ArticlePDF Available

Impaired Intracellular Trafficking of Adeno-Associated Virus Type 2 Vectors Limits Efficient Transduction of Murine Fibroblasts

American Society for Microbiology
Journal of Virology
Authors:

Abstract and Figures

Although adeno-associated virus type 2 (AAV) has gained attention as a potentially useful alternative to the more commonly used retrovirus- and adenovirus-based vectors for human gene therapy, efficient gene transfer and transgene expression by AAV vectors require that the following two obstacles be overcome. First, the target cell must express the receptor and the coreceptor for AAV infection, and second, the cell must allow for viral second-strand DNA synthesis. We now describe a third obstacle, impaired intracellular trafficking of AAV to the nucleus, which results in the lack of transgene expression in murine fibroblasts which do express the AAV receptor and the coreceptor and which are permissive for viral second-strand DNA synthesis. We document that AAV vectors bind efficiently and gain entry successfully into NIH 3T3 cells, but trafficking into the nucleus is significantly impaired in these cells. In contrast, viral trafficking to the nucleus in cells known to be efficiently transduced by AAV vectors is both rapid and efficient. The demonstration of yet another obstacle in AAV-mediated gene transfer has implications for the optimal use of these vectors in human gene therapy.
Content may be subject to copyright.
JOURNAL OF VIROLOGY,
0022-538X/00/$04.000
Jan. 2000, p. 992–996 Vol. 74, No. 2
Copyright © 2000, American Society for Microbiology. All Rights Reserved.
Impaired Intracellular Trafficking of Adeno-Associated Virus
Type 2 Vectors Limits Efficient Transduction of
Murine Fibroblasts
JONATHAN HANSEN,
1,2,3
KEYUN QING,
1,2,3
HYUNG-JOO KWON,
1,2,3
CATHRYN MAH,
4
AND ARUN SRIVASTAVA
1,2,3,5
*
Department of Microbiology and Immunology,
1
Walther Oncology Center,
2
and Division of Hematology/Oncology,
Department of Medicine,
5
Indiana University School of Medicine, and Walther Cancer Institute,
3
Indianapolis,
Indiana 46202, and Gene Therapy Center, University of Florida College of Medicine,
Gainesville, Florida 32610
4
Received 9 September 1999/Accepted 18 October 1999
Although adeno-associated virus type 2 (AAV) has gained attention as a potentially useful alternative to the
more commonly used retrovirus- and adenovirus-based vectors for human gene therapy, efficient gene transfer
and transgene expression by AAV vectors require that the following two obstacles be overcome. First, the target
cell must express the receptor and the coreceptor for AAV infection, and second, the cell must allow for viral
second-strand DNA synthesis. We now describe a third obstacle, impaired intracellular trafficking of AAV to
the nucleus, which results in the lack of transgene expression in murine fibroblasts which do express the AAV
receptor and the coreceptor and which are permissive for viral second-strand DNA synthesis. We document
that AAV vectors bind efficiently and gain entry successfully into NIH 3T3 cells, but trafficking into the nucleus
is significantly impaired in these cells. In contrast, viral trafficking to the nucleus in cells known to be efficiently
transduced by AAV vectors is both rapid and efficient. The demonstration of yet another obstacle in AAV-
mediated gene transfer has implications for the optimal use of these vectors in human gene therapy.
Adeno-associated virus type 2 (AAV) is a nonpathogenic
human parvovirus which requires coinfection with a helper
virus, such as adenovirus, for its optimal replication (3). In the
absence of a helper virus, the wild-type AAV genome inte-
grates in a site-specific way into human chromosome 19 and
establishes a latent infection (14, 15, 27). AAV possesses a
wide host range that transcends the species barrier (18). These
properties of AAV have been instrumental in the development
of recombinant AAV vectors for their use in human gene
therapy (4–6, 18, 33–35, 40). We and others have reported
AAV-mediated transduction and transgene expression both in
vitro and in vivo (1, 2, 9, 10, 12, 13, 17, 19–26, 28, 30, 41).
However, the transduction efficiency of AAV vectors varies
greatly in different cell types. This problem has been attributed
to the following two obstacles that AAV must overcome. First,
AAV must bind to a cellular receptor as well as to a coreceptor
for successful entry into the target cell (19, 21, 24, 36, 37), and
second, because AAV is a single-stranded DNA-containing
virus, the target cell must allow for the conversion of the
single-stranded viral genome to a transcriptionally active dou-
ble-stranded intermediate (7, 8). We have documented the
existence of a host cell protein, which we have designated the
single-stranded D sequence-binding protein (ssD-BP), which
plays a crucial role in viral second-strand DNA synthesis (23,
25). The ssD-BP is phosphorylated at tyrosine residues by
epidermal growth factor receptor protein tyrosine kinase
(EGFR-PTK) activity, and the phosphorylated form of the
ssD-BP prevents viral second-strand DNA synthesis and con-
sequently AAV-mediated transgene expression (16). In this
report, we provide evidence for the existence of a third obsta-
cle, impaired intracellular trafficking into the nucleus, which
AAV must also overcome prior to high-efficiency transduction.
We recently documented that AAV binds to murine NIH
3T3 cells efficiently but that little transgene expression occurs
(24). This observation was interpreted as the inability of AAV
to enter these cells. However, Southern blot analyses (31) of
low-M
r
DNA isolated from NIH 3T3 cells soon after infection
with a recombinant AAV vector containing the human cyto-
megalovirus (CMV) immediate-early gene promoter-driven
bacterial -galactosidase (lacZ) reporter gene revealed the
presence of AAV single-stranded DNA. These results are
shown in Fig. 1A. In these experiments, equivalent numbers of
Raji (nonpermissive for AAV binding and entry) (24), 293
(permissive for viral binding and entry) (23), and NIH 3T3
cells were either mock infected or infected with 10
4
particles/
cell of cesium chloride density gradient-purified recombinant
vCMVp-lacZ vector for2hat37°C, following which cells were
treated with 0.05% trypsin and washed extensively with phos-
phate-buffered saline (PBS) to remove any virus particles ad-
sorbed to cellular receptors in the plasma membrane, and
low-M
r
DNA samples were isolated. Equivalent amounts were
analyzed on Southern blots with a
32
P-labeled DNA probe
specific for lacZ gene sequences as previously described (19,
24). These data suggest that AAV does indeed gain entry into
NIH 3T3 cells. The apparent lack of transgene expression in
NIH 3T3 cells noted previously (24) was not due to the inac-
tivity of the CMV promoter, since abundant expression of the
lacZ gene could be detected in plasmid DNA transfection
experiments (data not shown).
