor growth in nude mice by portal vein
infusions of matrix-targeted retroviral vectors bearing a cytocidal
cyclin G1 construct. Cancer Res. 60: 3343-7.
Goulaouic, H. and Chow, S. A. (1996) Directed integration of viral DNA
mediated by fusion proteins consisting of human immunodeficiency
virus type 1 integrase and Escherichia coli LexA protein. J. Virol., 70:
37-46.
Greco, E., Fogar, P., Basso, D., Stefani, A. L., Navaglia, F., Zambon, C. F.,
Mazza, S., Gallo, N., Piva, M. G., Scarpa, A., Pedrazzoli, S. and
Plebani, M. (2002) Retrovirus-mediated herpes simplex virus
thymidine kinase gene transfer in pancreatic cancer cell lines: an
incomplete antitumor effect. Pancreas, 25: e21-9.
Hacein-Bey-Abina, S., von Kalle, C., Schmidt, M., Le Deist, F., Wulffraat,
N., McIntyre, E., Radford, I., Villeval, J. L., Fraser, C. C., Cavazzana-
Calvo, M. and Fischer, A. (2003) A serious adverse event after
successful gene therapy for X-linked severe combined
immunodeficiency. N. Engl. J. Med., 348: 255-6.
Hafenrichter, D. G., Ponder, K. P., Rettinger, S. D., Kennedy, S. C., Wu,
X., Saylors, R. S. and Flye, M. W. (1994) Liver-directed gene therapy:
evaluation of liver specific promoter elements. J. Surg. Res., 56: 510-
7.
Hall, F. L., Liu, L., Zhu, N. L., Stapfer, M., Anderson, W. F., Beart, R. W.
and Gordon, E. M. (2000) Molecular engineering of matrix-targeted
retroviral vectors incorporating a surveillance function inherent in von
Willebrand factor. Hum. Gene Ther., 11: 983-93.
Hatzoglou, M., Lamers, W., Bosch, F., Wynshaw-Boris, A., Clapp, D. W.
and Hanson, R. W. (1990) Hepatic gene transfer in animals using
retroviruses containing the promoter from the gene for
phosphoenolpyruvate carboxykinase. J. Biol. Chem., 265: 17285-93.
Heneine, W., Switzer, W. M., Sandstrom, P., Brown, J., Vedapuri, S.,
Schable, C. A., Khan, A. S., Lerche, N. W., Schweizer, M.,
Neumann-Haefelin, D., Chapman, L. E. and Folks, T. M. (1998)
Identification of a human population infected with simian foamy
viruses. Nat. Med., 4: 403-7.
Hunter, E. and Swanstrom, R. (1990) Retrovirus envelope glycoproteins.
Curr. Top. Microbiol. Immunol., 157: 187-253.
Ido, A., Nakata, K., Kato, Y., Nakao, K., Murata, K., Fujita, M., Ishii, N.,
Tamaoki, T., Shiku, H. and Nagataki, S. (1995) Gene therapy for
hepatoma cells using a retrovirus vector carrying herpes simplex virus
thymidine kinase gene under the control of human alpha-fetoprotein
gene promoter. Cancer Res., 55: 3105-9.
Ido, A., Uto, H., Moriuchi, A., Nagata, K., Onaga, Y., Onaga, M., Hori, T.,
Hirono, S., Hayashi, K., Tamaoki, T. and Tsubouchi, H. (2001) Gene
therapy targeting for hepatocellular carcinoma: selective and
enhanced suicide gene expression regulated by a hypoxia-inducible
34 Current Gene Therapy, 2005, Vol. 5, No. 1 Lee et al.
enhancer linked to a human alpha-fetoprotein promoter. Cancer Res.,
61: 3016-21.
Ismail, S. I., Kingsman, S. M., Kingsman, A. J. and Uden, M. (2000) Split-
intron retroviral vectors: enhanced expression with improved safety.
J. Virol., 74: 2365-71.
Ismail, S. I., Rohll, J. B., Kingsman, S. M., Kingsman, A. J. and Uden, M.
(2001) Use of intron-disrupted polyadenylation sites to enhance
expression and safety of retroviral vectors. J. Virol., 75: 199-204.
Jamieson, A. C., Miller, J. C. and Pabo, C. O. (2003) Drug discovery with
engineered zinc-finger proteins. Nat. Rev. Drug Discov., 2: 361-8.
Kaiser, J. (2003) Gene therapy. Seeking the cause of induced leukemias in
X-SCID trial. Science, 299: 495.
Katz, R. A., Merkel, G. and Skalka, A. M. (1996) Targeting of retroviral
integrase by fusion to a heterologous DNA binding domain: in vitro
activities and incorporation of a fusion protein into viral particles.
Virology, 217: 178-90.
Katz, R. A. and Skalka, A. M. (1994) The retroviral enzymes. Annu. Rev.
