ArticlePDF Available

Expression of Inducible Nitric Oxide Synthase and Elevation of Tyrosine Nitration of a 32-Kilodalton Cellular Protein in Brain Capillary Endothelial Cells from Rats Infected with a Neuropathogenic Murine Leukemia Virus

American Society for Microbiology
Journal of Virology
Authors:

Abstract and Figures

PVC-211 murine leukemia virus (MuLV) is a neuropathogenic variant of Friend MuLV (F-MuLV) which causes a rapidly progressive spongiform neurodegenerative disease in rodents. The primary target of PVC-211 MuLV infection in the brain is the brain capillary endothelial cell (BCEC), which is resistant to F-MuLV infection. Previous studies have shown that changes in the envelope gene of PVC-211 MuLV confer BCEC tropism to the virus. However, little is known about how infection of BCECs by PVC-211 MuLV induces neurological disease. Previous results suggest that nitric oxide (NO), which has been implicated as a potential neurotoxin, is involved in PVC-211 MuLV-induced neurodegeneration. In this study, we show that expression of inducible nitric oxide synthase (iNOS), which produces NO from L-arginine, is induced in BCECs from PVC-211 MuLV-infected rats. Furthermore, elevated levels of a 32-kDa cellular protein modified by 3-nitrotyrosine, which is a hallmark of NO production, were observed in virus-infected BCECs. BCECs from rats infected with BCEC-tropic but nonneuropathogenic PVF-e5 MuLV, which is a chimeric virus between PVC-211 MuLV and F-MuLV, fail to induce either iNOS expression or elevation of tyrosine nitration of a 32-kDa protein. These results suggest that expression of iNOS and nitration of tyrosine residues of a 32-kDa protein in PVC-211 MuLV-infected BCECs may play an important role in neurological disease induction.
Content may be subject to copyright.
JOURNAL OF VIROLOGY, May 2003, p. 5145–5151 Vol. 77, No. 9
0022-538X/03/$08.000 DOI: 10.1128/JVI.77.9.5145–5151.2003
Copyright © 2003, American Society for Microbiology. All Rights Reserved.
Expression of Inducible Nitric Oxide Synthase and Elevation of
Tyrosine Nitration of a 32-Kilodalton Cellular Protein in Brain
Capillary Endothelial Cells from Rats Infected with a
Neuropathogenic Murine Leukemia Virus
Atsushi Jinno-Oue,
1
Susan G. Wilt,
2
Charlotte Hanson,
1
Natalie V. Dugger,
2
Paul M. Hoffman,
2
Michiaki Masuda,
3
and Sandra K. Ruscetti
1
*
Basic Research Laboratory, National Cancer Institute, Frederick, Maryland 21702
1
; Research Service, Department of Veterans
Affairs Medical Center and Department of Neurology, University of Maryland, Baltimore, Maryland 21201
2
; and
Department of Microbiology, School of Medicine, Dokkyo University, Tochigi 321-0293, Japan
3
Received 7 October 2002/Accepted 4 February 2003
PVC-211 murine leukemia virus (MuLV) is a neuropathogenic variant of Friend MuLV (F-MuLV) which
causes a rapidly progressive spongiform neurodegenerative disease in rodents. The primary target of PVC-211
MuLV infection in the brain is the brain capillary endothelial cell (BCEC), which is resistant to F-MuLV
infection. Previous studies have shown that changes in the envelope gene of PVC-211 MuLV confer BCEC
tropism to the virus. However, little is known about how infection of BCECs by PVC-211 MuLV induces
neurological disease. Previous results suggest that nitric oxide (NO), which has been implicated as a potential
neurotoxin, is involved in PVC-211 MuLV-induced neurodegeneration. In this study, we show that expression
of inducible nitric oxide synthase (iNOS), which produces NO from
L-arginine, is induced in BCECs from
PVC-211 MuLV-infected rats. Furthermore, elevated levels of a 32-kDa cellular protein modified by 3-nitro-
tyrosine, which is a hallmark of NO production, were observed in virus-infected BCECs. BCECs from rats
infected with BCEC-tropic but nonneuropathogenic PVF-e5 MuLV, which is a chimeric virus between PVC-211
MuLV and F-MuLV, fail to induce either iNOS expression or elevation of tyrosine nitration of a 32-kDa
protein. These results suggest that expression of iNOS and nitration of tyrosine residues of a 32-kDa protein
in PVC-211 MuLV-infected BCECs may play an important role in neurological disease induction.
A number of murine leukemia viruses (MuLVs) have been
shown to induce diseases of the central nervous system (CNS)
that are characterized by progressive loss of neuronal function
(35, 39). The major cell types within the CNS that are prom-
inently infected with the MuLVs are glial and endothelial cells,
with neurons being infrequently infected. The most commonly
observed pathological changes are gliosis, neuronal loss, and
demyelination. The mechanism(s) by which the MuLVs induce
neurological diseases remains to be elucidated.
PVC-211 MuLV is a neuropathogenic variant of the leuke-
mia-inducing Friend MuLV (F-MuLV) (21). Infection of sus-
ceptible rats with PVC-211 MuLV causes a rapidly progressive
neurodegenerative disease characterized by tremor, spasticity,
ataxia, and hind limb paralysis. Neuropathological changes in-
clude widespread perivascular gliosis, neuropil vacuolation
without inflammation, and neuronal degeneration in the brain
stem, cerebellum, and spinal cord (19, 28).
The primary target of PVC-211 MuLV infection in the CNS
is the brain capillary endothelial cell (BCEC), which is resis-
tant to F-MuLV infection (19). The determinant of the BCEC
tropism of PVC-211 MuLV was mapped to two amino acids
(G
116
and K
129
) which lie within the putative receptor binding
domain of the envelope surface glycoprotein (SU) (30). Within
the CNS, reactive astrocytes and degenerating neurons showed
no evidence of virus infection (19). BCEC tropism of the virus
has been shown to be necessary for neuropathogenesis (29),
suggesting that CNS injury is indirect and that molecular
events in virus-infected BCECs play a very important role in
neurological disease induction.
Nitric oxide (NO) is an important messenger and effector
molecule involved in a number of biological functions (31). NO
is synthesized from
L-arginine by three isoforms of NO syn-
thases (NOS). Endothelial cell NOS (eNOS) and neuronal
NOS are constitutively expressed, and their activities are reg-
ulated by Ca
2
. In contrast, inducible NOS (iNOS) is inducible
and Ca
2
independent (13). In the CNS, NO may play impor
-
tant roles in neurotransmitter release, neurotransmitter re-
uptake, neurodevelopment, synaptic plasticity, and regulation
of gene expression, although excessive production of NO can
lead to neurotoxicity (9, 27). iNOS is an attractive candidate
for mediating NO-associated neurotoxicities, because long
bursts of large amounts of NO are produced by iNOS (7, 9, 32).
Indeed, elevated iNOS expression has been demonstrated in
such human neurological diseases as Alzheimer’s disease (24)
and Parkinson’s disease (26).
Recently, the spongiform vacuolation observed in PVC-211
MuLV-infected brains was reported to be associated with ox-
idative damage as detected by increased immunoreactivity for
* Corresponding author. Mailing address: Building 469, Room 205,
National Cancer Institute at Frederick, Frederick, MD 21702-1201.
Phone: (301) 846-5740. Fax: (301) 846-6164. E-mail: ruscetti@ncifcrf
.gov.
† Present address: Department of Virology and Preventive Medi-
cine, Gunma University School of Medicine, Gunma 371-8511, Japan.
‡ Present address:R&DProgram 151, Department of Veterans
Affairs Medical Center, Bronx, NY 10468.
5145
3-nitrotyrosine (NTyr) in infected brains (43). NTyr is widely
used as an indicator of NO formation, because nitration of
tyrosine is mediated by reactive nitrogen species derived from
NO (2, 11, 15). Elevated expression of NTyr has also been
reported in human neurodegenerative diseases such as familial
amyotrophic lateral sclerosis (41), Alzheimers disease (17),
Parkinsons disease (12), and human immunodeciency virus
type 1 dementia complex (5). In this study, we examined ex-
pression of iNOS and elevated expression of NTyr in PVC-211
MuLV-infected BCECs to evaluate the contribution of NO
produced by infected BCECs to the neuropathogenicity in-
duced by PVC-211 MuLV infection.
MATERIALS AND METHODS
Viruses. Neuropathogenic PVC-211 and PVF-e5 MuLV, a nonneuropatho-
genic variant of PVC-211, were grown in NIH 3T3 cells as described previously
(30). The viral supernatants had titers of 10
5
to 10
6
PFU/ml as determined by an
XC assay (40). Virus samples were stored at 80°C until use.
Animals. Pregnant Fisher 344 (F344) rats were obtained from Charles River
(Raleigh, N.C.) and housed in the Small Animal Facility at the Department of
Veterans Affairs Medical Center (Baltimore, Md.). All experiments were per-
formed in accordance with Public Health Service guidelines, using an IACUC-
approved protocol (P. M. Hoffman). Two-day-old F344 rats were inoculated
intracerebrally with 0.03 ml of supernatant from virus-producing NIH 3T3 cells.