Although treatment of NIH 3T3 cells with tyrphostin 1, a
specific inhibitor of EGFR-PTK (16), led to dephosphorylation
of the ssD-BP as determined by electrophoretic mobility shift
assays (EMSAs), transgene expression in tyrphostin 1-treated
NIH 3T3 cells could not be detected by a cytochemical staining
* Corresponding author. Mailing address: Department of Microbi-
ology and Immunology, Indiana University School of Medicine, 635
Barnhill Dr., Medical Science Building, Room 257, Indianapolis, IN
46202-5120. Phone: (317) 274-2194. Fax: (317) 274-4090. E-mail:
asrivast@iupui.edu.
992
method (24). EMSAs were performed as previously described
(16, 25). Briefly, 10 g of each whole-cell extract was preincu-
bated with 2 g of poly(dI-dC), 2 g of bovine serum albumin
(BSA), and 12% glycerol in HEPES buffer (pH 7.9) for 10 min
at 25°C. Following preincubation, 10,000 cpm of
32
P-labeled
D() sequence synthetic oligonucleotide (5-AGGAACCCCT
AGTGATGGAG-3) was added to the reaction mixture and
incubated for 30 min at 25°C. The bound complexes were
separated from the free probe by electrophoresis on 4% poly-
acrylamide gels. The ratio of dephosphorylated to phosphory-
lated forms of ssD-BP was determined by densitometric
scanning of autoradiograms with a Digital Imaging System
Alphaimager (Alpha Innotech Corp., San Leandro, Calif.).
These results are shown in Fig. 1B. Tyrphostin 1 treatment of
NIH 3T3 cells resulted in a change of the ratio of dephospho-
rylated to phosphorylated forms of ssD-BP from 0.8 to 1.2.
Accordingly, when NIH 3T3 cells were infected with the re-
combinant AAV vector (5 10
3
particles/cell) by a more
sensitive assay than cytochemical staining, a low level of AAV-
mediated transgene expression could be detected 48 h later in
these cells treated with higher concentrations of tyrphostin 1.
In these experiments, -galactosidase activity was measured
with the Galacto-Light Plus chemiluminescent reporter assay
(Tropix, Inc., Bedford, Mass.) according to the manufacturer’s
instructions; results were within the linear range. These data,
expressed as relative light units (RLU) per microgram of total
protein, are shown in Fig. 1C. HeLa cells were used as a
positive control in these experiments, since tyrphostin 1 treat-
FIG. 1. (A) Southern blot analysis of viral DNA entry into cells. Equivalent numbers of Raji, 293, and NIH 3T3 cells were either mock infected or infected with
the recombinant vCMVp-lacZ vector (10
4
particles/cell), trypsinized, and washed extensively, and low-M
r
DNA samples were isolated and analyzed with the
32
P-labeled
lacZ DNA probe as described previously (19, 24). ssDNA, viral single-stranded DNA genomes. (B) EMSA for detection of the phosphorylation status of the ssD-BP
in NIH 3T3 cells. Equivalent numbers of cells were either mock treated or treated with the indicated amounts of tyrphostin 1 at 37°C for 2 h. Whole-cell extracts were
prepared and used in EMSAs with the
32
P-labeled single-stranded AAV DNA probe, as described previously (16, 25). Phosphorylated and dephosphorylated forms of
the ssD-BP are denoted by the arrow and the arrowhead, respectively. (C) AAV-mediated transgene expression. Equivalent numbers of HeLa (left) and NIH 3T3
(right) cells were infected with 5 10
3
particles/cell of vCMVp-lacZ and then treated with the indicated amounts of tyrphostin 1 for 2 h under identical conditions.
Forty-eight hours postinfection, -galactosidase activity was determined as described in the text.
VOL. 74, 2000 NOTES 993
ment has previously been shown to significantly augment
AAV-mediated transgene expression in these cells by decreas-
ing the ratio of phosphorylated to dephosphorylated forms of
ssD-BP (16). Whereas there was a dose-dependent increase in
AAV transduction efficiency in both cell types, the extent of
transgene expression was roughly 2 orders of magnitude lower
in NIH 3T3 cells. These results corroborate that AAV does
indeed enter the NIH 3T3 cells.
Entry of the virus into NIH 3T3 cells was further confirmed
with AAV that was fluorescently labeled with Cy3 (Amersham
Life Sciences, Pittsburgh, Pa.) as described previously (2). 293
and NIH 3T3 cells were plated onto polylysine-treated cover-
slips and 24 h later were infected at 37°C with 10
4
particles per
cell of Cy3-labeled AAV in medium containing 0.1% BSA. At
various time points, cells were washed three times with me-
dium containing 0.1% BSA and fixed at 4°C for 20 min with
PBS containing 2% formaldehyde and 0.2% glutaraldehyde.
Cellular nucleic acid was subsequently counterstained with 1
M Syto-16 (Molecular Probes, Inc., Eugene, Oreg.) for 30
min at 25°C. Cells were then washed three times with PBS,
mounted on glass slides, and visualized with a Zeiss LSM 510
confocal microscope. A series of 0.3-m optical sections were
made through the cells and images representative of the center
of the cells were compared to assess the localization of viral
particles. These results are shown in Fig. 2. It is evident that
within 15 min, AAV could bind efficiently to both 293 (Fig. 2,
panel 1) and NIH 3T3 (panel 3) cells. By 2 h, entry of AAV
into 293 (panel 2) and NIH 3T3 (panel 4) cells was also clearly
seen. As expected, M07e cells, known not to bind AAV due to
lack of expression of one of the coreceptors (16, 21, 24), failed
to bind the fluorescent-labeled virus (data not shown).
Since these data are more qualitative than quantitative and
FIG. 2. Confocal microscopy for localization of AAV particles in 293 and NIH 3T3 cells. At 15 min (panels 1 and 3) and 2 h (panels 2 and 4) postinfection with
10
4
particles/cell of Cy3-labeled AAV, 293 (panels 1 and 2) and NIH 3T3 (panels 3 and 4) cells were visualized as described in the text. Cy3-labeled AAV (red) and
cellular nucleic acids (green) are shown in images representative of the center of the cells. Magnification, 630.
994 NOTES J. VIROL.
do not take into account the kinetics of intracellular trafficking
and uncoating of AAV in the two cell types being compared,
we hypothesized that despite successful entry, a significant
fraction of AAV vectors fails to enter the nucleus in NIH 3T3
cells. This hypothesis was experimentally tested as follows.
Southern blot analyses of low-M
r
DNA isolated at various
times postinfection were carried out with cytoplasmic as well as
nuclear fractions isolated from 293 and NIH 3T3 cells infected
under identical conditions with the CMVp-lacZ vector (10
4
particles/cell) as described above. Nuclear and cytoplasmic
fractions were prepared as described previously (32) from
equivalent numbers of cells, followed by isolation of low-M
r
DNA. These results are shown in Fig. 3A. It is evident that
within 2 h, a substantial amount (76%) of the input single-
stranded AAV DNA is present in the nucleus in 293 cells,
whereas essentially all (99%) of the signal is detected in the
cytoplasm in NIH 3T3 cells, as determined by densitometric
scanning of autoradiographs. By 48 h, a small portion (18%)
of the input AAV DNA does enter the nucleus in NIH 3T3
cells, but the majority (82%) of the signal is still in the
cytoplasm. In contrast, 78% of the input viral DNA is in the
nucleus, and 22% of the signal is in the cytoplasm of 293 cells
48 h postinfection. The purity of each fraction was determined
to be 95%, as measured by the absence of acid phosphatase
activity (in the nuclear fraction) and the absence of histone H3
(in the cytoplasmic fraction), by Western blot analysis with
H3 antibody (Upstate Biotechnology, Lake Placid, N.Y.)
