Departments of Internal Medicine and of Integrative Biology, Pharmacology, and Physiology, The University of Texas Medical School at Houston, Houston, Texas 77030
![]() |
ABSTRACT |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
Nitric oxide synthase-2 (NOS2) is responsible for high-output nitric oxide production important in renal inflammation and injury. Using a yeast two-hybrid assay, we identified Rac2, a Rho GTPase member, as a NOS2-interacting protein. NOS2 and Rac2 proteins coimmunoprecipitated from activated RAW 264.7 macrophages. The two proteins colocalized in an intracellular compartment of these cells. Glutathione-S-transferase (GST) pull-down assays revealed that both Rac1 and Rac2 associated with GST-NOS2 and that the NOS2 oxygenase domain was necessary and sufficient for the interaction. [35S]methionine-labeled NOS2 interacted directly with GST-Rac2 in the absence of GTP, calmodulin, or NOS2 substrates or cofactors. Stable overexpression of Rac2 in RAW 264.7 cells augmented LPS-induced nitrite generation (~60%) and NOS2 activity (~45%) without measurably affecting NOS2 protein abundance and led to a redistribution of NOS2 to a high-speed Triton X-100-insoluble fraction. We conclude that Rac1 and Rac2 physically interact with NOS2 in activated macrophages and that the interaction with Rac2 correlates with a posttranslational stimulation of NOS2 activity and likely its spatial redistribution within the cell.
inducible nitric oxide synthase; guanosine 5'-triphosphatase; protein-protein interaction; lipopolysaccharide; phagocyte
![]() |
INTRODUCTION |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
THE FREE RADICAL NITRIC OXIDE (NO) plays critical roles in numerous physiological and pathophysiological processes. In the kidney, NOS2-derived NO appears to play important roles in the evolution of ischemic acute renal failure, glomerulonephritis, and acute tubulointerstitial nephritis (see Ref. 22 for review). NO is synthesized from L-arginine by the family of NO synthases (NOS). The three types of NOS share a bidomain structure consisting of an NH2-terminal oxygenase domain, which binds heme, tetrahydrobiopterin, and L-arginine, and a COOH-terminal reductase domain that binds FAD, flavin adenine mononucleotide (FMN), and NADPH (40). NOS1 and NOS3 are basally expressed in selected tissues and are principally activated by intracellular Ca2+ transients that promote calmodulin binding sufficiently to produce low amounts of NO for signaling purposes (4). In contrast, expression of NOS2 generally requires provocation with immunoactive agents, such as bacterial lipopolysaccharide (LPS) and/or certain cytokines that activate de novo synthesis of the enzyme. Once formed, NOS2 sustains catalysis because calmodulin is tightly bound even at resting levels of intracellular Ca2+. In humans, NOS2 expression is most readily detected in monocytes or macrophages from patients with infectious and/or inflammatory diseases (31, 49). The large amounts of NO generated by NOS2 have been implicated in inflammation, host defense, tissue injury, and the profound hypotension accompanying sepsis (22, 27).
Because of its importance in cell injury and inflammation, the mechanisms governing the biosynthesis and function of NOS2 have been the subject of intense investigation. A complex array of transcriptional and posttranscriptional regulatory controls on NOS2 expression and activity have been described, including influences on transactivation of the NOS2 gene, NOS2 mRNA stability, substrate and cofactor binding and availability, and dimerization (27). For NOS1 and NOS3, interactions with heterologous proteins in addition to calmodulin have been shown to influence the activity or intracellular distribution of these isoenzymes. For example, a 10-kDa peptide, protein inhibitor of neuronal NOS (PIN), was shown to bind and inhibit the activity of NOS1 in vitro (18). Similarly, caveolins-1 and -3 interact directly with NOS1 and NOS3 (19) and inhibit NOS3 activity (19). NOS1 has been shown to interact physically with several proteins bearing PDZ domains that influence NOS1 targeting to discrete membrane domains of excitable tissues (5, 6). In the case of NOS2, thus far kalirin (34) and NOS-associated protein-110 (35) have been bound to interact with NOS2 and inhibit its activity.
In the present study, we determined that Rac1 and Rac2, members of the Rho GTPase family, physically interact with NOS2. The Rac isoforms play central roles in several important cellular functions including cytoskeletal reorganization, gene expression, and assembly and activation of the phagocytic and nonphagocytic NADPH oxidase. We further demonstrate that the point of interaction for Rac2 is the NOS2 oxygenase domain and that overexpression of Rac2 augments NO production in immune-activated murine RAW 264.7 macrophage cells. This remarkable convergence of two pluripotent signaling pathways reveals the unexpected complexity of NOS2 regulation, identifies a new target protein for Rac1 and Rac2, and suggests the possibility of coordinate regulation of the NADPH oxidase and NO production in phagocytes.