Biochem., 63: 133-73.
Kellum, R. and Schedl, P. (1991) A position-effect assay for boundaries of
higher order chromosomal domains. Cell, 64: 941-50.
Khare, P. D., Shao-Xi, L., Kuroki, M., Hirose, Y., Arakawa, F.,
Nakamura, K. and Tomita, Y. (2001) Specifically targeted killing of
carcinoembryonic antigen (CEA)-expressing cells by a retroviral
vector displaying single-chain variable fragmented antibody to CEA
and carrying the gene for inducible nitric oxide synthase. Cancer Res.,
61: 370-5.
Kim, J. S. and Pabo, C. O. (1998) Getting a handhold on DNA: design of
poly-zinc finger proteins with femtomolar dissociation constants. Proc.
Natl. Acad. Sci. USA, 95: 2812-7.
Knipper, R., Kuehlcke, K., Schiedlmeier, B., Hildinger, M., Lindemann, C.,
Schilz, A. J., Fauser, A. A., Fruehauf, S., Zeller, W. J., Ostertag, W.,
Eckert, H. G. and Baum, C. (2001) Improved post-transcriptional
processing of an MDR1 retrovirus elevates expression of multidrug
resistance in primary human hematopoietic cells. Gene Ther., 8: 239-
46.
Kohn, D. B., Sadelain, M. and Glorioso, J. C. (2003) Occurrence of
leukaemia following gene therapy of X-linked SCID. Nat. Rev.
Cancer, 3: 477-88.
Kurre, P., Morris, J., Thomasson, B., Kohn, D. B. and Kiem, H. P. (2003)
Scaffold attachment region-containing retrovirus vectors improve
long-term proviral expression after transplantation of GFP-modified
CD34+ baboon repopulating cells. Blood, 102: 3117-9.
Kwaks, T. H., Barnett, P., Hemrika, W., Siersma, T., Sewalt, R. G., Satijn,
D. P., Brons, J. F., Van Blokland, R., Kwakman, P., Kruckeberg, A. L.,
Kelder, A. and Otte, A. P. (2003) Identification of anti-repressor
elements that confer high and stable protein production in mammalian
cells. Nat. Biotechnol., 21: 553-8.
Labrador, M. and Corces, V. G. (2002) Setting the boundaries of
chromatin domains and nuclear organization. Cell, 111: 151-4.
Lee, C. W., Martinek, V., Usas, A., Musgrave, D., Pickvance, E. A.,
Robbins, P., Moreland, M. S., Fu, F. H. and Huard, J. (2002) Muscle-
based gene therapy and tissue engineering for treatment of growth
plate injuries. J. Pediatr. Orthop., 22: 565-72.
Lee, K. H., Song, S. U., Hwang, T. S., Yi, Y., Oh, I. S., Lee, J. Y., Choi, K.
B., Choi, M. S. and Kim, S. J. (2001) Regeneration of hyaline cartilage
by cell-mediated gene therapy using transforming growth factor beta
1-producing fibroblasts. Hum. Gene Ther., 12: 1805-13.
Lieberman, J. R., Ghivizzani, S. C. and Evans, C. H. (2002) Gene transfer
approaches to the healing of bone and cartilage. Mol. Ther., 6: 141-7.
Lin, A. H., Kasahara, N., Wu, W., Stripecke, R., Empig, C. L., Anderson,
W. F. and Cannon, P. M. (2001) Receptor-specific targeting mediated
by the coexpression of a targeted murine leukemia virus envelope
protein and a binding-defective influenza hemagglutinin protein. Hum.
Gene Ther., 12: 323-32.
Lorimer, I. A. and Lavictoire, S. J. (2000) Targeting retrovirus to cancer
cells expressing a mutant EGF receptor by insertion of a single chain
antibody variable domain in the envelope glycoprotein receptor
binding lobe. J. Immunol. Methods, 237: 147-57.
Lotti, F. and Mavilio, F. (2003) The choice of a suitable lentivirus vector:
transcriptional targeting. Methods Mol. Biol., 229: 17-27.
Martin, F., Chowdhury, S., Neil, S., Phillipps, N. and Collins, M. K. (2002)
Envelope-targeted retrovirus vectors transduce melanoma xenografts
but not spleen or liver. Mol. Ther., 5: 269-74.
Maurice, M., Mazur, S., Bullough, F. J., Salvetti, A., Collins, M. K., Russell,
S. J. and Cosset, F. L. (1999) Efficient gene delivery to quiescent
interleukin-2 (IL-2)-dependent cells by murine leukemia virus-
derived vectors harboring IL-2 chimeric envelope glycoproteins.
Blood, 94: 401-10.
Mavria, G., Jager, U. and Porter, C. D. (2000) Generation of a high titre
retroviral vector for endothelial cell-specific gene expression in vivo.