(data not shown).
We have previously reported that while AAV binds to HeLa
and KB cells more efficiently than to 293 cells, AAV-mediated
transduction of HeLa and KB cells is significantly lower than
that of 293 cells (23). Therefore, we compared the efficiency of
AAV trafficking into the nucleus of these three cell types.
These results are shown in Fig. 3B. It is clear that despite
less-efficient binding, the extent of AAV entry as well as traf-
ficking into the nucleus in 293 cells is significantly higher than
that in HeLa and KB cells. Thus, in addition to the phosphor-
ylation status of the cellular ssD-BP (23), AAV transduction
efficiency among permissive human cells also correlates well
with the extent of viral trafficking into the nucleus. Taken
together, these studies establish that efficient translocation to
the nucleus is essential for successful transduction of cells by
AAV vectors.
It is interesting to note that although AAV transduction
efficiency of murine cells in general has been reported to be
low (18), the exceptions include the muscle and the brain
tissues (10, 12, 17, 41). We have previously suggested that this
might be due to overabundant expression of fibroblast growth
factor receptor 1, a coreceptor for AAV infection (24), and the
lack of expression of the EGFR, PTK activity of which cata-
lyzes the phosphorylation of the ssD-BP, resulting in failure to
synthesize the viral second-strand DNA (16). It would be of
interest to now examine the kinetics of AAV trafficking into
the nucleus in primary murine cells as well as human cells that
are transduced differentially by AAV vectors, to substantiate
the observations reported here. Indeed, our recent studies
suggest that impaired intracellular trafficking of AAV into the
nucleus in primary murine hematopoietic progenitor cells lim-
its high-efficiency transduction of these cells, both in vitro and
in vivo (39).
Virtually nothing is known about the intracellular trafficking
of AAV particles following infection. However, a wealth of
information is available on the underlying mechanisms of cy-
toplasmic transport and nuclear import of other viruses (11).
For instance, herpesvirus binds to the cellular protein, dynein,
a minus-end-directed motor protein, which transports the viral
particle along microtubules toward the nucleus where the viral
DNA is released through the nuclear pore complex into the
nucleus (29). Adenovirus, on the other hand, first enters the
endosomal pathway and, after acidification of the endosome, is
released into the cytoplasm where it appears to bind microtu-
bules prior to nuclear entry (38). Since both herpesviruses and
adenoviruses provide the helper function for a productive in-
fection by AAV (3, 18) and since both adenovirus and AAV
use V5 integrin as a coreceptor (36), it is reasonable to
suggest that trafficking of AAV into the nucleus might be
accomplished by similar mechanisms as those employed by its
helper viruses. Thus, further studies on the mechanisms of
postreceptor entry, transport into the nucleus, and uncoating
of AAV should allow a clearer understanding of molecular
events involved in AAV-mediated high-efficiency transduction
which, in turn, should lead to improvements in the optimal use
of AAV vectors in human gene therapy.
We thank Hal E. Broxmeyer for a critical review of the manuscript
as well as for his support. We also thank Johnny He for his helpful
suggestions.
This research was supported in part by Public Health Service grants
(HL-53586, HL-58881, and DK-49218; Centers of Excellence in Mo-
lecular Hematology) from the National Institutes of Health and a
grant from the Phi Beta Psi Sorority.
REFERENCES
1. Alexander, I. E., D. W. Russell, A. M. Spence, and A. D. Miller. 1996. Effects
of gamma irradiation on the transduction of dividing and nondividing cells in
brain and muscle of rats by recombinant adeno-associated virus vectors.
Hum. Gene Ther. 7:841–850.
2. Bartlett, J. S., R. J. Samulski, and T. J. McCown. 1998. Selective and rapid
uptake of adeno-associated virus type 2 in brain. Hum. Gene Ther. 9:1181–
1186.
3. Berns, K. I., and R. A. Bohenzky. 1987. Adeno-associated viruses: an update.
Adv. Virus Res. 32:243–307.
4. Berns, K. I., and C. Giraud. 1996. Biology of adeno-associated virus. Curr.
Top. Microbiol. Immunol. 218:1–23.
5. Carter, B. J., and T. R. Flotte. 1996. Development of adeno-associated virus
vectors for gene therapy of cystic fibrosis. Curr. Top. Microbiol. Immunol.
218:119–144.
6. Chatterjee, S., D. Lu, G. Podsakoff, and K. K. Wong, Jr. 1995. Strategies for
efficient gene transfer into hematopoietic cells: the use of adeno-associated
FIG. 3. (A) Southern blot analysis of subcellular distribution of the viral
DNA in 293 and NIH 3T3 cells. Approximately 2 10
6
cells of each type were
either mock infected or infected with the recombinant vCMVp-lacZ vector (10
4
particles/cell), trypsinized, and washed extensively. Low-M
r
DNA samples were
isolated at various indicated time points from purified nuclear (Nuc.) and cyto-
plasmic (Cyt.) fractions and analyzed with the
32
P-labeled lacZ DNA probe as
described in the legend to Fig. 1. (B) Southern blot analyses of AAV entry and
trafficking into the nucleus of permissive human cell lines. Equivalent numbers of
HeLa, KB, and 293 cells were infected with the recombinant vCMVp-lacZ
vector, trypsinized, and washed extensively, and low-M
r
DNA samples were
isolated 2 h postinfection from purified nuclear and cytoplasmic fractions. The
rest of the analysis on Southern blots was performed as described above.
VOL. 74, 2000 NOTES 995
virus vectors in gene therapy. Ann. N. Y. Acad. Sci. 770:79–90.
7. Ferrari, F. K., T. Samulski, T. Shenk, and R. J. Samulski. 1996. Second-
strand synthesis is a rate-limiting step for efficient transduction by recombi-
nant adeno-associated virus vectors. J. Virol. 70:3227–3234.
8. Fisher, K. J., G.-P. Gao, M. D. Weitzman, R. DeMatteo, J. F. Burda, and
J. M. Wilson. 1996. Transduction with recombinant adeno-associated virus
for gene therapy is limited by leading-strand synthesis. J. Virol. 70:520–532.
9. Herzog, R. W., E. Y. Yang, L. B. Couto, J. N. Hagstrom, D. Elwell, P. A.
Fields, M. Burton, D. A. Bellinger, M. S. Read, K. M. Brinkhous, G. M.