Cells. Primary rat BCECs were isolated from the brains of virus- or medium-
inoculated 3-week-old F344 rats as described previously (6, 19) and grown for 2
weeks in minimum essential medium (MEM) with
D-valine (Life Technologies,
Inc., Gaithersburg, Md.) supplemented with 20% fetal calf serum, 2 mM
L-
glutamine, 50 U of penicillin/ml, 50 g of streptomycin/ml, 1 mM MEM nones-
sential amino acids solution (Life Technologies), 1 mM vitamin solution (Life
Technologies), 50 g of endothelial mitogen (Biomedical Technology Inc.,
Stoughton, Mass.)/ml, and 16 U of heparin (Life Technologies)/ml at 37°Cina
5% CO
2
humidied incubator for 2 weeks. Isolated BCEC populations were
95% pure as determined by immunohistochemistry with the endothelial cell
marker factor VIII.
Immunohistochemistry on brain sections. Brain tissue was collected from
medium- or virus-inoculated rats following intracardiac perfusion. The tissues
were then xed, and serial sections were stained for expression of iNOS, eNOS,
and MuLV SU (gp70). Briey, sections were rst incubated for 48 h with mouse
anti-iNOS (610329 from BD Transduction Laboratories, Lexington, Ky.); mouse
anti-eNOS (610296 from BD Transduction Laboratories); or goat anti-Rauscher
MuLV (R-MuLV) gp70 (National Cancer Institute, Bethesda, Md.). After wash-
ing in Tris-buffered saline, sections were incubated for 1 h with the appropriate
biotinylated secondary antibody (Southern Biotechnology, Birmingham, Ala.).
Antibody complexes were visualized using the Vectastain Elite ABC kit (Vector
Laboratories, Burlingame, Calif.). Sections were then dehydrated and cover-
slipped with Permount (Sigma, St. Louis, Mo.). Three to six sections per rat brain
were analyzed. Images were collected using a Nikon Eclipse 600 equipped with
a Nikon DMX 1200 digital camera.
RT- PCR analysis. To examine expression of NOS isoforms, we obtained total
cellular RNA from BCECs isolated from the brains of rats inoculated with
medium, PVC-211 MuLV, or PVF-e5 MuLV by using the RNA-STAT60 reagent
(TEL-TEST, Inc., Friendswood, Tex.). RNA from BCECs stimulated with lipo-
polysaccharide (LPS) was used as a positive control for iNOS expression. RNA
(2 g) was reverse transcribed in the presence of 50 mM random hexamers using
200 U of Superscript II reverse transcriptase (RT; Invitrogen, Carlsbad, Calif.)
according to the protocol supplied by the manufacturer. PCR was performed to
amplify iNOS, eNOS, or -actin sequences. Specic primer pairs used for each
amplication were as follows: iNOS sense, 5-ATGGAACAGTATAAGGCAA
ACACC-3; iNOS antisense, 5-GTTTCTGGTCGATGTCATGAGCAAAGG-
3; eNOS sense, 5-TACGGAGCAGCAAATCCAC-3; eNOS antisense, 5-GA
TCAAAGGACTGCAGCCTG-3; -actin sense, 5-CGTAAAGACCTCTATG
CCAA-3; and -actin antisense, 5-AGCCATGCCAAATGTCTCAT-3.
Each amplication reaction mixture contained 100 ng of cDNA, 400 ng of each
specic primer, a 0.2 mM concentration of each deoxyribonucleoside triphos-
phate, 2.5 U of Taq polymerase (Invitrogen), and 10 reaction buffer. Ampli-
cation conditions were as follows: 95°C for 5 min for 1 cycle; 30 cycles of 95°C for
1 min, 55°C for 1 min, and 72°C for 1 min; and 72°C for 10 min for 1 cycle.
Western blot analysis. To prepare cell lysates, cells were washed with cold
phosphate-buffered saline and lysed with immunoprecipitation buffer (50 mM
Tris-HCl [pH 7.5], 150 mM NaCl, 1% Triton X-100, 1% sodium deoxycholate,
0.1% sodium dodecyl sulfate [SDS], 10 mM EDTA, 10 g of aprotinin/ml, and
1 mM phenylmethylsulfonyl uoride). The protein samples were prepared in
buffer containing 625 mM Tris-HCl (pH 6.8), 2% SDS, 10% glycerol, and 50 mM
dithiothreitol (SDS-sample buffer) and boiled for 5 min. Then, SDS-polyacryl-
amide gel electrophoresis (SDS-PAGE) was performed, and protein samples
were electrophoretically transferred to nitrocellulose membranes (Invitrogen).
The membranes were then incubated with goat antibody to R-MuLV gp70 or p30
(National Cancer Institute); rabbit antibody to iNOS (Santa Cruz Biotechnology,
Santa Cruz, Calif.); rabbit antibody to NTyr (Upstate Biotechnology, Lake
Placid, N.Y.); or mouse antibody to -tubulin (Sigma). Bound antibodies were
detected with peroxidase-labeled secondary antibodies (DAKO, Carpinteria,
Calif.) by using the enhanced chemiluminescence system (Amersham Pharmacia
Biotech, Piscataway, N.J.). Blocking experiments for NTyr immunoreactivity
were performed by preincubating (25°C, 30 min) the anti-NTyr antibody with 10
mM NTyr (Sigma) prior to incubation with the membrane. For reprobing the
membrane, antibodies bound to the membrane were removed by incubating with
buffer containing 62.5 mM Tris-HCl (pH 6.7), 2% SDS, and 100 mM 2-mercap-
toethanol at 55°C for 30 min. After washing the membrane, subsequent proce-
dures for the binding of antibodies were carried out as described above. A cell
lysate from BCECs stimulated with LPS was used as a positive control for iNOS
expression. Stimulation was performed by incubation with 100 ng of LPS/ml for
24 h.
Detection of iNOS enzymatic activity. iNOS enzymatic activity was measured
by production of
L-[
3
H]citrulline from L-[
3
H]arginine (14) with the use of the
NOSdetect kit (Stratagene, La Jolla, Calif.) according to the manufacturers
specications. The reaction was performed in the presence of EGTA (1 mM) to
determine the calcium-independent activity of the inducible enzyme. Boiled
samples were used to determine the background level.
Inoculation of PVC-211 MuLV to primary BCECs in vitro. One day before
inoculation (day 0), primary rat BCECs (passage 2) were seeded at 10
4
per well
into 24-well tissue culture plates (Costar, Cambridge, Mass.). The following day
(day 1), PVC-211 MuLV (multiplicity of infection 10) was incubated with cells
in the presence of Polybrene (5 g/ml) for1hat37°C. Then the cells were
washed with medium, and fresh culture medium was added. This inoculation step
was performed once a day and completed by day 4. At day 14, cell extracts were
prepared and analyzed for expression of viral protein and iNOS by Western
blotting as described above.
RESULTS
Specic expression of iNOS in brains and BCECs from
PVC-211 MuLV-infected rat brain. To evaluate the molecular
events in PVC-211 MuLV-infected brains that may be contrib-
uting to neuropathogenicity, we examined brains from virus-
infected rats for expression of iNOS. As a control, we used
brains from rats infected with PVF-e5 MuLV, a chimeric virus
between F-MuLV and PVC-211 MuLV that efciently infects
BCECs but fails to induce neurological disease (29, 43). We
rst used immunohistochemistry to examine brain sections for
expression of iNOS. As shown in Fig. 1, iNOS could be de-
tected only in brains from PVC-211 MuLV-infected rats, and
the protein was expressed predominantly in BCECs (Fig. 1A).
Although only one time point is shown (14 days postinfection),
iNOS immunoreactivity was also observed at 21 and 28 days
after PVC-211 MuLV inoculation (data not shown) and was
not restricted to specic brain regions. Roughly 20 to 30% of
all BCECs in microvessels exhibited iNOS immunoreactivity,
compared with 90 to 100% of all BCECs in microvessels that
express the viral SU gp70 glycoprotein (Fig. 1G). Other cell
populations known to be activated following PVC-211 MuLV
infection, either microglia (43) or astrocytes (19), failed to
express iNOS immunoreactivity. In contrast to brains from
PVC-211 MuLV-infected rats, iNOS immunoreactivity could
not be detected in brains from rats infected with the nonneu-
5146 JINNO-OUE ET AL. J. VIROL.
ropathogenic PVF-e5 MuLV at the day 14 time point shown
(Fig. 1B) or at other time points examined (21 days and 16
weeks postinfection [data not shown]), despite comparable
levels of expression of viral SU gp70 (compare Fig. 1G with H).
As expected, brain sections from medium-inoculated rats ex-
hibited essentially no iNOS immunoreactivity at any time point
examined (Fig. 1C, 14 days postinfection). Neither PVC-211
MuLV (Fig. 1D) nor PVF-e5 MuLV (Fig. 1E) infection af-
fected the expression of eNOS compared to that observed in
brains from medium-inoculated rats (Fig. 1F) at the day 14
postinfection time point shown or at other time points exam-
ined (21 days and 16 weeks postinfection [data not shown]).