Gene Ther., 7: 368-76.
Mergia, A., Chari, S., Kolson, D. L., Goodenow, M. M. and Ciccarone, T.
(2001) The efficiency of simian foamy virus vector type-1 (SFV-1) in
nondividing cells and in human PBLs. Virology, 280: 243-52.
Modlich, U., Pugh, C. W. and Bicknell, R. (2000) Increasing endothelial
cell specific expression by the use of heterologous hypoxic and
cytokine-inducible enhancers. Gene Ther., 7: 896-902.
Namciu, S. J., Blochlinger, K. B. and Fournier, R. E. (1998) Human matrix
attachment regions insulate transgene expression from chromosomal
position effects in Drosophila melanogaster. Mol. Cell Biol., 18: 2382-
91.
Nemunaitis, J., Fong, T., Robbins, J. M., Edelman, G., Edwards, W.,
Paulson, R. S., Bruce, J., Ognoskie, N., Wynne, D., Pike, M., Kowal,
K., Merritt, J. and Ando, D. (1999) Phase I trial of interferon-gamma
(IFN-gamma) retroviral vector administered intratumorally to patients
with metastatic melanoma. Cancer Gene Ther., 6: 322-30.
Nilson, B. H., Morling, F. J., Cosset, F. L. and Russell, S. J. (1996)
Targeting of retroviral vectors through protease-substrate interactions.
Gene Ther., 3: 280-6.
Nussenbaum, B., Rutherford, R. B., Teknos, T. N., Dornfeld, K. J. and
Krebsbach, P. H. (2003) Ex vivo gene therapy for skeletal
regeneration in cranial defects compromised by postoperative
radiotherapy. Hum. Gene Ther., 14: 1107-15.
Orchard, P. J., Blazar, B. R., Burger, S., Levine, B., Basso, L., Nelson, D.
M., Gordon, K., McIvor, R. S., Wagner, J. E. and Miller, J. S. (2002)
Clinical-scale selection of anti-CD3/CD28-activated T cells after
transduction with a retroviral vector expressing herpes simplex virus
thymidine kinase and truncated nerve growth factor receptor. Hum.
Gene Ther., 13: 979-88.
Overbaugh, J., Miller, A. D. and Eiden, M. V. (2001) Receptors and entry
cofactors for retroviruses include single and multiple transmembrane-
spanning proteins as well as newly described glycophosphatidyl-
inositol-anchored and secreted proteins. Microbiol. Mol. Biol. Rev.,
65: 371-89, table of contents.
Ozturk-Winder, F., Renner, M., Klein, D., Muller, M., Salmons, B. and
Gunzburg, W. H. (2002) The murine whey acidic protein promoter
directs expression to human mammary tumors after retroviral
transduction. Cancer Gene Ther., 9: 421-31.
Palmer, G., Pascher, A., Gouze, E., Gouze, J. N., Betz, O., Spector, M.,
Robbins, P. D., Evans, C. H. and Ghivizzani, S. C. (2002)
Development of gene-based therapies for cartilage repair. Crit. Rev.
Eukaryot. Gene Expr., 12: 259-73.
Pandya, S., Boris-Lawrie, K., Leung, N. J., Akkina, R. and Planelles, V.
(2001) Development of an Rev-independent, minimal simian
immunodeficiency virus-derived vector system. Hum. Gene Ther., 12:
847-57.
Pannell, D. and Ellis, J. (2001) Silencing of gene expression: implications
for design of retrovirus vectors. Rev. Med. Virol., 11: 205-17.
Peng, K. W., Morling, F. J., Cosset, F. L., Murphy, G. and Russell, S. J.
(1997) A gene delivery system activatable by disease-associated
matrix metalloproteinases. Hum. Gene Ther., 8: 729-38.
Peng, K. W., Vile, R., Cosset, F. L. and Russell, S. (1999) Selective
transduction of protease-rich tumors by matrix-metalloproteinase-
targeted retroviral vectors. Gene Ther., 6: 1552-7.
Poeschla, E. M., Wong-Staal, F. and Looney, D. J. (1998) Efficient
transduction of nondividing human cells by feline immunodeficiency
virus lentiviral vectors. Nat. Med., 4: 354-7.
Ramezani, A. and Hawley, R. G. (2003) Human immunodeficiency virus
type 1-based vectors for gene delivery to human hematopoietic stem
cells. Methods Mol. Med., 76: 467-92.
Rivella, S., Callegari, J. A., May, C., Tan, C. W. and Sadelain, M. (2000)
The cHS4 insulator increases the probability of retroviral expression
at random chromosomal integration sites. J Virol. 74: 4679-87.
Robbins, P. D., Evans, C. H. and Chernajovsky, Y. (2003) Gene therapy
for arthritis. Gene Ther., 10: 902-11.