![]() |
MATERIALS AND METHODS |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
Reagents. L-Glutamine, heat-inactivated FBS, penicillin-streptomycin, DMEM, and LipofectAMINE PLUS reagent were from Life Technologies. DMEM lacking phenol red, LPS from Escherichia coli O111:B4 and protease inhibitor cocktail were from Sigma. Glutathione-Sepharose 4B beads and RediPack columns, pGEX-3X, pGEX-5X, factor Xa, ECL reagents, and recombinant glutathione-S-transferase (GST) were from Amersham Pharmacia Biotech (Piscataway, NJ). Radiochemicals were purchased from Amersham. RNAzol II was acquired from TEL-TEST "B," The MATCHMAKER Two-Hybrid System 2, the Mouse Kidney MATCHMAKER cDNA library, the GAL4-activation domain hybrid cloning vector pGAD10, and pAS2-1 (GAL4-DNA binding domain hybrid cloning vector) were purchased from Clontech. Zeocin and pcDNA3.1/Zeo were acquired from Invitrogen. Mouse monoclonal anti-NOS2 antibody, anti-nitrotyrosine antibody, and an anti-nitrotyrosine-positive control were obtained from Transduction Laboratories. NG-monomethyl-L-arginine (L-NMMA) was purchased from Alexis Biochemicals. Mouse monoclonal anti-nitrotyrosine antibody was from Zymed (San Francisco, CA). Rabbit polyclonal IgGs specific for Rac1 and for Rac2 and protein A/G PLUS-agarose were purchased from Santa Cruz Biotechnologies. Oligonucleotides were custom synthesized by Genosys, the BCA protein assay kit was from Pierce Chemical, and the LipofectAMINE PLUS reagent was purchased from Promega.
Plasmids and constructs. A fusion of a segment of murine NOS2 with the GAL4-DNA binding domain of pAS2-1 was constructed by subcloning an NcoI-NcoI fragment of the NOS2 cDNA (encoding M325-H934) into these restriction sites of pAS2-1 to yield the recombinant molecule pAS2-NOS2325-934. An expression plasmid for full-length NOS2 cloned into pCDNA3.1/Neo (Invitrogen) has been previously described (23). To generate fusions with GST, cDNA inserts encoding murine Rac2, human Rac1 (a gift from Dr. J. David Lambeth, Emory University), or various portions of murine NOS2 (M1-L1144, M1-E498, K499-L1144) were PCR amplified from cloned cDNAs or, in the case of Rac2, cDNA prepared from LPS-treated RAW 264.7 cells. The cDNAs were subcloned into convenient restriction sites in pGEX-3X or pGEX-5X vectors to maintain appropriate reading frames. For construction of a Rac2 expression plasmid, the entire coding region of Rac2 (nucleotides 35-614, GenBank Accession No. X53247) was PCR amplified from RAW 264.7 cell cDNA by using primers based on the published sequence. The Rac2 cDNA was subcloned into the BamH1 and EcoR1 sites of the mammalian expression vector pcDNA3.1/Zeo to create pcDNA3.1/Rac2-Zeo. All constructs were analyzed by DNA sequencing to confirm their authenticity and reading frames within the vectors.
Yeast two-hybrid assay.
The MATCHMAKER Two-Hybrid System 2 and the mouse kidney MATCHMAKER cDNA
library were used to identify proteins interacting with NOS2.
Saccharomyces cerevisiae strain CG-1945, which is
auxotrophic for tryptophan and leucine and contains lacZ and
HIS3 reporter genes, was grown at 30°C in either YPD
medium (1% yeast extract, 2% peptone, and 2% glucose) or minimal
synthetic dropout (SD) medium (0.5% yeast nitrogen base without amino
acids, 2% glucose, and 1% desired amino acid dropout solution).
pAS2-NOS2325-934 and pGAD10-kidney cDNA library
fusions were introduced into the CG-1945 yeast reporter strain by the
lithium acetate transformation procedure (2).
Cotransformants were grown on SD plates containing 5 mM
3-amino-1,2,4-triazole (to reduce background from "leaky" HIS3 expression) and lacking histidine, tryptophan, and
leucine to select for cotransformants in which the pAS2-1- and
pGAD10-hybrid proteins interact. The resultant colonies were then
screened for -galactosidase activity by a colony lift filter assay
(2). Insert cDNAs of clones demonstrating activation of
the HIS3 and lacZ marker genes were amplified by
PCR of total yeast DNA by using flanking primer sequences derived from
pGAD10. Selected clones were then sequenced on both strands by using an
automated method. In control experiments, the GAL4-DNA binding domain
and GAL4-activation domain fusion constructs were confirmed not to activate reporter gene transcription by themselves, and GAL4-activation domain fusion constructs were confirmed not to activate transcription when combined with the unrelated pLAM 5'-1, which encodes a GAL4-DNA binding domain-human lamin C hybrid in pAS2-1.