To further examine the expression of iNOS in virus-infected
BCECs, we isolated primary BCECs from virus-infected rat
brains and propagated them in vitro for 2 weeks, which is the
minimum time required to obtain sufcient cell numbers for
subsequent analysis. As a positive control, we used BCECs
stimulated with LPS, a known inducer of iNOS. When RT-
PCR analysis using specic primers was carried out on BCEC
RNA from rats inoculated with medium, PVC-211 MuLV, or
PVF-e5 MuLV, iNOS transcripts could be detected in BCECs
stimulated with LPS as well as in BCECs from PVC-211
MuLV-infected rats, but not in BCECs from rats injected with
PVF-e5 MuLV or medium (Fig. 2A). eNOS was expressed at
equivalent levels in all samples (Fig. 2A). Consistent with this
result, a 130-kDa iNOS protein was specically detected in
only LPS-stimulated BCECs or BCECs from PVC-211 MuLV-
infected rats (Fig. 2B). The successful infection of BCECs with
both PVC-211 MuLV and PVF-e5 MuLV was demonstrated
by the expression of viral envelope SU proteins (gPr85 and
gp70) (Fig. 2B). All samples expressed low levels of eNOS
protein (data not shown).
Detection of iNOS enzymatic activity in PVC-211 MuLV-
infected BCECs. iNOS catalytic activity was also directly mea-
sured in extracts of BCECs from either virus-infected or me-
dium-inoculated rats. iNOS enzymatic activity was measured
by the conversion of
L-arginine to NO and L-citrulline. Since it
has been reported that the expression of ecotropic MuLV
envelope proteins can affect the intracellular arginine concen-
tration due to down-modulation of the cationic amino acid
transporter CAT-1 (42), iNOS enzymatic activity was deter-
mined by measuring the conversion of radioactive
L-arginine
into
L-citrulline in the presence of EGTA to inhibit Ca
2
-
dependent eNOS activity. As a positive control, we used a
murine macrophage cell line, RAW264.7, which showed a sev-
enfold increase in iNOS activity after LPS stimulation (data
FIG. 1. Expression of iNOS in PVC-211 MuLV-infected brains. Brain sections (cerebellum) from rats inoculated 14 days previously with
PVC-211 MuLV (A, D, and G), PVF-e5 MuLV (B, E, and H) or medium (C, F, and I) were xed and stained for iNOS (A to C), eNOS (D to
F), or viral envelope proteins (G to I) using the ABC peroxidase technique, with diaminobenzidine as substrate. Panels A, B, C, G, H, and I show
40-m-thick frozen xed sections. Panels D, E, and F show 8-m-thick parafn sections. Final magnication for all panels is 25.
V
OL. 77, 2003 iNOS LEVELS AND Tyr NITRATION IN MuLV-INFECTED BCECs 5147
not shown). As shown in Fig. 3, the iNOS activities in extracts
of BCECs from both medium- and PVF-e5 MuLV-inoculated
rats were almost undetectable. In contrast, a marked increase
in iNOS activity (20-fold) was detected in extracts of BCECs
from PVC-211-infected rats, in agreement with the results ob-
tained for the specic expression of iNOS mRNA and protein
(Fig. 1 and 2). This difference is signicant as determined by
Students t test (P 0.00001).
Detection of tyrosine-nitrated protein in PVC-211 MuLV-
infected BCECs. NTyr is a biomarker for the presence of NO.
Since elevated immunoreactivity to NTyr was recently reported
in brain sections from PVC-211 MuLV-infected rats (43), we
tested more precisely for the presence of tyrosine-nitrated pro-
teins in extracts of BCECs from PVC-211 MuLV-infected rats.
The specicity of the anti-NTyr antibody was conrmed by the
recognition of tyrosine-nitrated bovine serum albumin (NTyr-
BSA) by Western blot analysis (Fig. 4A). This antiserum de-
tected multiple bands with a wide range of molecular masses,
mostly between 20 and 94 kDa, in all BCEC samples (Fig. 4B,
lanes 1 to 3). Interestingly, a more intense immunoreactive
band with a molecular mass of 32 kDa was found in PVC-211
MuLV-infected BCECs (lane 2). This 32-kDa protein appears
to be a cellular protein, not a viral protein, because it migrates
slower than MuLV p30
gag
(lane 4) and, unlike MuLV p30, it is
not present in a 65-kDa precursor (p65
gag
). Specic binding of
anti-NTyr was conrmed by blocking experiments using excess
NTyr (lanes 5 to 7).
Infection of BCECs by PVC-211 MuLV in vitro fails to
induce iNOS expression. We next examined whether infection
of primary BCECs by PVC-211 MuLV in vitro induces iNOS.
In vivo, BCECs are repeatedly infected by circulating virus in
the blood. To create a similar situation in vitro, BCECs were
seeded on day 0 and inoculated with PVC-211 MuLV one time
(on day 1), two times (on days 1 and 2), or four times (on days
1, 2, 3, and 4) (Fig. 5). Compared to a single inoculation,
inoculation of BCECs on two or four separate occasions sig-
FIG. 2. Specic expression of iNOS in PVC-211 MuLV-infected BCECs. (A) RNA was isolated from BCECs cultured from the brains of rats
inoculated with medium, PVC-211 MuLV, or PVF-e5 MuLV. RT-PCR was then carried out using specic primers to detect iNOS, eNOS, and
-actin. RNA from LPS-stimulated BCECs was used as a positive control for iNOS. (B) Protein extracts (20 g) of BCECs cultured from the brains
of rats inoculated with medium, PVC-211 MuLV, or PVF-e5 MuLV were separated by SDS8% PAGE, and Western blot analysis was performed
using anti-iNOS antibody. The anti-iNOS antibody was stripped from the membrane, and it was reprobed with goat anti-MuLV envelope SU
(gp70) antibody. Viral envelope proteins, gp70 and its noncleaved precursor gPr85, were detected. The same membrane was reprobed with
anti--tubulin antibody as an internal control. LPS-stimulated BCECs were analyzed separately as a positive control for iNOS.
FIG. 3. Measurement of enzymatic iNOS activity. Cell extracts of
BCECs from rats inoculated with medium or virus were tested for
iNOS enzymatic activity as described in Materials and Methods. All
measurements were determined in triplicate, and the data are shown
as means standard deviations. iNOS activity detected in BCECs
from rats inoculated with PVC-211 MuLV is signicantly higher (P
0.00001 by Students t test) than that in BCECs from rats inoculated
with medium or PVF-e5 MuLV.
5148 JINNO-OUE ET AL. J. VIROL.
nicantly improved the efciency of virus infection (compare
lane 4 to lanes 5 and 6). Almost equivalent levels of viral
envelope proteins were detected between BCECs infected with
PVC-211 MuLV in vivo and in vitro (compare lane 2 to lanes
5 and 6). The expression of iNOS protein, however, was not
detected in BCECs infected with virus in vitro (lanes 4 to 6).
This result suggests that an additional factor(s) induced in vivo
after PVC-211 MuLV infection might be necessary to induce
iNOS in BCECs.
DISCUSSION
Our previous studies have shown that neurodegeneration
induced by PVC-211 MuLV is indirect and occurs after the
infection of BCECs by the virus. Since excessive production of
NO can lead to neurotoxicity, we carried out studies to deter-
mine if infection of BCECs results in activation of iNOS or
eNOS, which produce NO from
L-arginine. Not only could we
detect iNOS immunoreactivity localized to capillaries in brain
sections from PVC-211 MuLV-infected rats, but also BCECs
isolated from these brains expressed elevated levels of iNOS
RNA, protein, and enzymatic activity compared to levels in
BCECs from medium-inoculated rats or from rats infected
with the nonneuropathogenic PVF-e5 MuLV. This is in con-
trast to eNOS, which continues to be expressed at a low,
constitutive level after PVC-211 MuLV infection. Since NTyr
is a biomarker for the presence of NO, we also examined
infected BCECs for proteins modied by tyrosine nitration and
detected elevated amounts of a 32-kDa tyrosine-nitrated cel-
lular protein in BCECs from PVC-211 MuLV-infected rats.
Thus, our data suggest that PVC-211 MuLV may be causing
neuronal death indirectly by infecting BCECs and triggering,
through iNOS activation, the production of large amounts of
NO.
The precise mechanism by which PVC-211 MuLV induces
the expression of iNOS in BCECs remains to be elucidated.
One possibility is that a viral protein expressed within BCECs
is initiating a cascade of events culminating in activation of
iNOS. However, we failed to detect elevated iNOS protein in
BCECs after in vitro infection with PVC-211 MuLV, suggest-
ing that expression of a viral protein(s) in these cells is not
sufcient to activate iNOS. Recently, Wilt et al. (43) reported
the appearance of activated microglia adjacent to PVC-211
MuLV-infected BCECs before neuronal damage. Activation
of microglia appears to be important for disease induction,
since activated microglia could not be detected in brains from
rats infected with the nonneuropathogenic PVF-e5 MuLV
(43). Activated microglia release a variety of inammatory
molecules which are known mediators of iNOS induction (16).