Romano, G., Reiss, K., Tu, X., Peruzzi, F., Belletti, B., Wang, J. Y.,
Zanocco-Marani, T. and Baserga, R. (2001) Efficient in vitro and in
vivo gene regulation of a retrovirally delivered pro-apoptotic factor
under the control of the Drosophila HSP70 promoter. Gene Ther., 8:
600-7.
Royer-Pokora, B., Loos, U. and Ludwig, W. D. (1991) TTG-2, a new gene
encoding a cysteine-rich protein with the LIM motif, is overexpressed
Retroviral Gene Therapy Current Gene Therapy, 2005, Vol. 5, No. 1 35
in acute T-cell leukaemia with the t(11;14)(p13;q11). Oncogene, 6:
1887-93.
Schneider, R. M., Medvedovska, Y., Hartl, I., Voelker, B., Chadwick, M.
P., Russell, S. J., Cichutek, K. and Buchholz, C. J. (2003) Directed
evolution of retroviruses activatable by tumour-associated matrix
metalloproteases. Gene Ther., 10: 1370-80.
Schroder, A. R., Shinn, P., Chen, H., Berry, C., Ecker, J. R. and Bushman,
F. (2002) HIV-1 integration in the human genome favors active genes
and local hotspots. Cell, 110: 521-9.
Schweizer, M., Falcone, V., Gange, J., Turek, R. and Neumann-Haefelin,
D. (1997) Simian foamy virus isolated from an accidentally infected
human individual. J. Virol., 71: 4821-4.
Sclimenti, C. R., Thyagarajan, B. and Calos, M. P. (2001) Directed
evolution of a recombinase for improved genomic integration at a
native human sequence. Nucleic Acids Res., 29: 5044-51.
Somia, N. and Verma, I. M. (2000) Gene therapy: trials and tribulations.
Nat. Rev. Genet., 1: 91-9.
Somia, N. V., Miyoshi, H., Schmitt, M. J. and Verma, I. M. (2000)
Retroviral vector targeting to human immunodeficiency virus type 1-
infected cells by receptor pseudotyping. J. Virol., 74: 4420-4.
Sparmann, G., Walther, W., Gunzburg, W. H., Uckert, W. and Salmons, B.
(1994) Conditional expression of human TNF-alpha: a system for
inducible cytotoxicity. Int. J. Cancer, 59: 103-7.
Stover, M. L., Wang, C. K., McKinstry, M. B., Kalajzic, I., Gronowicz, G.,
Clark, S. H., Rowe, D. W. and Lichtler, A. C. (2001) Bone-directed
expression of Col1a1 promoter-driven self-inactivating retroviral
vector in bone marrow cells and transgenic mice. Mol. Ther., 3: 543-
50.
Sun, F. L. and Elgin, S. C. (1999) Putting boundaries on silence. Cell, 99:
459-62.
Tanaka, A. S., Tanaka, M. and Komuro, K. (1998) A highly efficient
method for the site-specific integration of transfected plasmids into the
genome of mammalian cells using purified retroviral integrase. Gene,
216: 67-76.
Thomas, C. E., Ehrhardt, A. and Kay, M. A. (2003) Progress and problems
with the use of viral vectors for gene therapy. Nat. Rev. Genet., 4:
346-58.
Ting, Y. T., Wilson, C. A., Farrell, K. B., Chaudry, G. J. and Eiden, M. V.
(1998) Simian sarcoma-associated virus fails to infect Chinese
hamster cells despite the presence of functional gibbon ape leukemia
virus receptors. J. Virol., 72: 9453-8.
Uch, R., Gerolami, R., Faivre, J., Hardwigsen, J., Mathieu, S., Mannoni, P.
and Bagnis, C. (2003) Hepatoma cell-specific ganciclovir-mediated
toxicity of a lentivirally transduced HSV-TkEGFP fusion protein gene
placed under the control of rat alpha-fetoprotein gene regulatory
sequences. Cancer Gene Ther., 10: 689-95.
Vile, R., Miller, N., Chernajovsky, Y. and Hart, I. (1994) A comparison of
the properties of different retroviral vectors containing the murine
tyrosinase promoter to achieve transcriptionally targeted expression of
the HSVtk or IL-2 genes. Gene Ther., 1: 307-16.
Walther, W., Stein, U., Fichtner, I., Alexander, M., Shoemaker, R. H. and
Schlag, P. M. (2000) Mdr1 promoter-driven tumor necrosis factor-
alpha expression for a chemotherapy-controllable combined in vivo
gene therapy and chemotherapy of tumors. Cancer Gene Ther., 7:
893-900.
Walther, W., Wendt, J. and Stein, U. (1997) Employment of the mdr1
promoter for the chemotherapy-inducible expression of therapeutic
genes in cancer gene therapy. Gene Ther., 4: 544-52.