Podsakoff, T. C. Nichols, G. J. Kurtzman, and K. A. High. 1999. Long-term
correction of canine hemophilia B by gene transfer of blood coagulation
factor IX mediated by adeno-associated viral vector. Nat. Med. 5:56–63.
10. Kaplitt, M. G., P. Leone, R. J. Samulski, X. Xiao, D. W. Pfaff, K. L. O’Malley,
and M. J. During. 1994. Long-term gene expression and phenotypic correc-
tion using adeno-associated virus vectors in the mammalian brain. Nat.
Genet. 8:148–153.
11. Kasamatsu, H., and A. Nakanishi. 1998. How do animal viruses get to the
nucleus? Annu. Rev. Microbiol. 52:627–686.
12. Kessler, P. D., G. M. Podsakoff, X. Chen, S. A. McQuiston, P. C. Colosi, L. A.
Matelis, G. J. Kurtzman, and B. J. Byrne. 1996. Gene delivery to skeletal
muscle results in sustained expression and systemic delivery of a therapeutic
protein. Proc. Natl. Acad. Sci. USA 93:14082–14087.
13. Koeberl, D. D., I. E. Alexander, C. L. Halbert, D. W. Russell, and A. D.
Miller. 1997. Persistent expression of human clotting factor IX from mouse
liver after intravenous injection of adeno-associated virus vectors. Proc. Natl.
Acad. Sci. USA 94:1426–1431.
14. Kotin, R. M., J. C. Menninger, D. C. Ward, and K. I. Berns. 1991. Mapping
and direct visualization of a region-specific viral DNA integration site on
chromosome 19q13-qter. Genomics 10:831–834.
15. Kotin, R. M., M. Siniscalco, R. J. Samulski, X. D. Zhu, L. A. Hunter, C. A.
Laughlin, S. K. McLaughlin, N. Muzyczka, M. Rocchi, and K. I. Berns. 1990.
Site-specific integration by adeno-associated virus. Proc. Natl. Acad. Sci.
USA 87:2211–2215.
16. Mah, C., K. Y. Qing, B. Khuntirat, S. Ponnazhagan, X.-S. Wang, D. M.
Kube, M. C. Yoder, and A. Srivastava. 1998. Adeno-associated virus 2-me-
diated gene transfer: role of epidermal growth factor receptor protein ty-
rosine kinase in transgene expression. J. Virol. 72:9835–9843.
17. McCown, T. J., X. Xiao, J. Li, G. R. Breese, and R. J. Samulski. 1996.
Differential and persistent expression patterns of CNS gene transfer by an
adeno-associated virus (AAV) vector. Brain Res. 713:99–107.
18. Muzyczka, N. 1992. Use of adeno-associated virus as a general transduction
vector for mammalian cells. Curr. Top. Microbiol. Immunol. 158:97–129.
19. Ponnazhagan, S., P. Mukherjee, X.-S. Wang, K. Y. Qing, D. M. Kube, C.
Mah, C. Kurpad, M. C. Yoder, E. F. Srour, and A. Srivastava. 1997. Adeno-
associated virus type 2-mediated transduction of primary human bone mar-
row-derived CD34
hematopoietic progenitor cells: donor variation and
correlation of transgene expression with cellular differentiation. J. Virol.
71:8262–8267.
20. Ponnazhagan, S., P. Mukherjee, M. C. Yoder, X.-S. Wang, S. Z. Zhou, J.
Kaplan, S. Wadsworth, and A. Srivastava. 1997. Adeno-associated virus
2-mediated gene transfer in vivo: organ-tropism and expression of trans-
duced sequences in mice. Gene 190:203–210.
21. Ponnazhagan, S., X.-S. Wang, M. J. Woody, F. Luo, L. Y. Kang, M. L.
Nallari, N. C. Munshi, S. Z. Zhou, and A. Srivastava. 1996. Differential
expression in human cells from the p6 promoter of human parvovirus B19
following plasmid transfection and recombinant adeno-associated virus 2
(AAV) infection: human megakaryocytic leukaemia cells are non-permissive
for AAV infection. J. Gen. Virol. 77:1111–1122.
22. Ponnazhagan, S., M. C. Yoder, and A. Srivastava. 1997. Adeno-associated
virus type 2-mediated transduction of murine hematopoietic cells with long-
term repopulating ability and sustained expression of a human globin gene in
vivo. J. Virol. 71:3098–3104.
23. Qing, K. Y., B. Khuntirat, C. Mah, D. M. Kube, X.-S. Wang, S. Ponnazha-
gan, S. Z. Zhou, V. J. Dwarki, M. C. Yoder, and A. Srivastava. 1998.
Adeno-associated virus type 2-mediated gene transfer: correlation of ty-
rosine phosphorylation of the cellular single-stranded D sequence-binding
protein with transgene expression in human cells in vitro and murine tissues
in vivo. J. Virol. 72:1593–1599.
24. Qing, K. Y., C. Mah, J. Hansen, S. Z. Zhou, V. J. Dwarki, and A. Srivastava.
1999. Human fibroblast growth factor receptor 1 is a co-receptor for infec-
tion by adeno-associated virus 2. Nat. Med. 5:71–77.
25. Qing, K. Y., X.-S. Wang, D. M. Kube, S. Ponnazhagan, A. Bajpai, and A.
Srivastava. 1997. Role of tyrosine phosphorylation of a cellular protein in
adeno-associated virus 2-mediated transgene expression. Proc. Natl. Acad.
Sci. USA 94:10879–10884.
26. Rivera, V. M., X. Ye, N. L. Courage, J. Sachar, F. Cerasoli, J. M. Wilson, and
M. Gilman. 1999. Long-term regulated expression of growth hormone in
mice after intramuscular gene transfer. Proc. Natl. Acad. Sci. USA 96:8657–
8662.
27. Samulski, R. J., X. Zhu, X. Xiao, J. D. Brook, D. E. Houseman, N. Epstein,
and L. A. Hunter. 1991. Targeted integration of adeno-associated virus
(AAV) into human chromosome 19. EMBO J. 10:3941–3950.
28. Snyder, R. O., C. Miao, L. Meuse, J. Tubb, B. A. Donahue, H.-F. Lin, D. W.
Stafford, S. Patel, A. R. Thompson, T. C. Nichols, M. S. Read, D. A. Bell-
inger, K. M. Brinkhous, and M. A. Kay. 1999. Correction of hemophilia B in
canine and murine models using recombinant adeno-associated viral vectors.
Nat. Med. 5:64–70.
29. Sodeik, B., M. Ebersold, and A. Helenius. 1997. Dynein mediated transport
of incoming herpes simplex virus 1 capsids to the nucleus. J. Cell Biol.
136:1007–1021.