FIG. 4. Expression of tyrosine-nitrated proteins in BCECs from PVC-211 MuLV-infected rats. (A) BSA (0.1 g) (lane 1) or tyrosine-nitrated
BSA (0.1 g) (lane 2) was separated by SDS4-to-20% PAGE, and Western blot analysis using anti-NTyr antibody was performed. The arrow
represents the migration of BSA. (B) Protein extracts (20 g) of cultured BCECs from medium- (lanes 1 and 5), PVC-211 MuLV- (lanes 2 and
6), or PVF-e5 MuLV- (lanes 3 and 7) inoculated rats were separated by SDS4-to-20% PAGE, and Western blot analysis was performed using
anti-NTyr antibody. The asterisk represents a 32-kDa protein. The anti-NTyr antibody was stripped from the membrane, and it was reprobed with
goat anti-MuLV gag capsid (p30
gag
) antibody. Gag proteins, p30
gag
and its noncleaved precursor p65
gag
, were detected (lane 4). The right panel
is the result of Western blot analysis performed using anti-NTyr antibody pretreated with 10 mM NTyr. The same membrane was reprobed with
anti--tubulin antibody as an internal control.
V
OL. 77, 2003 iNOS LEVELS AND Tyr NITRATION IN MuLV-INFECTED BCECs 5149
If such cytokines or chemokines are released from the acti-
vated microglia in the brains of PVC-211 MuLV-infected rats,
they may bind to receptors on BCECs and induce the expres-
sion of iNOS. It is unclear how microglia become activated in
the brains of PVC-211 MuLV-infected rats. Virus cannot be
detected in the microglia (19), but the virus or a viral protein
could be interacting with a transmembrane receptor on these
cells that results in their activation. An attractive candidate is
TLR4, the transmembrane component of the receptor complex
mediating the cellular response to LPS (3). It was recently
reported that the envelope protein of Moloney MuLV, a ret-
rovirus highly related to PVC-211 MuLV, can bind to TLR4
(36) and that TLR4 is expressed on rat microglia (25). Thus,
activation of microglia in the brains of PVC-211 MuLV-in-
fected rats could be due to interaction of the viral envelope
protein with TLR4 on these cells. BCECs may also express
TLR4 that can be activated by the viral envelope protein. This
could lead to direct activation of iNOS in the BCEC, since one
of the signals activated by TLR4 binding is NF-B, a transcrip-
tional activator of iNOS (45). The failure of PVF-e5 MuLV to
cause neurological disease could be due to the inability of its
envelope glycoprotein, which differs from that of PVC-211
MuLV, to interact with TLR4 on microglia or BCECs. Exper-
iments to test these hypotheses are in progress. Interestingly,
C3H/HeJ mice, which do not express TLR4 due to a missense
mutation in the gene (34), are resistant to PVC-211 MuLV-
induced neurological disease (19), suggesting that TLR4 may
indeed play a role in PVC-211 MuLV disease induction.
The exact pathways by which excessive NO production by
PVC-211 MuLV-infected BCECs causes neuronal death are
not known. Reactive nitrogen species potentially generated
from NO by multiple pathways (8, 11, 15, 27) can modify
proteins by nitration of tyrosine residues (1, 2). Tyrosine ni-
tration may alter a proteins conformation, structure, catalytic
activity, and/or susceptibility to protease digestion (23, 38, 46)
and has been shown to disrupt protein tyrosine kinase-related
signal transduction (10, 18, 22). In this study, we showed in-
creased anti-NTyr immunoreactivity of a 32-kDa cellular pro-
tein in PVC-211 MuLV-infected BCECs. It is possible that the
modication of this protein by nitration might alter its physi-
ological properties and affect the function of the BCECs. Since
BCECs are a component of the blood-brain barrier and play an
important role in maintaining the ion homeostasis of the CNS
(33), disruption of their function by NO could indirectly lead to
the death of neurons. Alternatively, NO released from PVC-
211 MuLV-infected BCECs may move to the site of neurons
where it generates reactive nitrogen species that directly dam-
age the neurons. Although neurons are spatially separated
from BCECs, models of potential diffusion of NO indicate that
it can diffuse as far as 300 m from its site of origin, which
could include as many as 2 million synapses (44).
In addition to affecting BCECs and neurons, NO produced
by PVC-211 MuLV-infected BCECs may also have other ef-
fects in the brain. NO generated by iNOS in human microvas-
cular endothelial cells has been shown to inhibit the rolling and
adhesion of leukocytes (4), and NO has been shown to inhibit
proliferation of T cells that invade the CNS (20). Since PVC-
211 MuLV-induced neurodegeneration is not associated with
either inammation or inltration of leukocytes (19, 21), NO
produced in PVC-211 MuLV-infected BCECs may be inhibit-
ing adhesion and inltration of leukocytes into the brain pa-
renchyma or proliferation of T cells that do inltrate. In addi-
tion, NO produced in PVC-211 MuLV-infected BCECs may
also inhibit virus replication. Data from a number of labora-
tories using both RNA and DNA viruses suggest that NO may
inhibit an early stage in viral replication (37). The virus titer in
brains recovered from rats infected with PVC-211 MuLV is
always signicantly lower than that in the spleen (19, 21, 43). In
addition, microglia or neurons adjacent to the infected BCECs
show no evidence of PVC-211 MuLV infection despite the
release of virions into the basement membrane of infected
BCECs. Thus, it is possible that NO released from PVC-211
MuLV-infected BCECs inhibits replication and spread of the
virus in the brain.
In summary, our results show that infection of BCECs with
PVC-211 MuLV in vivo induces functionally active iNOS pro-
tein and elevates tyrosine nitration of a 32-kDa cellular pro-
tein. Our failure to detect iNOS expression in BCECs after in
vivo infection with the nonneuropathogenic but BCEC-tropic
PVF-e5 MuLV suggests that iNOS activation plays a crucial
role in the development of neurological disease induced by
PVC-211 MuLV. The use of iNOS inhibitors and the genera-
FIG. 5. Infection of BCECs by PVC-211 MuLV in vitro fails to
induce iNOS. Primary BCECs (twice passaged) were inoculated in
vitro with PVC-211 MuLV either one time, two times, or four times,
24 h apart. Fourteen days after the initial inoculation, cell extracts were
prepared and compared with extracts of BCECs obtained from rats
inoculated with PVC-211 MuLV in vivo. Lane 1, BCECs from rats
inoculated in vivo with medium; lane 2, BCECs from rats inoculated in
vivo with PVC-211 MuLV; lane 3, mock-infected BCECs; lane 4,
BCECs infected in vitro with PVC-211 MuLV one time; lane 5, BCECs
infected in vitro with PVC-211 MuLV two times; lane 6, BCECs
infected in vitro with PVC-211 MuLV four times. Extracts (10 g)
were separated by SDS4-to-20% PAGE, and Western blot analysis
was performed using goat anti-MuLV envelope SU (gp70) antibody
(upper panel). The anti-MuLV gp70 antibody was stripped from the
membrane and reprobed with anti-iNOS antibody (middle panel). The
same membrane was reprobed with anti--tubulin antibody as an in-
ternal control (bottom panel).
5150 JINNO-OUE ET AL. J. V
IROL.
tion of iNOS-decient rats should help to clarify how iNOS is
involved in the pathological process.
ACKNOWLEDGMENTS
We thank Takashi Yugawa, Karen Rulli, and Joan Cmarik for help-
ful advice.
REFERENCES
1. Beckman, J. S., T. W. Beckman, J. Chen, P. A. Marshall, and B. A. Freeman.
1990. Apparent hydroxyl radical production by peroxynitrite: implications for
endothelial injury from nitric oxide and superoxide. Proc. Natl. Acad. Sci.
USA 87:16201624.
2. Beckman, J. S. 1996. Oxidative damage and tyrosine nitration from peroxyni-
trite. Chem. Res. Toxicol. 9:836844.
3. Beutler, B. 2000. TLR4: central component of the sole mammalian LPS
sensor. Curr. Opin. Immunol. 12:2026.
4. Binion, D. G., S. Fu, K. S. Ramanujam, Y. C. Chai, R. A. Dweik, J. A. Drazba,
J. G. Wade, N. P. Ziats, S. C. Erzurum, and K. T. Wilson. 1998. iNOS
expression in human intestinal microvascular endothelial cells inhibits leu-
kocyte adhesion. Am. J. Physiol. 275:G592G603.
5. Boven, L. A., L. Gomes, C. Herry, F. Gray, J. Verhof, P. Portegies, M.
Tardieu, and H. S. Nottet. 1999. Increased peroxynitrite activity in AIDS
dementia complex: implications for the neuropathogenesis of HIV-1 infec-
tion. J. Immunol. 162:43194327.
6. Bowman, P. D., A. L. Betz, J. S. Wolinsky, J. B. Penny, R. R. Shivers, and
G. W. Goldstein. 1981. Primary culture of capillary endothelium from rat
brain. In Vitro 17:353362.
7. Bredit, D. S., and S. H. Synder. 1994. Nitric oxide: a physiologic messenger
molecule. Annu. Rev. Biochem. 63:175195.