Wei, S. Q., Mizuuchi, K. and Craigie, R. (1997) A large nucleoprotein
assembly at the ends of the viral DNA mediates retroviral DNA
integration. EMBO J., 16: 7511-20.
White, J. M. (1992) Membrane fusion. Science, 258: 917-24.
Wu, X., Li, Y., Crise, B. and Burgess, S. M. (2003) Transcription start
regions in the human genome are favored targets for MLV
integration. Science, 300: 1749-51.
Xu, K., Ma, H., McCown, T. J., Verma, I. M. and Kafri, T. (2001)
Generation of a stable cell line producing high-titer self-inactivating
lentiviral vectors. Mol. Ther., 3: 97-104.
Yamada, Y., Warren, A. J., Dobson, C., Forster, A., Pannell, R. and
Rabbitts, T. H. (1998) The T cell leukemia LIM protein Lmo2 is
necessary for adult mouse hematopoiesis. Proc. Natl. Acad. Sci. USA,
95: 3890-5.
Yannaki, E., Tubb, J., Aker, M., Stamatoyannopoulos, G. and Emery, D.
W. (2002) Topological constraints governing the use of the chicken
HS4 chromatin insulator in oncoretrovirus vectors. Mol. Ther., 5: 589-
98.
Yu, S. F., von Ruden, T., Kantoff, P. W., Garber, C., Seiberg, M., Ruther,
U., Anderson, W. F., Wagner, E. F. and Gilboa, E. (1986) Self-
inactivating retroviral vectors designed for transfer of whole genes
into mammalian cells. Proc. Natl. Acad. Sci. USA, 83: 3194-8.
Yu, S. S., Kim, J. M. and Kim, S. (2000) High efficiency retroviral vectors
that contain no viral coding sequences. Gene Ther., 7: 797-804.
Zufferey, R., Dull, T., Mandel, R. J., Bukovsky, A., Quiroz, D., Naldini, L.
and Trono, D. (1998) Self-inactivating lentivirus vector for safe and
efficient in vivo gene delivery. J. Virol., 72: 9873-80.
... The most often used gene therapy LVs have been developed based on HIV-1 (human immunode ciency virus type I). LV belongs to the retroviral family but can effectively infects both dividing and non-dividing cells (32)(33)(34). The major challenges for iv LV gene delivery include exposure to the host immune/complement system and offtarget gene transfer. ...
... LV pseudotyped with VSV-G envelope can infect both dividing and non-dividing cells at high e ciencies (32)(33)(34). However, major challenges exist for iv LV gene delivery due to the inhibitory host immune/complement system and off-target gene transfer. ...
Preprint
Full-text available
Chronic granulomatous disease (CGD) is a congenital immunodeficiency characterized by lack of reactive oxygen species in phagocytes. We developed an in vivo gene therapy strategy based on intravenous (iv) injection of lentiviral vectors (LVs) in X-CGD mice. A non-myeloablative chemo-conditioning regimen using busulfan, cyclophosphamide and dexamethasone was developed to improve iv LV gene delivery efficiency. The X-CGD mice received two LVs injections. After the second injection, antibody response to LV particle-associated p24-protein was examined by Western blot. We detected increased gene transfer without anti-p24 antibody response. However, the blood vector copy number (VCN) was gradually reduced after 3–12 months. To improve gene delivery into hematopoietic stem cells (HSCs), the mice were treated with AMD3100 to mobilize HSCs before LV injection. To confirm HSCs gene transfer, we transplanted the HSCs from the LV- CYBB -treated CGD mice into untreated CGD mice. The result showed successful passage of LV- CYBB HSCs to recipient mice. Thus, by combining chemo-conditioning and AMD3100 mobilization prior to the iv LV injection, improved in vivo long-term LV gene transfer into HSCs could be established. This improved iv LV gene delivery strategy could reduce both the risk and the cost of CGD gene therapy with great potential in translational applications.
... 50 However, several major drawbacks limit their applications. 51 First, retrovirus vectors require cell division to integrate its DNA into host genome, and thus they can only transduce dividing cells. In addition, retrovirus vectors have the risk to randomly insert its DNA into host chromosome and leads to insertional mutagenesis. ...
... In addition, retrovirus vectors have the risk to randomly insert its DNA into host chromosome and leads to insertional mutagenesis. 51 Self-inactivating vectors that have the promotor or enhancer of the long terminal repeat deleted have been developed to reduce the risk of insertional mutagenesis. 16 Due to these limitations, retrovirus vectors are currently not often used in clinical studies any more. ...