30. Song, S., M. Morgan, T. Ellis, A. Poirier, K. Chesnut, J. Wang, M. Brantley,
N. Muzyczka, B. J. Byrne, M. Atkinson, and T. R. Flotte. 1998. Sustained
secretion of human alpha-1-antitrypsin from murine muscle transduced with
adeno-associated virus vectors. Proc. Natl. Acad. Sci. USA 95:14384–14388.
31. Southern, E. M. 1975. Detection of specific sequences among DNA frag-
ments separated by gel electrophoresis. J. Mol. Biol. 98:503–517.
32. Sperinde, V., and M. A. Nugent. 1998. Heparan sulfate proteoglycans control
intracellular processing of bFGF in vascular smooth muscle cells. Biochem-
istry 37:13153–13164.
33. Srivastava, A. 1998. Delivery systems for gene therapy: adeno-associated
virus 2, p. 257–288. In P. J. Quesenberry, G. S. Stein, B. G. Forget, and S. M.
Weissman (ed.), Stem cell biology and gene therapy. Wiley-Liss, Inc., New
York, N.Y.
34. Srivastava, A. 1999. Parvoviral vectors for human hematopoietic gene ther-
apy, p. 89–122. In L. J. Fairbairn and N. Testa (ed.), Blood cell biochemistry,
vol. 8. Hematopoiesis and gene therapy. Kluwer Academic/Plenum Publish-
ers, New York, N.Y.
35. Srivastava, A., X.-S. Wang, S. Ponnazhagan, S. Z. Zhou, and M. C. Yoder.
1996. Adeno-associated virus 2-mediated transduction and erythroid lineage-
specific expression in human hematopoietic progenitor cells. Curr. Top.
Microbiol. Immunol. 218:93–117.
36. Summerford, C., J. S. Bartlett, and R. J. Samulski. 1999. V5 integrin: a
co-receptor for adeno-associated virus 2 infection. Nat. Med. 5:78–82.
37. Summerford, C., and R. J. Samulski. 1998. Membrane-associated heparan
sulfate proteoglycan is a receptor for adeno-associated virus type 2 virions.
J. Virol. 72:1438–1445.
38. Suomalainen, M., M. Y. Nakano, S. Keller, K. Boucke, R. P. Stidwell, and
U. F. Greber. 1999. Microtubule-dependent plus- and minus end-directed
motilities are competing processes for nuclear targeting of adenovirus.
J. Cell Biol. 144:657–672.
39. Tan, M. Q., K. Y. Qing, M. C. Yoder, and A. Srivastava. Adeno-associated
virus 2-mediated transduction of primary murine hematopoietic progenitor
cell in vitro and in vivo. Blood, in press.
40. Walsh, C. E., J. M. Liu, J. L. Miller, A. W. Nienhuis, and R. J. Samulski.
1993. Gene therapy for human hemoglobinopathies. Proc. Soc. Exp. Biol.
Med. 204:289–300.
41. Xiao, X., L. Li, and R. J. Samulski. 1996. Efficient long-term gene transfer
into muscle tissue of immunocompetent mice by adeno-associated virus
vector. J. Virol. 70:8098–8108.
996 NOTES J. VIROL.
... For instance, treatment with brefeldin A [74,76] or Golgicide A [74] disrupts the Golgi apparatus structure and abolishes AAV2 transduction. In NIH3T3 cells AAV2 successfully binds to the cells and is efficiently endocytosed, but transduction is very low [77]. Interestingly, it has been shown that intracellular trafficking of AAV2 in NIH3T3 cells differs from HeLa cells [77], and that in NIH3T3 cells the transport of AAV2 to the Golgi is severely impaired [74]. ...
... In NIH3T3 cells AAV2 successfully binds to the cells and is efficiently endocytosed, but transduction is very low [77]. Interestingly, it has been shown that intracellular trafficking of AAV2 in NIH3T3 cells differs from HeLa cells [77], and that in NIH3T3 cells the transport of AAV2 to the Golgi is severely impaired [74]. ...
Article
In the last two decades, recombinant adeno-associated virus has emerged as the most popular gene therapy vector. Recently AAV gene therapy has been approved by the FDA for the treatment of two rare genetic disorders, namely the early childhood blindness disease Leber congenital amaurosis and spinal muscular atrophy (SMA). As is the case for the treatment of SMA, if the AAV vector must be administered systemically, very high vector doses are often required for therapeutic efficacy. But higher vector doses inevitably increase the risk of adverse events. The tragic death of three children in a clinical trial to treat X-linked myotubular myopathy with an AAV vector has thrown this limitation into sharp relief. Regardless of the precise cause(s) that led to the death of the two children, it is critical that we develop better AAV vectors to achieve therapeutic levels of expression with lower vector doses. To transduce successfully a target cell, AAV has to overcome both systemic as well as cellular roadblocks. In this review, we discuss some of the most prominent cellular roadblocks that AAV must get past to deliver successfully its therapeutic payload. We also highlight recent advancements in our knowledge of AAV biology that can potentially be harnessed to improve AAV vector performance and thereby make AAV gene therapy safer.
... The expression of a dominant negative version of RAB5 leads to a 70% and 60% decrease in transduction of rAAV2 and rAAV8, respectively. Similarly, rAAV2 and 8 colocalize with EE antigen 1 (EEA1), an EE-specific marker [42,43]. EEs communicate with the trans-Golgi network (TGN) in a bidirectional vesicular exchange [44] and rAAV particles need to travel to the Golgi for efficient transduction [45,46]. ...
... The observation that golgicide A, which inhibits the transport of Shiga toxin from EEs to the TGN [47], also inhibits rAAV transduction demonstrates the importance of this exchange. In less-permissive cells such as NIH3T3 and murine fibroblasts, minimal accumulation of rAAV in the TGN is observed [43,48]. All of these observations demonstrate a central role of the Golgi in rAAV transduction. ...
Article
As adeno-associated virus (AAV)-based gene therapies are being increasingly approved for use in humans, it is important that we understand vector–host interactions in detail. With the advances in genome-wide genetic screening tools, a clear picture of AAV–host interactions is beginning to emerge. Understanding these interactions can provide insights into the viral life cycle. Accordingly, novel strategies to circumvent the current limitations of AAV-based vectors may be explored. Here, we summarize our current understanding of the various stages in the journey of the vector from the cell surface to the nucleus and contextualize the roles of recently identified host factors.
... Endosome escape, capsid degradation, intracellular transport, and nuclear translocation are significant hurdles. Only 20% of all AAV2 virions that infected fibroblasts were able to deliver their genetic payload to the nucleus and express the transgene (Hansen et al., 2000). This difference between virion uptake and 'functional transduction' resulting in transgene expression is AAV subtype, target cell and species dependent (Lisowski et al., 2014;Westhaus et al., 2020). ...