8. Brennan, M.-L., W. Wu, X. Fu, Z. Shen, W. Song, H. Frost, C. Vadseth, L.
Narine, E. Lenkiewicz, M. T. Borchers, A. J. Lusis, and S. L. Hazen. 2002. A
tale of two controversies: dening both the role of peroxidases in nitroty-
rosine formation in vivo using eosinophil peroxidase and myeloperoxidase-
decient mice, and the nature of peroxidase-generating reactive nitrogen
species. J. Biol. Chem. 277:1741517427.
9. Dawson, T. M., and V. L. Dawson. 1998. Nitric oxide in neurodegeneration.
Prog. Brain Res. 118:215229.
10. Di Stasi, A. M., C. Mallozzi, G. Macchia, T. C. Petrucci, and M. Minetti.
1999. Peroxynitrite induces tyrosine nitration and modulates tyrosine phos-
phorylation of synaptic proteins. J. Neurochem. 73:727735.
11. Eiserich, J. P., C. E. Cross, A. D. Jones, B. Halliwell, and A. van der Vliet.
1996. Formation of nitrating and chlorinating species by reaction of nitrite
with hypochlorous acid. J. Biol. Chem. 271:1919919208.
12. Ferrante, R. J., P. Hantray, E. Brouillet, and M. F. Beal. 1999. Increased
nitrotyrosine immunoreactivity in substantia nigra neurons in MPTP treated
baboons is blocked by inhibition of neuronal nitric oxide synthase. Brain Res.
823:177182.
13. Forstemann, U., E. I. Closs, J. S. Pollock, M. Nakane, P. Schwarz, I. Gath,
and H. Kleinert. 1994. Nitric oxide synthase isozymes. Characterization,
purication, molecular cloning, and functions. Hypertension 23:11211131.
14. Galea, E., D. L. Feinstein, and D. L. Reis. 1992. Induction of calcium-
independent nitric oxide synthase activity in primary rat glial cultures. Proc.
Natl. Acad. Sci. USA 89:1094510949.
15. Gaut, J. P., J. Byun, H. D. Tran, W. M. Lauber, J. A. Carroll, R. S. Hotchkiss,
A. Belaaouaj, and J. W. Heinecke. 2002. Myeloperoxidase produces nitrating
oxidants in vivo. J. Clin. Investig. 109:13111319.
16. Gehrmann, J., Y. Matsumoto, and G. W. Kreutzberg. 1995. Microglia: in-
trinsic immuno-effector cell of brain. Brain Res. Rev. 20:269287.
17. Good, P. F., P. Werner, A. Hsu, C. W. Olanow, and D. P. Perl. 1996. Evidence
for neuronal oxidative damage in Alzheimers disease. Am. J. Pathol. 49:21
27.
18. Hevel, J. M., K. A. White, and M. A. Marletta. 1991. Purication of the
inducible murine macrophage nitric oxide synthase. Identication as a a-
voprotein. J. Biol. Chem. 266:2278922791.
19. Hoffman, P. M., E. H. Cimino, D. S. Robbins, R. D. Broadwell, J. M. Powers,
and S. K. Ruscetti. 1992. Cellular tropism and localization in the rodent
nervous system of a neuropathogenic variant of Friend murine leukemia
virus. Lab. Investig. 67:314321.
20. Juedes, A. E., and N. H. Ruddle. 2001. Resident and inltrating central
nervous system APCs regulate the emergence and resolution of experimental
autoimmune encephalomyelitis. J. Immunol. 166:51685175.
21. Kai, K., and T. Furuta. 1984. Isolation of paralysis-inducing murine leuke-
mia viruses from Friend virus passaged in rats. J. Virol. 50:970973.
22. Kong, S. K., M. B. Yim, E. R. Stadtman, and P. B. Chock. 1996. Peroxynitrite
disables the tyrosine phosphorylation regulatory mechanism: lymphocyte-
specic tyrosine kinase fails to phosphorylate nitrated cdc2(620)NH2 pep-
tide. Proc. Natl. Acad. Sci. USA 93:33773382.
23. Kuo, W. N., J. M. Kreahling, V. P. Shanbhag, P. P. Shanbhag, and M.
Mewar. 2000. Protein nitration. Mol. Cell. Biochem. 214:121129.
24. Lee, S. C., M. L. Zhao, A. Hirano, and D. W. Dickson. 1999. Inducible nitric
oxide synthase immunoreactivity in the Alzheimer disease hippocampus:
association with Hirano bodies, neurobrillary tangles, and senile plaques.
J. Neuropathol. Exp. Neurol. 58:11631169.
25. Lehnardt, S., C. Lachance, S. Patrizi, S. Lefebvre, P. L. O. Follett, F. E.
Jensen, P. A. Rosenberg, J. J. Volpe, and T. Vartanian. 2002. The toll-like
receptor TLR4 is necessary for lipopolysaccharide-induced oligodendrocyte
injury in the CNS. J. Neurosci. 22:24782486.
26. Liberatore, G. T., V. Jackson-Lewies, S. Vukosavic, M. Vila, W. G. McAuliffe,
T. M. Dawson, and S. Przedorski. 1999. Inducible nitric oxide synthase
stimulates dopaminergic neurodegeneration in the MPTP model of Parkin-
son disease. Nat. Med. 5:14031409.
27. Lipton, S. A. 1999. Neuronal protection and destruction by NO. Cell Death
Differ. 6:943951.
28. Masuda, M., M. P. Remington, P. M. Hoffman, and S. K. Ruscetti. 1992.
Molecular characterization of a neuropathogenic and nonerythroleukemic
variant of Friend murine leukemia virus PVC-211. J. Virol. 66:27982806.
29. Masuda, M., P. M. Hoffman, and S. K. Ruscetti. 1993. Viral determinants
that control neuropathogenicity of PVC-211 murine leukemia virus in vivo
determine brain capillary endothelial cell tropism of the virus in vitro. J. Vi-
rol. 67:45804587.
30. Masuda, M., C. A. Hanson, W. G. Alvord, P. M. Hoffman, S. K. Ruscetti, and
M. Masuda. 1996. Effect of subtle changes in the SU protein of ecotropic
murine leukemia virus on its brain capillary endothelial cell tropism and
interference properties. Virology 215:142151.
31. Moncada, S., R. M. Palmer, and E. A. Higgs. 1991. Nitric oxide: physiology,
pathophysiology, and pharmacology. Pharmacol. Rev. 43:109142.
32. Nathan, C., and Q.-W. Xie. 1994. Nitric oxide synthases: roles, tolls, and
controls. Cell 78:915918.
33. Pardridge, W. M. 1999. Blood-brain barrier biology and methodology.
J. Neurovirol. 5:556569.
34. Poltorak, A., X. He, I. Smirnova, M.-Y. Liu, C. Van Huffel, X. Du, D.
Birdwell, E. Alejos, M. Silva, C. Galanos, M. Freudenberg, P. Ricciardi-
Castagnoli, B. Layton, and B. Beutler. 1998. Defective LPS signaling in
C3H/HeJ and C57BL/10ScCr mice: mutations in Tlr4 gene. Science 282:
20852088.
35. Power, C. 2001. Retroviral diseases of the nervous system: pathogenic host
response or viral gene-mediated neurovirulence? Trends Neurosci. 24:162
169.
36. Rassa, J. C., J. L. Meyers, Y. Zhang, R. Kudaravalli, and S. R. Ross. 2002.
Murine retroviruses activate B cells via interaction with toll-like receptor 4.
Proc. Natl. Acad. Sci. USA 99:22812286.
37. Reiss, C. S., and T. Komatsu. 1998. Does nitric oxide play a critical role in
viral infections? J. Virol. 72:45474551.
38. Roberts, E. S., H. L. Lin, J. R. Crowley, J. L. Vuletich, Y. Osawa, and P. F.
Hollenberg. 1998. Peroxynitrite-mediated nitration of tyrosine and inactiva-
tion of the catalytic activity of cytochrome P450 2B1. Chem. Res. Toxicol.
11:10671074.
39. Rosenberg, N., and P. Jolicoeur. 1997. Retroviral pathogenesis, p. 475586.
In J. M. Cofn, S. H. Hughes, and H. E. Varmus (ed.), Retroviruses. Cold
Spring Harbor Laboratory Press, Cold Spring Harbor, N.Y.
40. Rowe, W. P., W. E. Pugh, and J. W. Hartley. 1970. Plaque assay techniques
for murine leukemia viruses. Virology 42:136138.
41. Sasaki, S., N. Shibata, T. Komori, and M. Iwata. 2000. iNOS and nitroty-
rosine immunoreactivity in amyotrophic lateral sclerosis. Neurosci. Lett.
291:4448.
42. Wang, H., E. Dechant, M. Kavanaugh, R. A. North, and D. Kabat. 1982.
Effects of ecotropic murine retroviruses on the dual-function cell surface
receptor/basic amino acid transporter. J. Biol. Chem. 267:2361723624.
43. Wilt, S. G., N. V. Dugger, N. D. Hitt, and P. M. Hoffman. 2000. Evidence for
oxidative damage in a murine leukemia virus-induced neurodegeneration.