Article
Full-text available
Gene therapies are currently one of the most investigated therapeutic modalities in both the preclinical and clinical settings and have shown promise in treating a diverse spectrum of diseases. Gene therapies aim at introducing a gene material in target cells and represent a promising approach to cure diseases that were thought to be incurable by conventional modalities. In many cases, a gene therapy requires a vector to deliver gene therapeutics into target cells; viral vectors are among the most widely studied vectors owing to their distinguished advantages such as outstanding transduction efficiency. With decades of development, viral vector‐based gene therapies have achieved promising clinical outcomes with many products approved for treating a range of diseases including cancer, infectious diseases and monogenic diseases. In addition, a number of active clinical trials are underway to further expand their therapeutic potential. In this review, we highlight the diversity of viral vectors, review approved products, and discuss the current clinical landscape of in vivo viral vector‐based gene therapies. We have reviewed 13 approved products and their clinical applications. We have also analyzed more than 200 active trials based on various viral vectors and discussed their respective therapeutic applications. Moreover, we provide a critical analysis of the major translational challenges for in vivo viral vector‐based gene therapies and discuss possible strategies to address the same.
... Designing a safe and effective transfection vector is the key to gene therapy. Although viral vectors are currently considered to be most effective, safety concerns have limited their clinical application, making non-viral vectors the focus of attention of clinicians and researchers [109,110]. ...
Article
Full-text available
Hepatocellular carcinoma (HCC) has the sixth-highest new incidence and fourth-highest mortality worldwide. Transarterial chemoembolization (TACE) is one of the primary treatment strategies for unresectable HCC. However, the therapeutic effect is still unsatisfactory due to the insufficient distribution of antineoplastic drugs in tumor tissues and the worsened post-embolization tumor microenvironment (TME, e.g., hypoxia and reduced pH). Recently, using nanomaterials as a drug delivery platform for TACE therapy of HCC has been a research hotspot. With the development of nanotechnology, multifunctional nanoplatforms have been developed to embolize the tumor vasculature, creating conditions for improving the distribution and bioavailability of drugs in tumor tissues. Currently, the researchers are focusing on functionalizing nanomaterials to achieve high drug loading efficacy, thorough vascular embolization, tumor targeting, controlled sustained release of drugs, and real-time imaging in the TACE process to facilitate precise embolization and enable therapeutic procedures follow-up imaging of tumor lesions. Herein, we summarized the recent advances and applications of functionalized nanomaterials based on TACE against HCC, believing that developing these functionalized nanoplatforms may be a promising approach for improving the TACE therapeutic effect of HCC.
... This group of viruses has lipid capsids, and each virus particle has two copies of single-stranded linear positive RNA and 7-11 kb in size [56]. Retroviruses are a large family of viruses that have prone their potential to be used for gene therapy applications. ...
Article
Full-text available
Genetics and molecular genetic techniques have changed many perspectives and paradigms in medicine. Using genetic methods, many diseases have been cured or alleviated. Gene therapy, in its simplest definition, is application of genetic materials and related techniques to treat various human diseases. Evaluation of the trends in the field of medicine and therapeutics clarifies that gene therapy has attracted a lot of attention due to its powerful potential to treat a number of diseases. There are various genetic materials that can be used in gene therapy such as DNA, single- and double-stranded RNA, siRNA and shRNA. The main gene editing techniques used for in vitro and in vivo gene modification are ZNF, TALEN and CRISPR-Cas9. The latter has increased hopes for more precise and efficient gene targeting as it requires two separate recognition sites which makes it more specific and can also cause rapid and sufficient cleavage within the target sequence. There must be carriers for delivering genes to the target tissue. The most commonly used carriers for this purpose are viral vectors such as adenoviruses, adeno-associated viruses and lentiviruses. Non-viral vectors consist of bacterial vectors, liposomes, dendrimers and nanoparticles. Graphical abstract
... The gene transfer to epidermal cells has been commonly performed through retroviral methods, which, despite its high efficiency, have disadvantages such as immunogenicity, oncogenicity, and development of mutagenic processes. 21,22 Consequently, the use of synthetic nonviral vectors has generated a great interest in protocols for clinical use in gene therapy. In particular, polyethylenimine (PEI) polyplexes have been widely used for nonviral transfection in mammalian cells due to their ability to protect nucleic acids against premature degradation, and thus facilitate their delivery through the plasma membranes. ...
Article
Full-text available
Inefficient autologous tissue recovery in skin wounds increases the susceptibility of patients to infections caused by multidrug resistant microorganisms, resulting in a high mortality rate. Genetic modification of skin cells has become an important field of study because it could lead to the construction of more functional skin grafts, through the overexpression of antimicrobial peptides that would prevent early contamination and infection with bacteria. In this study, we produce and evaluate Human Skin Equivalents (HSEs) containing transfected human primary fibroblasts and keratinocytes by polyplexes to express the antimicrobial peptide LL-37. The effect of LL-37 on the metabolic activity of normal HSEs was evaluated before the construction of the transfected HSEs, and the antimicrobial efficacy against Pseudomonas aeruginosa and Staphylococcus aureus was evaluated. Subsequently, the levels of LL-37 in the culture supernatants of transfected HSEs, as well as the local expression were determined. It was found that LL-37 treatment significantly promoted the cellular proliferation of HSEs. Furthermore, HSEs that express elevated levels of LL-37 were shown to possess histological characteristics close to the normal skin and display enhanced antimicrobial activity against S. aureusin vitro. These findings demonstrate that HSEs expressing LL-37 through non-viral modification of skin cells is a promising approach for the prevention of bacterial colonization in wounds.