Coimmunoprecipitation. RAW 264.7 cells were incubated with basal medium or stimulated with LPS for 16 h to induce NOS2 expression. Lysates of the cells were prepared with ice-cold RIPA buffer (PBS containing 1% Nonidet P-40, 0.5% sodium deoxycholate, 0.1% SDS, 3% protease inhibitor cocktail). One milliliter of lysate was precleared with 10 µl normal rabbit serum together with 20 µl protein A/G PLUS-agarose for 1 h at 4°C. The lysates were centrifuged at 1,500 rpm at 4°C for 5 min to remove insoluble material. Two hundred and fifty microliters of the resulting supernatant were subjected to immunoprecipitation with anti-NOS2 mouse monoclonal antibody (2.5 µg; or with an equal amount of mouse IgG) or with rabbit polyclonal anti-Rac2 (1 µg; or with an equal amount of rabbit IgG) for 1 h at 4°C together with 20 µl protein A/G PLUS-agarose for 2 h at 4°C. The immunoprecipitates were collected by centrifugation at 2,500 rpm for 5 min at 4°C, and washed four times with 1 ml RIPA buffer. Proteins precipitated by anti-NOS2 or mouse IgG were separated by SDS-PAGE, electroblotted to Hybond-ECL membranes, and immunoblotted with anti-Rac2. Proteins precipitated by anti-Rac2 or rabbit IgG were immunoblotted with anti-NOS2. Bound antibody was visualized by the ECL chemiluminescent detection system, using sheep anti-mouse (for NOS2 and mouse IgG) or goat anti-rabbit (for Rac2 and rabbit IgG) IgG-peroxidase conjugate and autoradiography. The relative amount of NOS2 and Rac2 coprecipitated was grossly estimated by running 50- and 100-µg aliquots of the total lysate (not subjected to the immunoprecipitation procedure) on the final immunoblot of the coprecipitates and comparing the relative abundance of the bands.
In vitro translation. NOS2 was transcribed and translated from pCDNA3.1/Neo-NOS2 in the presence of [35S]methionine by using T7 RNA polymerase and a troponin T-coupled reticulocyte lysate kit (Promega) by methods previously described (24).
GST pull-down assays.
Three GST-fusion proteins, constructed to contain full-length NOS2
(M1-L1144), the isolated oxygenase domain
(M1-E498), or the isolated reductase domain
(K499-L1144), were purified from sonicates of
isopropyl -D-thiogalactoside-induced HB101 bacterial
cells according to the manufacturer's (Amersham Pharmacia Biotech)
instructions. For in vitro binding assays using cell lysates, ~4 µg
of GST-fusion protein constructs were added to RediPack
glutathione-Sepharose 4B columns. The columns were washed with 20 ml
ice-cold PBS. Two hundred microliters of total cell lysates of
LPS-treated RAW 264.7 cells prepared in RIPA buffer as above were used
as a source of NOS2 protein and added to the column, followed by
incubation with continuous rotation for 90 min at 4°C. The columns
were then washed with 30 ml of ice-cold PBS containing 0.1% Tween
followed by a final 3-ml wash of PBS, with eluates collected after each
wash for analysis by SDS-PAGE. Factor Xa (50 µl in 450 µl of PBS
containing 0.5% SDS) was added to the column for an overnight
incubation at room temperature to cleave the bait and prey proteins
from the GST tag. The final washes and the factor Xa-cleaved eluates
were analyzed by SDS-PAGE and immunoblotting with anti-Rac1 or
anti-Rac2 antibodies.
Cell culture and transfection.
The murine macrophage-like cell line RAW 264.7 was grown in complete
medium (DMEM supplemented with 2 mM L-glutamine, 100 U/ml
penicillin, 100 U/ml streptomycin, and 10% FBS) at 37°C in a
humidified incubator with 5% CO2. LPS (5 or 10 µg/ml)
with or without interferon (IFN)- (200 U/ml) was added to the cells
as indicated in the text and figure legends. For nitrite assays, DMEM
lacking phenol red was used to formulate the complete medium. For
stable transfections, plasmid preparations were made by using an
endotoxin-free Plasmid Maxi prep kit (Qiagen). Subconfluent RAW 264.7 cells were transfected with 4.5 µg/ml pcDNA3.1/Zeo (as a vector
control) or pcDNA3.1/Rac2-Zeo with the LipofectAMINE PLUS reagent
according to the manufacturer's protocol to yield RAW-Zeo and RAW-Rac2
cell lines, respectively. Forty-eight hours after transfection,
selection was initiated with complete medium containing 300 µg/ml
zeocin. The zeocin-containing medium was replaced every 3-4 days
until individual resistant colonies were isolated and established in
culture as individual lines. Because RAW 264.7 cells endogenously
express Rac2, the RAW-Rac2 lines were screened for successful
expression of vector-derived Rac2 mRNA by Northern analysis by using a
32P-labeled StuI-PvuII fragment of
pcDNA3.1/Rac2-Zeo (nucleotides 525-614 of Rac2 and nucleotides
951-1271 of pcDNA3.1/Zeo) that hybridizes to vector-derived, but
not endogenous, Rac2. Six representative lines of the highest
expressing RAW-Rac2 clones as well as the RAW-Zeo clones were
maintained in zeocin-containing medium, frozen after one to three in
vitro passages, and studied in further detail.