J. Neurosci. Res. 62:440450.
44. Wood, J., and J. Garthwaite. 1994. Models of the diffusional spread of nitric
oxide: implications for neural nitric oxide signalling and its pharmacological
properties. Neuropharmacology 33:12351244.
45. Xie, Q. W., Y. Kashiwabara, and C. Nathan. 1994. Role of transcription
factor NF-kappa B/Rel in induction of nitric oxide synthase. J. Biol. Chem.
269:47054708.
46. Yamakura, F., H. Taka, T. Fujimura, and K. Murayama. 1998. Inactivation
of human manganese-superoxide dismutase by peroxynitrite is caused by
exclusive nitration of tyrosine 34 to 3-nitrotyrosine. J. Biol. Chem. 273:
1408514089.
VOL. 77, 2003 iNOS LEVELS AND Tyr NITRATION IN MuLV-INFECTED BCECs 5151
... The spongiform vacuolation observed in PVC-211 MuLVinfected brains is associated with oxidative damage (47), and BCEC isolated from PVC-211 MuLV-infected rats produce inducible nitric oxide synthase (iNOS) (23). However, iNOS was not induced after in vitro infection of primary BCEC, suggesting that expression of the virus in BCEC is insufficient to activate iNOS. ...
... Virus infection of primary BCEC. Primary rat BCEC were isolated, cultured, and infected with virus as described previously (23). Briefly, cells were rinsed three times in phosphate-buffered saline (PBS) to remove heparin and pretreated for 30 min with DMEM containing 10% FCS and 5 g/ml of Polybrene. ...
... Proteins were separated by electrophoresis on an 8% Tris-glycine gel (Invitrogen) and then transferred to a nitrocellulose membrane (Invitrogen). The membrane was then incubated with anti-RLV gp70 serum as previously described (23). Bound antibody was detected using peroxidase-labeled secondary antibody and an enhanced chemiluminescence system (GE Healthcare, Buckinghamshire, United Kingdom). ...
Article
Full-text available
PVC-211 murine leukemia virus (MuLV) is a neuropathogenic retrovirus that has undergone genetic changes from its nonneuropathogenic parent, Friend MuLV, that allow it to efficiently infect rat brain capillary endothelial cells (BCEC). To clarify the mechanism by which PVC-211 MuLV expression in BCEC induces neurological disease, we examined virus-infected rats at various times during neurological disease progression for vascular and inflammatory changes. As early as 2 weeks after virus infection and before any marked appearance of spongiform neurodegeneration, we detected vessel leakage and an increase in size and number of vessels in the areas of the brain that eventually become diseased. Consistent with these findings, the amount of vascular endothelial growth factor (VEGF) increased in the brain as early as 1 to 2 weeks postinfection. Also detected at this early disease stage was an increased level of macrophage inflammatory protein 1 alpha (MIP-1 alpha), a cytokine involved in recruitment of microglia to the brain. This was followed at 3 weeks postinfection by a marked accumulation of activated microglia in the spongiform areas of the brain accompanied by an increase in tissue plasminogen activator, a product of microglia implicated in neurodegeneration. Pathological observations at the end stage of the disease included loss of neurons, decreased myelination, and mild muscle atrophy. Treatment of PVC-211 MuLV-infected rats with clodronate-containing liposomes, which specifically kill microglia, significantly blocked neurodegeneration. Together, these results suggest that PVC-211 MuLV infection of BCEC results in the production of VEGF and MIP-1 alpha, leading to the vascular changes and microglial activation necessary to cause neurodegeneration.
... In CasBrE, the Env protein accompanied with differential processing of N-linked sugar in glial cells is correlated with neuropathogenicity (37). On the other hand, some cellular effectors involving several cytokines and iNOS have been found to be aberrantly expressed in the brains of infected animals, but it is not clear whether these effectors are required for the development of neurodegenerative lesions (38)(39)(40)(41)(42)(43)(44)(45). ...
... This finding indicates that, in A8 virus, accumulation of the Env precursor polyprotein or differential glycosylation of Env protein is not correlated with neuropathogenicity. Likewise, in PVC211, which is a neuropathogenic variant of Fr-MLV (6,15) and closely related to A8, it was reported that accumulation of the Env precursor polyprotein was not detected (44). ...
Article
Friend murine leukemia virus clone A8 causes spongiform neurodegeneration in the rat brain, and the env gene of A8 is a primary determinant of neuropathogenicity. In order to narrow down the critical region within the env gene that determines neuropathogenicity, we constructed chimeric viruses having chimeric env between A8 and non-neuropathogenic 57 on the background of A8 virus. After replacement of the BamHI (at nucleotide 5715)-AgeI (at nucleotide 6322) fragment of A8 virus with the corresponding fragment of 57, neuropathogenicity was lost. In contrast, the chimeric viruses that have the BamHI (5715)-AgeI (6322) fragment of A8 induced spongiosis in 100% of infected rats at the same or slightly lower intensity than A8 virus. These results indicate that the BamHI (5715)-AgeI (6322) fragment of A8, which contains the signal sequence and the N-terminal half of RBD, is crucial for the induction of spongiform neurodegeneration. In the BamHI (5715)-AgeI (6322) fragment, seven amino acids differed between A8 and 57, one in the signal sequence and six in RBD, which suggests that these amino acids significantly contribute to the neuropathogenicity of A8.
... Proteins were separated by electrophoresis on an 8% Tris-glycine gel (Invitrogen) and then transferred to a nitrocellulose membrane (Invitrogen). The membrane was then incubated with anti-RLV gp70 serum as previously described (Jinno-Oue et al., 2003). Bound antibody was detected using peroxidase-labeled secondary antibody and the enhanced chemiluminescence system (GE Healthcare, Buckinghamshire, UK). ...
Article
We recently reported that infection of rats with the neurodegenerative disease-causing retrovirus PVC-211 MuLV results in elevated levels of the chemokine MIP-1α followed by the accumulation of activated microglia in the brain. To investigate the importance of MIP-1α in recruitment of microglia to the brain, we treated rats with MIP-1α antibodies before and after PVC-211 MuLV infection. This caused a delay in the development of paralysis which was associated with a decrease in activated microglia without affecting virus expression. To determine the source of activated microglia, rats were splenectomized 4 days after virus infection. Splenectomized rats showed a delay in disease development that was associated with decreased numbers of activated microglia without affecting virus expression. Together, these results suggest that MIP-1α is directly involved in the neurodegeneration induced in rats by PVC-211 MuLV by recruiting macrophages/microglia from the periphery into regions of the brain that eventually become diseased.
Article
In addition to the env gene, a 0.3-kb fragment containing the R-U5-5' leader sequence is essential for the induction of spongiform neurodegeneration by Friend murine leukemia virus (Fr-MLV) clone A8 and it also influences expression of the Env protein. Kinetic studies were carried out using two recombinant viruses, R7f, carrying the A8 0.3-kb fragment, and Rec5, carrying the 0.3-kb fragment of the non-neuropathogenic Fr-MLV clone 57. These analyses suggested that the 0.3-kb fragment influenced the expression level of the Env protein by regulating the amount of spliced env-mRNA rather than the amount of total viral mRNA or viral production.
Article
This chapter examines the current information on endothelial cell activation, providing overviews of some of the typical endothelial responses and activation stimuli. It focuses on the activators of vascular endothelium, providing current understanding of their activating mechanisms. The multitude of stimuli is broadly grouped by similarity of structure and function, ranging from cell-derived mediators, to microbial organisms, and to mechanical stimuli. Endothelial cells respond to activating stimuli by undergoing phenotypic changes, with an increased metabolic activity and increases in (1) adhesion for leukocytes, (2) permeability, (3) proliferation, (4) pro-coagulant activity, and (5) secretion. A particular stimulus can activate one or a combination of these endothelial phenotypes. More commonly, a stimulus triggers a combination of these phenotypes, the predominance of phenotype(s) being characteristic of the particular stimulus.
Article
Full-text available
Renal lipid metabolism may play important roles in renal inflammation, glomerulosclerosis and tubulointerstitial injury in diabetic nephropathy. These alterations in lipids are associated with (1) decreased expression of PPAR-γ mRNA and protein, (2) increased abundance of the sterol regulatory element binding protein-1 (SREBP-1), key regulator of fatty acid synthesis, (3) decreased abundance of farnesoid X receptor (FXR), a negative regulator of fatty acid synthesis and promoter of fatty acid oxidation, (4) downregulation of peroxisome proliferator-activated receptor delta (PPAR-γ), key regulator of fatty acid oxidation, (5) increased abundance of the sterol regulatory element binding protein-2 (SREBP-2), key regulator of cholesterol synthesis, and (6) downregulation of ATP binding cassette A1 (ABCA1), key regulator of cholesterol efflux. These lipid alterations are also associated with marked downregulation of the podocyte markers podocin and zonula occludens-1 (ZO-1) and proteinuria. Treatment of ZDF rats with the PPAR-γ agonist rosiglitazone results in normalization of the renal lipid metabolism pathways and prevention of lipid and adipophilin accumulation, restoration of podocin and ZO-1 expression, and prevention of proteinuria. Thus, our results indicate that renal lipid accumulation significantly contributes to renal cell injury and treatment with PPAR agonist significantly ameliorates podocyte injury, glomerulosclerosis, tubulointerstitial fibrosis, and proteinuria. PPAR-FXR-SREBP pathway may play a critical role in regulation of lipid homeostasis and fibrosis in the kidney. [Nature and Science. 2009;7(1):91-95]. (ISSN: 1545-0740).