... However, to obtain gene expression from DNA, it is necessary to deliver DNA into the nucleus of the target cell, which is a difficult hurdle. The use of viral vectors allows for efficient gene delivery owing to the inherent ability of the viruses; however, there are concerns about their safety [13,14]. Development of non-viral gene delivery is preferable due to the advantage of industrial production. ...
Article
Full-text available
Nucleic acid and genetic medicines are increasingly being developed, owing to their potential to treat a variety of intractable diseases. A comprehensive understanding of the in vivo fate of these agents is vital for the rational design, discovery, and fast and straightforward development of the drugs. In case of intravascular administration of nucleic acids and genetic medicines, interaction with blood components, especially plasma proteins, is unavoidable. However, on the flip side, such interaction can be utilized wisely to manipulate the pharmacokinetics of the agents. In other words, plasma protein binding can help in suppressing the elimination of nucleic acids from the blood stream and deliver naked oligonucleotides and gene carriers into target cells. To control the distribution of these agents in the body, the ligand conjugation method is widely applied. It is also important to understand intracellular localization. In this context, endocytosis pathway, endosomal escape, and nuclear transport should be considered and discussed. Encapsulated nucleic acids and genes must be dissociated from the carriers to exert their activity. In this review, we summarize the in vivo fate of nucleic acid and gene medicines and provide guidelines for the rational design of drugs.
Article
The treatment of patients with advanced-stage solid tumours typically involves a multimodality approach (including surgery, chemotherapy, radiotherapy, targeted therapy and/or immunotherapy), which is often ultimately ineffective. Nucleic acid-based drugs, either as monotherapies or in combination with standard-of-care therapies, are rapidly emerging as novel treatments capable of generating responses in otherwise refractory tumours. These therapies include those using viral vectors (also referred to as gene therapies), several of which have now been approved by regulatory agencies, and nanoparticles containing mRNAs and a range of other nucleotides. In this Review, we describe the development and clinical activity of viral and non-viral nucleic acid-based treatments, including their mechanisms of action, tolerability and available efficacy data from patients with solid tumours. We also describe the effects of the tumour microenvironment on drug delivery for both systemically administered and locally administered agents. Finally, we discuss important trends resulting from ongoing clinical trials and preclinical testing, and manufacturing and/or stability considerations that are expected to underpin the next generation of nucleic acid agents for patients with solid tumours.
Article
Full-text available
The mature lymphocyte population of a healthy individual has the remarkable ability to recognise an immense variety of antigens. Instead of encoding a unique gene for each potential antigen receptor, evolution has used gene rearrangements, also known as variable, diversity, and joining gene segment (V(D)J) recombination. This process is critical for lymphocyte development and relies on recombination-activating genes-1 (RAG1) and RAG2, here collectively referred to as RAG. RAG serves as powerful genome editing tools for lymphocytes and is strictly regulated to prevent dysregulation. However, in the case of dysregulation, RAG has been implicated in cases of cancer, autoimmunity and severe combined immunodeficiency (SCID). This review examines functional protein domains and motifs of RAG, describes advances in our understanding of the function and (dys)regulation of RAG, discuss new therapeutic options, such as gene therapy, for RAG deficiencies, and explore in vitro and in vivo methods for determining RAG activity and target specificity.
Article
The manipulation of an individual’s genetic information to treat a disease has revolutionized the biomedicine field. Despite the promise of gene therapy, this treatment can have long-term side-effects. Efforts in the field and recent discoveries have already led to several improvements, including efficient gene delivery and transfer, as well as in patient safety. Several studies to treat a wide range of pathologies—such as cancer or monogenic diseases— are currently being conducted. Here we provide a broad overview of methodologies available for gene therapy, placing a strong emphasis on treatments for central nervous system diseases. Finally, we give a perspective on current delivery strategies to treat such diseases, with a special focus on systems that use peptides as delivery vectors.
Article
The art and science of gene therapy has received much attention of late. The tragic death of 18-year-old Jesse Gelsinger, a volunteer in a Phase I clinical trial, has overshadowed the successful treatment of three children suffering from a rare but fatal immunological disease. In the light of the success and tragedy, it is timely to consider the challenges faced by gene therapy - a novel form of molecular medicine that may be poised to have an important impact on human health in the new millennium.