Article
Full-text available
Central Nervous System (CNS) homeostasis and function rely on intercellular synchronization of metabolic pathways. Developmental and neurochemical imbalances arising from mutations are frequently associated with devastating and often intractable neurological dysfunction. In the absence of pharmacological treatment options, but with knowledge of the genetic cause underlying the pathophysiology, gene therapy holds promise for disease control. Consideration of leukodystrophies provide a case in point; we review cell type - specific expression pattern of the disease - causing genes and reflect on genetic and cellular treatment approaches including ex vivo hematopoietic stem cell gene therapies and in vivo approaches using Adeno-associated virus (AAV) vectors. We link recent advances in vectorology to glial targeting directed towards gene therapies for specific leukodystrophies and related developmental or neurometabolic disorders affecting the CNS white matter and frame strategies for therapy development in future.
... However, essential viral processes post-entry can limit the ability of a virus to cause clinical disease in a host. Within-host barriers to pathogenicity following cell entry can include the ability to evade host immune defenses (vesicular stomatitis virus [12]; rinderpest virus [13]), the ability to make viral proteins and replicate in the cell (influenza A [14]; deformed wing virus, [15]), intracellular trafficking (adeno-associated virus [16]), and the ability to exit the cell and spread to new hosts (influenza A [17]; H5N1 avian influenza virus [18]). Intrinsic differences between host species affecting any of these processes, in either or both lymphatic and epithelial cells, may need to be overcome to induce a disease state, i.e., distemper, in a novel host. ...
Article
Full-text available
Canine distemper virus (CDV) is a multi-host pathogen with variable clinical outcomes of infection across and within species. We used whole-genome sequencing (WGS) to search for viral markers correlated with clinical distemper in African lions. To identify candidate markers, we first documented single-nucleotide polymorphisms (SNPs) differentiating CDV strains associated with different clinical outcomes in lions in East Africa. We then conducted evolutionary analyses on WGS from all global CDV lineages to identify loci subject to selection. SNPs that both differentiated East African strains and were under selection were mapped to a phylogenetic tree representing global CDV diversity to assess if candidate markers correlated with documented outbreaks of clinical distemper in lions (n = 3). Of 54 SNPs differentiating East African strains, ten were under positive or episodic diversifying selection and 20 occurred in the clinical strain despite strong purifying selection at those loci. Candidate markers were in functional domains of the RNP complex (n = 19), the matrix protein (n = 4), on CDV glycoproteins (n = 5), and on the V protein (n = 1). We found mutations at two loci in common between sequences from three CDV outbreaks of clinical distemper in African lions; one in the signaling lymphocytic activation molecule receptor (SLAM)-binding region of the hemagglutinin protein and another in the catalytic center of phosphodiester bond formation on the large polymerase protein. These results suggest convergent evolution at these sites may have a functional role in clinical distemper outbreaks in African lions and uncover potential novel barriers to pathogenicity in this species.
... In other studies, comparison of various viral vectors in rat, rabbit and human MSCs showed that different serotypes of AAV vectors were the least effective in accomplishing gene delivery [55,56]. Finally, AAV2-mediated gene delivery in fibroblasts has been associated with~30-fold lower levels of AAV2 DNA replication when compared with 293T or HeLa cells [57,58]. ...
Article
Full-text available
Adeno-associated viral vectors (AAV) are unique in their ability to transduce a variety of both dividing and nondividing cells, with significantly lower risk of random genomic integration and with no known pathogenicity in humans, but their role in ex vivo regional gene therapy for bone repair has not been definitively established. The goal of this study was to test the ability of AAV vectors carrying the cDNA for BMP-2 to transduce human mesenchymal stem cells (MSCs), produce BMP-2, and induce osteogenesis in vitro as compared with lentiviral gene therapy with a two-step transcriptional amplification system lentiviral vector (LV-TSTA). To this end, we created two AAV vectors (serotypes 2 and 6) expressing the target transgene; eGFP or BMP-2. Transduction of human MSCs isolated from bone marrow (BMSCs) or adipose tissue (ASCs) with AAV2-eGFP and AAV6-eGFP led to low transduction efficiency (BMSCs: 3.57% and 8.82%, respectively, ASCs: 6.17 and 20.2%, respectively) and mean fluorescence intensity as seen with FACS analysis 7 days following transduction, even at MOIs as high as 106. In contrast, strong eGFP expression was detectable in all of the cell types post transduction with LV-TSTA-eGFP. Transduction with BMP-2 producing vectors led to minimal BMP-2 production in AAV-transduced cells 2 and 7 days following transduction. In addition, transduction of ASCs and BMSCs with AAV2-BMP-2 and AAV6-BMP-2 did not enhance their osteogenic potential as seen with an alizarin red assay. In contrast, the LV-TSTA-BMP-2-transduced cells were characterized by an abundant BMP-2 production and induction of the osteogenic phenotype in vitro (p < 0.001 vs. AAV2 and 6). Our results demonstrate that the AAV2 and AAV6 vectors cannot induce a significant transgene expression in human BMSCs and ASCs, even at MOIs as high as 106. The LV-TSTA vector is significantly superior in transducing human MSCs; thus this vector would be preferable when developing an ex vivo regional gene therapy strategy for clinical use in orthopedic surgery applications.
... However, drug-induced inhibition of clathrin-coated vesicles showed an independency of AAV2 of this particular route, and alternative pathways such as the GPI-anchored-protein-enriched endosomal compartment as well as the clathrin-independent carriers were suggested instead 114 . After internalization, the AAV particle has to traffic to the nucleus, which was demonstrated to be a rate-limiting hurdle for the outcome of the infection 115 . ...
Thesis
Full-text available
In recent years, the adeno-associated virus (AAV) gained considerable attention mainly due to the approval of the first AAV-based gene therapy treatment in the Western hemisphere in 2012, named Glybera®. It not only conveyed the feasibility of utilizing this parvovirus to introduce healthy gene copies but simultaneously reinforced further interest in developing more specific and efficient synthetic vectors by capsid engineering approaches such as DNA family shuffling or random peptide display. However, the characterization of lead candidates resulting from these directed evolution strategies is labor-intensive and therefore excludes the possibility to validate multiple promising variants. Therefore, a comprehensive high-throughput capsid validation pipeline was established in this work adapting a previously reported approach in which a DNA barcode-comprising AAV genome is assigned to a chosen capsid variant during virus production. Thus, the identification of the respective capsid in the complex physiological environment of living animals is enabled by solely detecting the barcode sequence via next generation sequencing. The principle was further improved by placing the barcode into the 3’UTR of a CMV promoter-driven eyfp transgene permitting tracking on the DNA and RNA level. Hence, next to information about transduction efficiency, the especially crucial transcriptional activity in a certain tissue was measured. Using this design, three barcoded AAV libraries were generated comprising up to 157 variants including 12 commonly used serotypes, >70 peptide-displaying mutants based on these naturally occurring wild types and several published benchmarks such as AAVDJ, AAV9_PHP.B and AAVAnc80L65. After intravenously injecting the library into C57BL/6J mice and analyzing the RNA and DNA data from >20 collected tissues, prior observations for the literature variants could be confirmed thus validating the workflow. Most impressively, a peptide display mutant previously created in our laboratory exhibited drastically improved efficiencies in the diaphragm, heart and skeletal muscles in comparison to AAV9wt on the cDNA and protein level while in addition demonstrating pronounced muscle specificity. In conclusion, in the course of this PhD thesis a highly robust barcode-based capsid screening pipeline was established that facilitates and accelerates the identification of promising candidates for gene therapies, best exemplified by the discovery of the muscle-tropism of our lead candidate.