Article
A readily obtainable in vitro paradigm of the blood-brain barrier (BBB) would offer considerable benefits. Toward this end, in this study, we describe a novel method for purifying murine brain microvascular endothelial cells (BMEC) for culture. The method uses limited collagenase-dispase digestion of enriched brain microvessels, followed by immunoisolation of digested, microvascular fragments by magnetic beads coated with antibody to platelet-endothelial cell adhesion molecule-1. When plated onto collagen IV-coated surfaces, these fragments elaborated confluent monolayers of BMEC that expressed, as judged by immunocytochemistry, the adherens junction-associated proteins, VE-cadherin and beta-catenin, as well as the tight junction (TJ)-associated proteins, claudin-5, occludin, and zonula occludin-1 (ZO-1), in concentrated fashion along intercellular borders. In contrast, cultures of an immortalized and transformed line of murine brain capillary-derived endothelial cells, bEND.3, displayed diffuse cytoplasmic localization of occludin and ZO-1. This difference in occludin and ZO-1 staining between the two endothelial cell types was also reflected in the extent of association of these proteins with the detergent-resistant cytoskeletal framework (CSK). Although both occludin and ZO-1 largely partitioned with the CSK fraction in BMEC, they were found predominantly in the soluble fraction of bEND.3 cells, and claudin-5 was found associated equally with both fractions in BMEC and bEND.3 cells. Moreover, detergent-extracted cultures of the BMEC retained pronounced immunostaining of occludin and ZO-1, but not claudin-5, along intercellular borders. Because both occludin and ZO-1 are thought to be functionally coupled to the detergent-resistant CSK and high expression of TJs is considered a seminal characteristic of the BBB, these results impart that this method of purifying murine BMEC provides a suitable platform to investigate BBB properties in vitro.
Article
Full-text available
Some murine leukemia viruses (MuLVs), among them Cas-Br-E and ts-1 MuLVs, are neurovirulent, inducing spongiform myeloencephalopathy and hind limb paralysis in susceptible mice. It has been shown that the env gene of these viruses harbors the determinant of neurovirulence. It appears that neuronal loss occurs by an indirect mechanism, since the target motor neurons have not been found to be infected. However, the pathogenesis of the disease remains unclear. Several lymphokines, cytokines, and other cellular effectors have been found to be aberrantly expressed in the brains of infected mice, but whether these are required for the development of the neurodegenerative lesions is not known. In an effort to identify the specific effectors which are indeed required for the initiation and/or development of spongiform myeloencephalopathy, we inoculated gene-deficient (knockout [KO]) mice with ts-1 MuLV. We show here that interleukin-6 (IL-6), inducible nitric oxide synthetase (iNOS), ICE, Fas, Fas ligand (FasL), and TNF-R1 KO mice still develop signs of disease. However, transgenic mice overexpressing Bcl-2 in neurons (NSE/Bcl-2) were largely protected from hind limb paralysis and had less-severe spongiform lesions. These results indicate that motor neuron death occurs in this disease at least in part by a Bcl-2-inhibitable pathway not requiring the ICE, iNOS, Fas/FasL, TNF-R1, and IL-6 gene products.
Article
Full-text available
The promoter of the murine gene encoding inducible nitric oxide synthase (iNOS) contains an NF-kappa B site beginning 55 base pairs upstream of the TATA box, designated NF-kappa Bd. Reporter constructs containing truncated promoter regions, when transfected into macrophages, revealed that NF-kappa Bd is necessary to confer inducibility by bacterial lipopolysaccharide (LPS). Oligonucleotide probes containing NF-kappa Bd plus the downstream 9 or 47 base pairs bound proteins that rapidly appeared in the nuclei of LPS-treated macrophages. The nuclear proteins bound to both probes in an NF-kappa Bd-dependent manner, but binding was resistant to cycloheximide only for the shorter probe. The proteins binding both probes reacted with antibodies against p50 and c-rel but not RelB; those binding the shorter probe also reacted with anti-RelA (p65). Pyrrolidine dithiocarbamate, which acts as a specific inhibitor of NF-kappa B, blocked both the activation of the NF-kappa Bd-binding proteins and the production of NO in LPS-treated macrophages. Thus, activation of NF-kappa Bd/Rel is critical in the induction of iNOS by LPS. However, additional, newly synthesized proteins contribute to the NF-kappa Bd-dependent transcription factor complex on the iNOS promoter in LPS treated mouse macrophages.
Article
Full-text available
Exposure of primary cultures of neonatal rat cortical astrocytes to bacterial lipopolysaccharide (LPS) results in the appearance of nitric oxide synthase (NOS) activity. The induction of NOS, which is blocked by actinomycin D, is directly related to the duration of exposure and dose of LPS, and a 2-hr pulse can induce enzyme activity. Cytosol from LPS-treated astrocyte cultures, but not from control cultures, produces a Ca(2+)-independent conversion of L-arginine to L-citrulline that can be completely blocked by the specific NOS inhibitor NG-monomethyl-L-arginine. The induced NOS activity exhibits an apparent Km of 16.5 microM for L-arginine and is dependent on NADPH, FAD, and tetrahydrobiopterin. LPS also induces NOS in C6 glioma cells and microglial cultures but not in cultured cortical neurons. The expression of NOS in astrocytes and microglial cells has been confirmed by immunocytochemical staining using an antibody to the inducible NOS of mouse macrophages and by histochemical staining for NADPH diaphorase activity. We conclude that glial cells of the central nervous system can express an inducible form of NOS similar to the inducible NOS of macrophages. Inducible NOS in glia may, by generating nitric oxide, contribute to the neuronal damage associated with cerebral ischemia and/or demyelinating diseases.
Article
Full-text available
The widely expressed Na(+)-independent transporter for basic amino acids (system y+) is the cell surface receptor (ecoR) for ecotropic host-range mouse retroviruses (murine leukemia viruses (MuLVs)), a class of retroviruses that naturally infects only mice or rats. Accordingly, expression of mouse ecoR cDNA in mink CCL64 fibroblasts yields cells (CEN cells) that have y+ transporter activity above the endogenous background and that bind and are infected by ecotropic MuLVs. The effect of ecotropic MuLV infection on expression of y+ transporter was analyzed in mouse and in mink CEN fibroblasts. Chronic infection with ecotropic MuLVs caused 50-70% loss (down-modulation) of mouse y+ transporter in plasma membranes, detected as a reduced Vmax for uptake and outflow of L-[3H]arginine with no effect on Km values. Down-modulation was specific for mouse y+ and did not affect other transporters or the endogenous mink y+, suggesting that it results from specific interaction between mouse y+ and the viral envelope glycoprotein gp70 in the infected cells. Because this partial loss of mouse y+ from cell surfaces is insufficient to explain the complete interference to superinfection that occurs in cells chronically infected with ecotropic MuLVs, alternative explanations for interference are proposed. In contrast to the y+ down-modulation caused by chronic infection, binding of extracellular envelope glycoprotein gp70 at 37 degrees C resulted in noncompetitive inhibition of amino acid import by mouse y+ but had no effect on export through this same transporter or on any transporter properties of mink y+. The effects of gp70 on transport kinetics suggest that it slows the rate-limiting step of the amino acid import cycle, a conformational transition of the empty transporter in which the binding site moves from the inside back to the outside of the cell, and that gp70 has no effect on the rate-limiting step of the amino acid export cycle. Infected cells retain substantial y+ activity. Moreover, the virus binding site on ecoR is in a mobile region that changes conformation during the amino acid transport cycle.
Article
Full-text available
PVC-211 murine leukemia virus (MuLV) is a replication-competent, ecotropic type C retrovirus that was isolated after passage of the Friend virus complex through F344 rats. Unlike viruses in the Friend virus complex, it does not cause erythroleukemia but causes a rapidly progressive hind limb paralysis when injected into newborn rats and mice. We have isolated an infectious DNA clone (clone 3d) of this virus which causes neurological disease in animals as efficiently as parental PVC-211 MuLV. The restriction map of clone 3d is very similar to that of the nonneuropathogenic, erythroleukemogenic Friend murine leukemia virus (F-MuLV), suggesting that PVC-211 MuLV is a variant of F-MuLV and that no major structural alteration was involved in its derivation. Studies with chimeric viruses between PVC-211 MuLV clone 3d and wild-type F-MuLV clone 57 indicate that at least one determinant for neuropathogenicity resides in the 2.1-kb XbaI-ClaI fragment containing the gp70 coding region of PVC-211 MuLV. Compared with nonneuropathogenic ecotropic MuLVs, the env gene of PVC-211 MuLV encodes four unique amino acids in the gp70 protein. Nucleotide sequence analysis also revealed a deletion in the U3 region of the long terminal repeat (LTR) of PVC-211 MuLV clone 3d compared with F-MuLV clone 57. In contrast to the env gene of PVC-211 MuLV, particular sequences within the U3 region of the viral LTR do not appear to be required for neuropathogenicity. However, the changes in the LTR of PVC-211 MuLV may be responsible for the failure of this virus to cause erythroleukemia, because chimeric viruses containing the U3 region of F-MuLV clone 57 were erythroleukemogenic whereas those with the U3 of PVC-211 MuLV clone 3d were not.