Deconvolution immunofluorescence microsopy. RAW 264.7, RAW-Zeo, and RAW-Rac2 cells grown on glass coverslips were stimulated with LPS for 16 h to induce NOS2 expression. The cells were fixed in buffered 3.7% formaldehyde for 20 min, permeabilized with PBS containing 0.5% Triton X-100, and blocked in PBS containing 10% goat serum at 37°C for 1 h. The cells were incubated for 1 h with anti-Rac2 and anti-NOS2 antibodies together with FITC-conjugated phalloidin diluted in PBS+10% normal goat serum at 37°C for 1 h. As a negative control, primary antibody was omitted from the immunostaining procedure. After 3 washes in PBS containing 0.05% Tween-20, the cells were incubated with Cy5-congugated goat anti-rabbit IgG (Molecular Probes) and Texas red-conjugated goat anti-mouse IgG (Molecular Probes) diluted 1:500 in PBS containing 10% normal goat serum and 0.05% Tween 20. After final washes in PBS containing 0.05% Tween, the cells were overlaid with Elvanol (DuPont) and mounted. The cells were then imaged using an Olympus IX70 inverted epifluorescence microscope. Data sets were acquired by using a mercury short-arc lamp and stored in digital format by using a cooled charge-coupled device (CCD) camera (Applied Precision, Delta Vision System). The data sets were then transferred to a Silicon Graphics workstation for deconvolution and three-dimensional reconstruction (9, 28). Delta Vision System SoftWoRx (Applied Precision, Issaquah, WA) software was used to deconvolve 0.1-µm optical sections before reconstruction. The data sets were then imported into Imaris 3 (Bitplane, Zurich, Switzerland) for digital image restoration and shadowing. For each cell, a series of 30-50 individual optical sections of 100-nm thickness in the vertical axis was obtained. Colors were arbitrarily computer-assigned to individual fluors (Cy5 = green, Texas red = red; FITC = blue).
Nitrite assays. Nitrite, the stable metabolite of NO, was measured in culture supernatants by a modification of the Griess reaction as previously detailed (30). Standard curves were generated from sodium nitrite. Triplicate determinations were performed for each condition and represent a single observation (n). The data were normalized to protein content as determined by the BCA assay.
NOS activity assay. Total cell homogenates were prepared at 4°C in homogenization buffer (25 mM Tris · HCl, pH 7.4, 1 mM EDTA, 1 mM EGTA). The NOS activity of the lysates was measured for 30 min at room temperature by the formation of L-[3H]citrulline from L-[3H]arginine (30 mM) by using the components of a NOSdetect Assay Kit. Parametric studies indicated that these conditions were within the linear range of the reaction. The reaction mixture contained 25 mM Tris · HCl (pH 7.4), 3 mM tetrahydrobiopterin, 25 nM calmodulin, 0.6 mM CaCl2, 1 µM FAD, 1 µM FMN, and 1 mM NADPH, with or without 1 mM NG-nitro-L-arginine. L-citrulline was eluted from the resin supplied with the kit, and L-[3H]citrulline was quantified by liquid scintillation counting. The data were normalized to protein content as determined by the BCA assay.
Superoxide anion assays.
Superoxide anion (O for 16 h at 37°C. The cells were then collected, and
106 cells were pelleted and resuspended in 100 µl
superoxide anion medium (Stratagene). Luminol and the proprietary
enhancer were added to a final concentration of 100 and 125 µM,
respectively. The luminescence, indicative of the superoxide anion
produced, was measured in a Turner Systems 20/20 luminometer. The data
were normalized to the protein content of the samples.
RT-PCR. Isolation of total RNA and RT-PCR for NOS1 and NOS3 was performed as described in previous work from our laboratory (29). For amplification of NOS1, primers corresponded to nucleotides 2462-2616 (forward) and 3037-3062 (reverse) of the rat sequence (3). NOS3 amplification was performed with primers directed to nucleotides 1889-1902 (forward) and 2594-2616 (reverse) of the murine sequence (13).