Preprint
Full-text available
Krabbe disease is caused by the mutation or deficiency of galactocerebrosidase (GALC) enzyme, which is located in the lysosome and hydrolyzes the galactolipid substrates like psychosine. Psychosine would accumulate abnormally in the myelin forming cells and result in demyelination in the nervous systems with the clinical symptoms of spastic paraparesis and seizures. Adeno-associated virus (AAV) is a well-established and safe viral vector for gene delivery. However, effective AAV serotype for the transduction of the human neural stem cells (NSCs) has not been identified. Here, we screened a variety of AAV serotypes to transduce NSCs-related disease model induced by Krabbe patient induced pluripotent stem cells (iPSC) differentiation. It has been found that AAV2 has a higher transduction effect for NSCs, and AAV2 carrying GALC gene rescued the GALC enzymatic activity of Krabbe NSCs. Our findings established Krabbe patient iPSCs-derived NSCs as a new model for study the pathogenesis of Krabbe disease, and also demonstrated the potential of using AAV2 as a vector in gene therapy for Krabbe disease, which also proved the potential of AAV in devising gene therapy strategies for the treatment of genetic neurodegenerative diseases.
Article
Aim: Adeno-associated virus (AAV) is the most preferred gene therapy vector. The purpose of our research is to compare the infection tropism and gene expression efficiency of vitreous injection of recombinant AAVs (rAAVs) and their capsid mutants in mouse retina. Materials & methods: We packaged wild-type rAAV2/2,6,8,9 and their capsid mutants carrying EGFP expression cassette using insect cells. The gene expression profiles of rAAVs and their mutants in mouse retina were evaluated by optical imaging of retinal tissue flat mount and cryosections. Results & conclusion: The results showed that rAAV2 and rAAV2-Y444F mainly targeted retinal ganglion cell; rAAV8, rAAV8-Y733F, rAAV9 and mutants had obvious EGFP expression in retinal pigment epithelium cells. Compared with the wild-type rAAVs, capsid mutants have an improved transduction efficiency in mouse retina cells.
Article
Efforts to identify mutations that underlie inherited genetic diseases combined with strides in the development of gene therapy vectors over the last three decades have culminated in the approval of several adeno-associated virus (AAV)-based gene therapies. Genetic diseases that manifest in the lung such as cystic fibrosis and surfactant deficiencies, however, have so far proven to be elusive targets. Early clinical trials in CF using AAV serotype 2 (AAV2) achieved safety but not efficacy endpoints, but importantly these studies provided critical information on barriers that need to be surmounted to translate AAV lung gene therapy towards clinical success. Bolstered with an improved understanding of AAV biology and more clinically relevant lung models, next generation molecular biology and bioinformatics approaches have given rise to novel AAV capsid variants that offer improvements in transduction efficiency, immunological profile, and the ability to circumvent physical barriers in the lung such as mucus. This review discusses the principal limiting barriers to clinical success in lung gene therapy and focuses on novel engineered AAV capsid variants that have been developed to overcome those challenges.
Article
Recombinant vectors based on a nonpathogenic parvovirus, the adeno-associated virus (AAV), have taken center stage in the past decade. The safety of AAV vectors in clinical trials and clinical efficacy in several human diseases are now well documented. Despite these achievements, it is increasingly clear that the full potential of AAV vectors composed of the naturally occurring capsids is unlikely to be realized. This article describes advances that have been made and challenges that remain in the optimal use of AAV vectors in human gene therapy applications.
Article
Full-text available
Cellular sequences flanking integrated copies of the adeno-associated virus (AAV) genome were isolated from a latently infected clonal human cell line and used to probe genomic blots derived from an additional 21 independently derived clones of human cells latently infected with AAV. In genomic blots of uninfected human cell lines and of primary human tissue, each flanking-sequence probe hybridized to unique bands, but in 15 of the 22 latently infected clones the flanking sequences hybridized not only to the original fragments but also to a total of 36 additional species. AAV probes also hybridized to 22 of these new bands, representing 11 of the 15 positive clones, but never to the fragment characteristic of uninfected cell DNA. From these data we conclude that the AAV genome preferentially integrates into a specific region of the cellular genome. We have determined that the integration site is unique to chromosome 19 by somatic cell hybrid mapping, and this sequence has been isolated from uninfected human DNA.
Article
Full-text available
Expression from the human parvovirus B19p6 promoter fused to the firefly luciferase ('Luc') reporter gene was evaluated in a non-erythroid human nasopharyngeal carcinoma cell line, KB, and a human megakaryocytic leukaemia cell line, MB-02, known to become permissive for B19 replication following erythroid-differentiation. The B19p6-Luc construct was introduced into KB and MB-02 cells, both in undifferentiated and differentiated states, either via DNA-mediated transfection, or via infection with recombinant adeno-associated virus 2 (AAV), a non-pathogenic human parvovirus known to possess a broad host-range. Although Luc activity was readily detected in KB cells following transfection of the B19p6-Luc plasmid DNA, no expression from the B19p6 promoter was observed following infection with recombinant virus. In addition, transfection of the reporter plasmid resulted in high-level expression of Luc in differentiated but not in undifferentiated MB-02 cells. However, no Luc activity was detected, even in differentiated MB-02 cells, following infection with recombinant virus. Further studies with an additional recombinant as well as wild-type (wt) AAV revealed that MB-02 cells were non-permissive for AAV infection. A second human megakaryocytic leukaemia cell line, M07e, was likewise resistant to infection by recombinant as well as wt AAV. Taken together, these studies identify the first human cell type that cannot be infected by AAV. They indicate that expression from the B19p6 promoter, in the context of an AAV genome, is restricted to primary human haematopoietic cells, perhaps because parvoviral DNA replication and transcription are intrinsically coupled.