Article
Detection of 3-nitrotyrosine has served as an in vivo marker for the production of the cytotoxic species peroxynitrite (ONOO−). We show here that reaction of nitrite (NO−2), the autoxidation product of nitric oxide (·NO), with hypochlorous acid (HOCl) forms reactive intermediate species that are also capable of nitrating phenolic substrates such as tyrosine and 4-hydroxyphenylacetic acid, with maximum yields obtained at physiological pH. Monitoring the reaction of NO−2 with HOCl by continuous flow photodiode array spectrophotometry indicates the formation of a transient species with spectral characteristics similar to those of nitryl chloride (Cl-NO2). Reaction of synthetic Cl-NO2 with N-acetyl-L-tyrosine results in the formation of 3-chlorotyrosine and 3-nitrotyrosine in ratios that are similar to those obtained by the NO−2/HOCl reaction (4:1). Tyrosine residues in bovine serum albumin are also nitrated and chlorinated by NO−2/HOCl and synthetic Cl-NO2. The reaction of N-acetyl-L-tyrosine with NO−2/HOCl or authentic Cl-NO2 also produces dityrosine, suggesting that free radical intermediates are involved in the reaction mechanism. Our data indicate that while chlorination reactions of Cl-NO2 are mediated by direct electrophilic addition to the aromatic ring, a free radical mechanism appears to be operative in nitrations mediated by NO−2/HOCl or Cl-NO2, probably involving the combination of nitrogen dioxide (·NO2) and tyrosyl radical. We propose that NO−2 reacts with HOCl by Cl+ transfer to form both cis- and trans-chlorine nitrite (Cl-ONO) and Cl-NO2 as intermediates that modify tyrosine by either direct reaction or after decomposition to reactive free and solvent-caged Cl· and ·NO2 as reactive species. Formation of Cl-NO2 and/or Cl-ONO in vivo may represent previously unrecognized mediators of inflammation-mediated protein modification and tissue injury, and offers an additional mechanism of tyrosine nitration independent of ONOO−.
Article
The addition of peroxynitrite to purified cytochrome P450 2B1 resulted in a concentration-dependent loss of the NADPH- and reductase-supported or tert-butylhydroperoxide-supported 7-ethoxy-4-(trifluoromethyl)coumarin O-deethylation activity of P450 2B1 with IC50 values of 39 and 210 μM, respectively. After incubation of P450 2B1 with 300 μM peroxynitrite, the heme moiety was not altered, but the apoprotein was modified as shown by HPLC and spectral analysis. Western blot analysis of peroxynitrite-treated P450 2B1 demonstrated the presence of an extensive immunoreactivite band after incubating with anti-nitrotyrosine antibody. However, the immunostaining was completely abolished after coincubation of the anti-nitrotyrosine antibody with 10 mM nitrotyrosine. These results indicated that one or more of the tyrosine residues in P450 2B1 were modified to nitrotyrosines. The decrease in the enzymatic activity correlated with the increase in the extent of tyrosine nitration. Further demonstration of tyrosine nitration was confirmed by GC/MS analysis by using 13C-labeled tyrosine and nitrotyrosine as internal standards; approximately 0.97 mol of nitrotyrosine per mole of P450 2B1 was found after treatment with peroxynitrite. The peroxynitrite-treated P450 2B1 was digested with Lys C, and the resulting peptides were separated by Tricine-sodium dodecyl sulfate−polyacrylamide gel electrophoresis (SDS−PAGE). The amino acid sequence of the major nitrotyrosine-containing peptide corresponded to a peptide containing amino acid residues 160−225 of P450 2B1, which contains two tyrosine residues. Thus, incubation of P450 2B1 with peroxynitrite resulted in the nitration of tyrosines at either residue 190 or 203 or at both residues of P450 2B1 concomitant with a loss of 2B1-dependent activity.
Article
Many virus infections elicit vigorous host immune responses, both innate and acquired. The immune responses are fre- quently successful in controlling and then clearing the virus, using both cellular effectors such as natural killer (NK) cells and cytolytic T lymphocytes and soluble factors such as inter- ferons (IFNs). However, some immune responses lead to pathologic changes or are unable to prevent the pathogen's growth. This review will not be devoted to the different strat- egies viruses have taken to promote their transmission or sur- vival but rather to one aspect of the innate immune response to infection: the role of nitric oxide (NO) in the antiviral reper- toire. Recently, data from many laboratories, using both RNA and DNA viruses in experimental systems, have implicated a role for NO in the immune response. The data do not indicate a magic bullet for all systems but suggest that NO may inhibit an early stage in viral replication and thus prevent viral spread, promoting viral clearance and recovery of the host. The earliest host responses to viral infections are nonspecific and involve the induction of cytokines, among them, IFNs and tumor necrosis factor alpha (TNF-a). Gamma IFN (IFN-g) and TNF-a have both been shown to be active in many cell types and induce cascades of downstream mediators (reviewed in references 25, 34, and 41). Others have found that NO synthase type 2 (NOS-2, iNOS) is an IFN-g-inducible protein in macrophages, requiring IRF-1 as a transcription factor (12, 17). We have observed that the isoform expressed in neurons, NOS-1, is IFN-g, TNF-a, and interleukin-12 (IL-12) inducible (20). Thus, NOS falls into the category of IFN-inducible pro- teins, activated during innate immune responses. NO is produced by the enzymatic modification of L-arginine to L-citrulline and requires many cofactors, including tetrahy- drobiopterine, calmodulin, NADPH, and O2. NO rapidly re- acts with proteins or with H2O2 to form ONOO 2 , peroxyni-
Chapter
Retroviruses are associated with a wide variety of diseases including an array of malignancies, immunodeficiencies, and neurologic disorders. Syndromes as seemingly diverse as arthritis, osteopetrosis, and anemia can all result from retroviral infection. These disorders afflict a large number of different creatures, ranging from clams and fish to birds and mammals, including humans. Some of these disorders have significant agricultural impact, crippling farm animals during their most productive years, whereas others have a devastating medical and economic impact on humans. Still others, particularly many of the retrovirus-induced malignancies of rodents, were found originally in laboratory settings and provide excellent model systems for probing the biological and molecular mechanisms of carcinogenesis. Study of these disorders has provided critical information about the ways in which retroviruses induce disease. In addition, the mechanisms involved in these diseases have shed important light on the way in which similar conditions, lacking a retroviral etiology, arise. As noted in Chapter 1 studies of tumor induction by retroviruses provide the basis of much of modern molecular tumor biology. The discovery of viral oncogenes and the ways in which such genes can induce tumors provided both the intellectual framework and the molecular tools that have led to a deeper understanding of the mechanisms important for the development of all cancers. These investigations also helped lay the foundation for our current understanding of signal transduction and normal cellular growth control.
Article
We studied PVC-211 murine leukemia virus (MuLV) (1), a neuropathogenic variant of Friend MuLV, to determine its cellular tropism and distribution in the nervous system of infected rats and the factors that affected disease expression. Rats from five different strains and mice from 3 strains were inoculated intracerebrally or intraperitoneally from birth to 10 days of age and observed for signs of neurologic disease and tumors for 24 weeks. Nervous system pathology, MuLV gp70 expression, and virus production were evaluated weekly for 4 weeks after perinatal infection of Fisher (F344) rats. Blood-brain-barrier integrity and ultrastructure were evaluated in 21-day-old symptomatic infected rats. Microvessel and mixed glial cell cultures were prepared from brains of infected and uninfected 21-day-old F344 rats and evaluated for virus production, MuLV gp70 expression, and the presence of PVC-211 MuLV DNA. Tremor, ataxia, spasticity, and hindlimb weakness occurred in rats and mice as early as 3 weeks after neonatal infection. Severity, latency, and progression varied among mouse and rat strains but exposure to PVC-211 MuLV before 6 days of age was required for disease expression. Rapid PVC-211 MuLV replication in brain capillary endothelial cells (BCEC) early in the perinatal period was followed by widespread astrogliosis, neuropil vacuolation, and finally, neuronal degeneration in the spinal cord, brainstem, cerebellum, and subcortex. MuLV gp70 expression in vivo increased during infection, was restricted to BCEC, but was not associated with perivascular inflammatory infiltrates. BCEC cultured from microvessel preparations but not astrocytes or microglia in mixed glial cell cultures isolated from infected rats contained PVC-211 MuLV DNA, expressed MuLV gp70, and produced infectious virus. The rapid replication of PVC-211 MuLV that occurs in the nervous system of infected rodents is restricted to BCEC. These infected BCEC appear to play a critical role in initiating the astroglial response in this neurodegenerative process through mechanisms that remain to be defined.