Triton X-100 fractionation. RAW-Rac2 and RAW-Zeo clonal cell lines grown on 100-mm2 culture dishes and treated with vehicle or LPS for 16 h were washed twice with ice-cold PBS and lysed for 10 min on ice in lysis buffer [125 mM NaCl, 25 mM Tris · HCl (pH 8.0), 10 mM EDTA, 10 mM NaF, 1% Triton X-100, protease inhibitor cocktail] or cytoskeleton-stabilizing buffer (20 mM Tris · HCl, pH 7.4, 3 mM MgCl2, 8% sucrose, 0.5% Triton X-100, protease inhibitor cocktail) (47). A "total lysate" sample was taken and, after boiling in Laemmli sample buffer, analyzed on Western blots. All fractionation steps were performed at 4°C. The remaining Triton X-100 lysate was centrifuged at 15,000 g for 10 min. The resulting pellet was washed once in lysis or cytoskeleton-stabilizing buffer, as appropriate, and taken as the "low-speed Triton X-100-insoluble fraction." The supernatant was collected and centrifuged at 200,000 g for 100 min. The resulting pellet, which was washed once with lysis buffer or cytoskeleton-stabilizing buffer, as appropriate, represents the "high-speed Triton X-100-insoluble fraction," whereas the supernatant represents the "high-speed soluble fraction." For analysis by SDS-PAGE and Western blotting, all samples were boiled in Laemmli sample buffer.
Immunoblotting for nitrosylated proteins.
Immunoblots of cell lysates were prepared from RAW-Zeo and RAW-Rac2
cells that had been treated with LPS+IFN- overnight as previously
described (14). A commercially available positive control
containing nitrated proteins was used. As an additional positive
control, lysates of RAW 264.7 cells that had been exposed 20 times to
exogenously applied 5 mM peroxynitrite (a gift from Drs. Yasu Irie and
Ferid Murad) were used. The blots were probed with two different
anti-nitrotyrosine antibodies and sheep anti-mouse or goat anti-rabbit
IgG-peroxidase conjugate as appropriate.
![]() |
RESULTS |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
NOS2 interacts with Rac1 and Rac2 in vivo. To identify proteins that physically interact with NOS2, a yeast two-hybrid screen was performed. A fusion gene, pAS2-NOS2325-934, containing the GAL4-DNA binding domain of pAS2-1 fused to cDNA encoding M325-H934 of murine NOS2, was constructed. Yeast CG-1945 cells expressing this construct were used to screen a library of GAL4-activation domain-tagged mouse kidney cDNA. Of the ~3,000,000 clones screened, 9 transformants exhibited pAS2-NOS2325-934-dependent activation of both the HIS3 and lacZ markers in CG-1945 cells. Sequence analysis of one of the nine pAS2-NOS2325-934-derived clones revealed it to contain nucleotides 363-899 of murine Rac2 in-frame with the GAL 4-activation domain sequence. This sequence of Rac2 corresponds to the coding sequence beginning at codon 110 and extending into the 3'-untranslated region.
To confirm the results of the two-hybrid experiments by an independent biochemical method and to establish the occurrence of the Rac2-NOS2 interaction in mammalian phagocytic cells, a coimmunoprecipitation strategy was used. RAW 264.7 macrophage cells were treated with LPS for 16 h to induce expression of NOS2. Cell lysates were then prepared and immunoprecipitated with either anti-Rac2 or anti-NOS2 antibodies. Proteins precipitated by anti-Rac2 were separated by gel electrophoresis, transferred to nylon membranes, and immunoblotted with anti-NOS2. In the reciprocal experiment, anti-NOS2 immunoprecipitates were immunoblotted with anti-Rac2. Coprecipitation of the two proteins was taken as evidence for a direct and stable interaction between them. Figure 1A shows the results of the coimmunoprecipitation experiments. Immunoblots of the anti-Rac2 immunoprecipitates with the anti-NOS2 antibody revealed the expected ~130 NOS2 monomer. Similarly, immunoblots of the anti-NOS2 immunoprecipitates probed with the anti-Rac2 antibody revealed the presence of the ~21 kD Rac2 protein. A gross comparison of the relative abundance of the immunoprecipitated proteins vs. the input amount revealed that the association appeared to be roughly stoichiometric in that ~20% of the total NOS2 and Rac2 was coimmunoprecipitated. Because some of the protein A/G PLUS-agarose is lost in the multiple washes during the immunoprecipitation procedure, this value probably underestimates the true amount of associated proteins. In both cases, immunoprecipitation with non-immune IgG failed to precipitate the Rac2 or NOS2 proteins (Fig. 1B). As an additional control, immunoprecipitation with the anti-Rac2 and blotting with anti-NOS2 failed to detect NOS2 in RAW 264.7 cells not pretreated with LPS (Fig. 1C). These results confirmed the association of full-length Rac2 with NOS2 in LPS-stimulated RAW 264.7 cells.
|
The NOS2-Rac2 interaction in vitro is direct, GTP independent, and
does not require calmodulin, NOS substrates, or cofactors.