Chapter
The concept of treating human diseases by introducing normal alleles of genes into appropriate target cells is a clinical milestone (Anderson, 1995). Although a number of physical and chemical methods for gene transfer have been developed, viruses have generally proven much more efficient in transferring genetic material into cells. Indeed, viral vectors based on retroviruses and adenoviruses have already been employed in a number of clinical trials (Crystal, et al., 1994; Rosenberg et al., 1990; Zabner et al., 1993). Although initial results with retroviral vectors have been encouraging (Blaese et al., 1995; Bordignon et al., 1995; Grossman et al., 1994) their use in nonhuman primate studies have been reported to lead to malignancy (Donahue et al., 1992), and the efficacy of adenoviral vectors has been questioned (Knowles et al., 1995), Because retroviruses and most of the DNA-containing viruses are the etiologic agents of, or are intimately associated with, malignant disorders (Tooze, 1981; Weiss et al., 1984), the search for an alternative viral vector continues (Hodgson, 1995; Miller and Vile, 1995). Parvoviruses, which are among the smallest of the DNA-containing viruses that infect a wide variety of vertebrates (Siegl et al., 1985), remain the only group of viruses that have thus far not been associated with malignant disease (Cotmore and Tattersall, 1987; Pattison, 1988). In fact, parvoviruses possess antitumor properties (Cukor et al., 1975; Mayor et al., 1973; Ostrove et al., 1981). One of the members of the Parvoviridae family, the adeno-associated virus 2 (AAV), has gained significant attention primarily because AAV is a nonpathogenic human parvovirus (Blacklow, 1988), and the wild-type (wt) AAV genome establishes a latent infection in human cells where the viral genome integrates into the chromosomal DNA in a site-specific manner (Kotin et al., 1989, 1990, 1991, 1992; Samulski et al., 1991). Although several reviews on AAV vectors have been published (Berns and Linden, 1995; Carter, 1993; Flotte and Carter, 1995; Kotin, 1994; Muzyczka, 1992; Samulski, 1993; Srivastava, 1994; Srivastava et al., 1996; Xiao et al., 1993), this article highlights salient features of parvoviruses in general and outlines a number of advantages that support the potential use of parvovirus vectors in human gene therapy.
Article
Guidelines for submitting commentsPolicy: Comments that contribute to the discussion of the article will be posted within approximately three business days. We do not accept anonymous comments. Please include your email address; the address will not be displayed in the posted comment. Cell Press Editors will screen the comments to ensure that they are relevant and appropriate but comments will not be edited. The ultimate decision on publication of an online comment is at the Editors' discretion. Formatting: Please include a title for the comment and your affiliation. Note that symbols (e.g. Greek letters) may not transmit properly in this form due to potential software compatibility issues. Please spell out the words in place of the symbols (e.g. replace “α” with “alpha”). Comments should be no more than 8,000 characters (including spaces ) in length. References may be included when necessary but should be kept to a minimum. Be careful if copying and pasting from a Word document. Smart quotes can cause problems in the form. If you experience difficulties, please convert to a plain text file and then copy and paste into the form.
Article
This chapter attempts to highlight the salient features of AAV in general and outline a number of advantages of AAV vectors for their potential use in human gene therapy.
Article
Adeno-associated virus 2 (AAV), a non-pathogenic human parvovirus, is gaining attention as a vector for its potential use in human gene therapy. However, few studies have examined the safety and the efficacy of this vector system in vivo. We report here that recombinant AAV vectors, when directly injected intravenously in mice, accumulated predominantly in liver cells, suggesting that AAV may possess in vivo organ-tropism for liver. The transduced lacZ reporter gene was expressed in hepatocytes in the liver and, at the level examined, did not appear to induce any detectable cytotoxic T lymphocyte response against βGal. AAV-mediated transduction of murine hematopoietic progenitor cells ex vivo followed by transplantation into lethally irradiated syngeneic mice also revealed high-efficiency gene transfer into progeny cells without any observable cytotoxicity or deleterious effect. The transduced reporter gene sequences were also expressed in mice in vivo. The AAV-based vectors may thus prove useful as a potentially safe alternative to the more commonly used retrovirus- and adenovirus-based vector systems.
Article
Adeno-associated virus is an integrating DNA parvovirus with the potential to be an important vehicle for somatic gene therapy. A potential barrier, however, is the low transduction efficiencies of recombinant adeno-associated virus (rAAV) vectors. We show in this report that adenovirus dramatically enhances rAAV transduction in vitro in a way that is dependent on expression of early region 1 and 4 (E1 and E4, respectively) genes and directly proportional to the appearance of double-stranded replicative forms of the rAAV genome. Expression of the open reading frame 6 protein from E4 in the absence of E1 accomplished a similar but attenuated effect. The helper activity of adenovirus E1 and E4 for rAAV gene transfer was similarly demonstrated in vivo by using murine models of liver- and lung-directed gene therapy. Our data indicate that conversion of a single-stranded rAAV genome to a duplex intermediate limits transduction and usefulness for gene therapy.
Article
This paper describes a method of transferring fragments of DNA from agarose gels to cellulose nitrate filters. The fragments can then be hybridized to radioactive RNA and hybrids detected by radioautography or fluorography. The method is illustrated by analyses of restriction fragments complementary to ribosomal RNAs from Escherichia coli and Xenopus laevis, and from several mammals.
Article
Parvoviruses are among the smallest of the DNA-containing viruses that infect a wide variety of vertebrates (Siegl et al. 1985). Two parvoviruses of human origin, the nonpathogenic adeno-associated virus 2 (AAV) and the parvovirus B19, a common human pathogen, have been studied extensively (Berns and Bohenzky 1987; Brown et al. 1994). AAV requires coinfection with a helper virus, such as adenovirus or herpesvirus, for its optimal replication (Berns 1990), but in the absence of a helper virus, the AAV genome establishes a latent infection in a site-specific manner (Kotin and Berns 1989; Kotin et al. 1990, 1991, 1992; Samulski et al. 1991). B19, by contrast, is an autonomously replicating virus with a remarkable tropism for human erythroid progenitor cells (Ozawa et al. 1986, 1987; Yaegashi et al. 1989; Srivastava and Lu 1988; Takahashi et al. 1990). We have described the construction of a recombinant AAV-B19 hybrid genome, in which we combined the remarkable features of these two parvoviruses, and speculated that such a hybrid vector may prove useful for high efficiency transduction of primary human hematopoietic progenitor cells (Srivastava et al. 1989). Indeed, it has become increasingly clear that the AAV-based vector system may prove to be a useful alternative to the more commonly used retroviral and adenoviral vectors for its potential use in human gene therapy (Muzyczka 1992; Carter 1993; Srivastava 1994). Despite these advances, a number of fundamental questions related to AAV remain unanswered. For example, the molecular details of viral assembly and the mechanism of viral entry into the host cell have not been rigorously analyzed. Furthermore, the feasibility of obtaining tissue-specific expression of an AAV-transduced gene has not been adequately addressed. Here, we provide experimental evidence to suggest that the vector assembly requires a precise signaling mechanism and that AAV infection of human cells is receptor-mediated. We also document erythroid lineage restricted expression following AAV-B19 hybrid vector-mediated transduction of primary human hematopoietic progenitor cells. Elucidation of the molecular details of these aspects of AAV biology will have important implications in the potential use of AAV as a vector in human gene therapy.