Article
Common themes are emerging from the study of viral, cell-cell, intracellular, and liposome fusion. Viral and cellular membrane fusion events are mediated by fusion proteins or fusion machines. Viral fusion proteins share important characteristics, notably a fusion peptide within a transmembrane-anchored polypeptide chain. At least one protein involved in a cell-cell fusion reaction resembles viral fusion proteins. Components of intracellular fusion machines are utilized in multiple membrane trafficking events and are conserved through evolution. Fusion pores develop during viral and intracellular fusion events suggesting similar mechanisms for many, if not all, fusion events.
Article
Current retroviral vectors based on murine leukemia virus (MuLV) are unable to efficiently transduce nondividing cells. Lentiviruses, such as the human immunodeficiency virus 1 (HIV-1) are efficient at transducing nondividing, growth-arrested, and post-mitotic cells, but due to complex safety considerations, they may have limited potential for human clinical gene transfer. For this reason, alternatives to MuLV and HIV-1 vectors need to be explored. In this paper, we have found that simian foamy virus vector (SFV-1) containing a CMV-LacZ expression cassette is able to efficiently transduce multiple cell types of various species that include epithelial, lymphoid, and hematopoietic-derived human cell lines and fibroblast cell lines of several species. Previously it was reported that foamy virus replication is cell cycle dependent (P. D. Bieniasz, R. A. Weiss, and M. O. McClure, 1995. J. Virol. 69, 7295–7299). However, others studies demonstrated nuclear import of viral DNA in arrested cells (A. Saibi, F. Puvion-Dutilleul, M. Schmid, J. Peries, and H. d. The 1997. J. Virol. 71, 1155–1161). Here, we show efficient LacZ transduction by SFV-1 vectors in several chemically arrested cell lines and terminally differentiated human neurons. SFV-1 vector can transduce cell lines arrested in G1/S phase of the cell cycle by aphidicolin treatment with similar efficiencies to that of dividing cells. The terminally differentiated human neural cell line, NT2N, was transduced with 30–50% efficiency, corroborating our data obtained with the arrested cell lines. To further examine whether the SFV-1 vector can efficiently deliver a gene into clinically important cells for gene therapy, we transduced primary human peripheral blood cells (PBLs) in the presence and absence of phytohemagglutanin (PHA) stimulation. We observed 81% transduction efficiency in non-stimulated PBLs and 87% in PHA-stimulated PBLs with vector infection carried out twice in 8 hours intervals at a multiplicity of infection of 1. Together, these data indicate that SFV-1 based retroviral vectors may provide a safe, efficient alternative to current onco- and lentiviral vectors for gene transfer in cells from a broad spectrum of lineages across species boundaries.
Article
Tumor necrosis factor-α (TNF-α) is currently being used in clinical trials for cancer treatment, but toxic side effects, due to systemic administration and high doses, are observed. Inducible expression of TNF may permit selective killing of tumour cells in gene therapy protocols without need for prolonged and/or high-level TNF expression. A conditional TNF expression vector has been constructed in which the coding sequences of human TNF have been placed under the transcriptional control of the glucocorticoid-regulated murine mammary tumour virus long terminal repeat (MMTV-LTR). Negligible levels of TNF expression, associated with no phenotypic alterations, are observed in cells transfected with MMTV-TNF vectors in the absence of glucocorticoid. Expression levels could be stimulated by the addition of the synthetic glucocorticoid dexamethasone. Increasing expression levels of TNF were associated with enhanced cytotoxicity. Our results suggest the potential use of inducible TNF systems for the treatment of tumours in gene therapy protocols.
Article
Transcriptional silencing of retroviruses poses a major obstacle to their use as gene therapy vectors. Silencing is most pronounced in stem cells which are desirable targets for therapeutic gene delivery. Many vector designs combat silencing through cis-modifications of retroviral vector sequences. These designs include mutations of known retroviral silencer elements, addition of positive regulatory elements and insulator elements to protect the transgene from negative position effects. Similar strategies are being applied to lentiviral vectors that readily infect non-dividing quiescent stem cells. Collectively these cis-modifications have significantly improved vector design but optimal expression may require additional intervention to escape completely the trans-factors that scan for foreign DNA, establish silencing in stem cells and maintain silencing in their progeny. Cytosine methylation of CpG sites was proposed to cause retroviral silencing over 20 years ago. However, several studies provide evidence that retrovirus silencing acts through methylase-independent mechanisms. We propose an alternative silencing mechanism initiated by a speculative stem cell-specific ‘somno-complex’. Further understanding of retroviral silencing mechanisms will facilitate better gene therapy vector design and raise new strategies to block transcriptional silencing in transduced stem cells. Copyright © 2001 John Wiley & Sons, Ltd.