In the yeast two-hybrid system and coimmunoprecipitation assays,
protein-protein interactions may be direct or mediated by a bridging
endogenous protein. Therefore, Rac2 was produced and purified as a
GST-fusion protein, and [35S]methionine-labeled NOS2 was
translated in vitro to test their ability to bind directly in vitro in
a GST pull-down assay conducted in the absence of calmodulin or NOS
substrates or cofactors. In addition, to test the GTP dependence of the
interaction, the GST-Rac2 fusion proteins were prebound with GDPS,
the nonhydrolyzable GTP
S, or were generated nominally free of
nucleotides and subjected to the pull-down assay. As shown in Fig.
2, [35S]methionine-labeled
NOS2 directly interacted with GST-Rac2, but not GST alone, under each
of these conditions.
|
The NH2-terminal oxygenase domain of NOS2 is necessary
and sufficient to mediate the association of NOS2 with Rac2.
To determine the regions of NOS2 to which Rac2 binds, GST-fusion
proteins encoding the isolated oxygenase and reductase domains of NOS2
as well as full-length NOS2 were used in a GST pull-down assay with
lysates of LPS-treated RAW 264.7 cells. As shown in Fig.
3, Rac2 in the lysates bound to the
GST-full-length NOS2 fusion protein (in agreement with the experiment
in Fig. 1B) and the GST-NOS2 oxygenase domain fusion
protein. However, no interaction with Rac2 was observed with the
GST-NOS2 reductase domain fusion protein (Fig. 3). Because the NOS2
oxygenase construct lacks the binding sites for calmodulin and flavin
nucleotides, this result indicates that the latter are not required for
the interaction with Rac2. This result substantiates the findings
presented in Fig. 2. The fact that Rac2 did not interact with the
GST-NOS2 reductase domain fusion protein also demonstrates specificity of the interactions observed with the GST-full-length NOS2 fusion protein and the GST-NOS2 oxygenase domain fusion protein.
|
Overexpression of Rac2 stimulates NOS2 activity. To determine whether Rac2 might influence NOS2 expression or activity, we stably transfected RAW 264.7 cells with a Rac2 expression plasmid or the parent vector pcDNA3.1/Zeo. To confirm that the transfection procedure did not somehow result in de novo expression of NOS1 or NOS3 in the resultant cell lines, RT-PCR analysis for these isoforms was performed by using isoform-specific primers and RNAs prepared from normal and LPS-treated wild-type RAW-Zeo cells, RAW-Rac2 cells, and mouse kidney (as a positive control). NOS1 and NOS3 transcripts were successfully amplified from mouse kidney RNA, but not from the cell lines (data not shown), indicating that NOS2 is the sole NOS isoform expressed in the transfected cell lines treated with LPS.
The RAW-Rac2 and RAW-Zeo cell lines were then examined for their ability to produce NO and NOS2 protein under basal conditions and in response to LPS. Under basal conditions, both cell lines produced comparable, but negligible, nitrite (Fig. 4A), indicating that overexpressed Rac2, in the absence of immune stimuli, cannot provoke NO generation. Figure 4A shows that LPS promoted, as expected, high-level nitrite production in the RAW-Zeo cells, but even greater (~60%) nitrite levels were observed in the RAW-Rac2 cells. Because nitrite production from the intact cell may be influenced by changes in the availability of NOS2 cofactors or substrate, the NOS2 activity of RAW-Zeo and RAW-Rac2 cell homogenates was measured by the L-[3H]arginine to L-[3H]citrulline conversion assay in the presence of fixed and saturating concentrations of cofactors and substrates. In response to LPS, the RAW-Rac2 cells exhibited NOS2 activity that was roughly 45% higher than that of the RAW-Zeo cells (Fig. 4B), in close agreement with the nitrite production of the intact cells. The enhanced LPS-induced nitrite production and NOS2 activity of the RAW-Rac2 cells appeared to be principally the result of a posttranslational event, because immunoblots of lysates prepared from the cells revealed NOS2 protein amounts comparable to that of the LPS-treated RAW-Zeo cells (Fig. 4C).
|
NOS2 and Rac2 colocalize in RAW 264.7 cells.
To determine whether NOS2 and Rac2 colocalize in macrophages, the
subcellular distribution of NOS2 and Rac2 was analyzed by immunofluorescence deconvolution microscopy (9, 28) of
LPS-treated RAW 264.7 cells. Both NOS2 and Rac2 immunofluorescence were
found in a perinuclear location (judged by DAPI staining of the nuclei; not shown) as well as in more peripheral cytoplasmic sites. An overlapping subcellular localization was observed between a pool of
NOS2 and Rac2 (Fig. 5). In control
experiments, no staining was evident in the absence of primary antibody
for either NOS2 or Rac2 (not shown).
|
Overexpression of Rac2 promotes NOS2 translocation to a high-speed
Triton X-100-insoluble fraction.
Because Rac2 has been shown to influence the organization of the
cytoskeleton and the spatial distribution of NADPH oxidase components
in activated phagocytes, we hypothesized that overexpression of Rac2
might influence the subcellular distribution of NOS2 and thereby
provide a molecular basis for observations of soluble and particulate
forms of NOS2. As an initial approach to this question, we used Triton
X-100 fractionation of lysates from RAW-Zeo and RAW-Rac2 cell lines to
examine this possibility. This method has been commonly used to study
Rac association with the components of the NADPH oxidase. Two separate
buffer systems, lysis buffer and cytoskeleton-stabilizing buffer, were
used. The results below were comparable regardless of the buffer system
used to prepare the lysates. The low-speed Triton X-100-insoluble
pellet has been reported to contain nuclei, large cytoskeletal
networks, and caveolae. The high-speed Triton X-100-insoluble pellet
contains submembranous cytoskeleton complexes and associated signaling
molecules (12, 42), including Rac2 (10). In
agreement with these studies, Rac2 was found in the high-speed Trixon
X-100-insoluble pellet (and the low-speed supernatant from which the
high-speed fractions were derived) but not high-speed supernatant in
control and LPS-treated (16 h) RAW-Zeo and RAW-Rac2 cells (Fig.
6A). Rac2 was also
consistently more abundant in the LPS-treated cells. NOS2 was
undetectable under basal conditions in RAW-Zeo and RAW-Rac2 cell lines
(Fig. 6A). After LPS treatment for 16 h, NOS2 was
apparent in the low-speed Triton X-100-soluble fraction and both the
high-speed Triton X-100-soluble and -insoluble fractions (Fig.
6B). Similar results were obtained in untransfected RAW
264.7 cells after LPS treatment (data not shown). In the LPS-treated
RAW-Rac2 cells, however, NOS2 was detected exclusively in the low-speed
Triton X-100-soluble fraction and the high-speed Triton X-100-insoluble
fraction derived from it (Fig. 6B). These results indicate
that overexpression of Rac2 promotes NOS2 translocation to and/or
retention in the high-speed Triton X-100-insoluble fraction. However,
deconvolution immunofluorescence microscopy, which has comparable
resolution to confocal microscopy, did not demonstrate an unequivocal
difference in the subcellular labeling of the NOS2 and Rac2 proteins in
the LPS-treated RAW-Zeo and RAW-Rac2 cell lines (data not shown).
|
Effects of overexpression of Rac2 on superoxide anion production.
To assess whether overexpression of Rac2 alters superoxide anion
production in macrophages, a chemiluminescence assay was used to
measure superoxide anion produced by resting or LPS+IFN--stimulated RAW-Zeo and RAW-Rac2 cells. As shown in Table
1, RAW-Rac2 cells generated greater
levels of superoxide anion than RAW-Zeo cells, suggesting, for the
first time, that Rac2 overexpression drives superoxide production by
the NADPH oxidase. LPS+IFN-
treatment resulted in a significant
reduction in the amount of superoxide anion produced by both groups of
cell lines. Because LPS+IFN-
treatment promotes NO production in
these cells, and because NO chemically inactivates superoxide anion by
forming peroxynitrite, we reasoned that the lower levels of superoxide
anion in the activated RAW-Rac2 and RAW-Zeo cell lines might reflect a
quenching of superoxide by NO. Accordingly, we assayed superoxide
production in LPS+IFN-
-treated cells when NOS2 was inhibited with
L-NMMA. With NOS2 inhibited, superoxide anion production
was enhanced in the LPS+IFN-
-treated RAW-Zeo cells and to a far
greater extent in the RAW-Rac2 cells (Table 1). If the
L-NMMA-dependent component of superoxide anion production
represents the superoxide anion inactivated by NO, this result suggests
that the RAW-Rac2 cells form far greater levels of peroxynitrite than
the RAW-Zeo cells after exposure to LPS+IFN-
. Alternatively, high
concentrations of NO peroxynitrite have been reported to inactivate the
NADPH oxidase in polymorphonuclear leukocytes (26).
|
![]() |
DISCUSSION |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
In this study, we report several novel findings important to an understanding of NOS2 biology, Rac1 and Rac2 signaling targets, and macrophage function. By demonstrating the physical interaction of NOS2 with Rac1 and Rac2 in activated murine macrophages, we provide new examples of heterologous proteins other than calmodulin that physically associate with NOS2 and identify NOS2 as a new target for the Rac isoforms. Our demonstration that overexpression of Rac2 augments NOS2-generated NO production and NOS activity in LPS-treated RAW 264.7 cells establishes not only the functional importance of the NOS2-Rac2 interaction but also suggests a novel posttranslational mechanism for controlling NO synthesis. In vitro binding studies showed that Rac2 binds to the oxygenase, but not the reductase domain of NOS2, that the interaction is direct, and that it can occur in vitro in the absence of calmodulin, heme, and NOS2 substrates and cofactors. Finally, our finding that NOS2 exclusively distributes to a high-speed Triton X-100-insoluble fraction in RAW 264.7 cells overexpressing Rac2 suggests that the interaction influences the spatial distribution of NOS2 in the cell, apparently in such a way that optimizes LPS-induced NO release. Given the diverse cell processes in which both NOS2 and the Rac isoforms participate, the physical association of these proteins in vivo suggests the possibility that the interaction may play important biological roles.
Rac2 is principally expressed in cells of myeloid origin, where it participates in the formation of the NADPH oxidase complex in phagocytic cells (8), and it is also believed to be a component of the NAD(P)H oxidase in vascular smooth muscle cells (33). Its close structural relative, Rac1, is more widely distributed among cell types and has been implicated in diverse cell processes, such as the organization of the actin cytoskeleton, protein kinase cascades, transcriptional activation, and cell cycle progression (15). The Rac proteins are believed to play a role in membrane ruffling and phagocytosis in macrophages. The NADPH oxidase of phagocytes generates superoxide anion and other reactive oxygen species such as H2O2, HOCl, and OH that are cytotoxic to invading pathogens (8). In resting cells, the phagocytic NADPH oxidase is composed of membrane-associated flavocytochrome b559 and several cytosolic proteins, including p47phox, p67phox, p40phox, and Rac2 or, less commonly, Rac1. When the phagocytic NADPH oxidase is activated to generate superoxide anion, p47phox, p67phox, and Rac2 translocate to a detergent-insoluble subcellular fraction (10) and become stably associated with the plasma membrane (8). Similarly, we found that overexpression of Rac2 in LPS-treated RAW 264.7 cells promotes redistribution of NOS2 from high-speed Triton X-100-soluble and -insoluble fractions exclusively to the Triton X-100-insoluble fraction. Whether Rac2, through its physical association, ushers NOS2 from the high-speed Triton X-100-soluble compartment to the high-speed Triton X-100-insoluble compartment, facilitates its retention in the latter compartment, or acts indirectly to promote NOS2 translocation from one compartment to the other will require additional study.
In suggesting that the distribution of NOS2 within the cell might be
regulated and that Rac2 is involved in this process, the present data
extend earlier reports of soluble and particulate forms of NOS2
(16, 17, 38, 48). Hiki and co-workers (17) demonstrated nitrite-producing activity in the high-speed pellet of
sonicated peritoneal macrophages from Calmette-Guerin bacillus-treated rats. A portion of the particulate activity was Triton X-100 insoluble. Hecker et al. (16), studying IFN--treated J774.2
macrophage-like cells, recovered roughly one-half of the cGMP-elevating
activity in a 200,000-g pellet, but the molecular identity
and detergent solubility of the activity were not established. Schmidt
et al. (38) found that roughly one-third of NOS activity
in LPS and IFN-
-treated RAW 264.7 cells distributed to a
105,000-g, KCl-washed pellet. The majority of the
particulate activity remained insoluble in
3-[(3-cholamidopropyl)dimethylammonio]-1-propanesulfonate. Finally, Vodovotz and co-workers (48), using activated
primary macrophages, found approximately one-half of the NOS2 activity and immunoblotted NOS2 protein in a 100,000-g, KCl-washed
pellet. Of this particulate activity, about one-half was Triton X-100 insoluble. Immunoelectron microscopy demonstrated NOS2 immunoreactivity in 50- to 80-nm vesicles that were not labeled by lysosomal or peroxisomal antibodies. Spatial restrictions of NOS2 within the cell
presumably serve functional purposes. Because NO and
O
), is favored with the prospect of
injury not only to pathogens but also to the host cell.
Compartmentalizing NO production may represent a means for the cell to
optimize microbial killing or inhibition of viral replication and
minimize exposure of host cell components to these toxic molecules.
Whether Rac2 ushers a pool of NOS2 to the NADPH oxidase remains to be
clarified. Further studies will also be needed to determine whether the
NOS2-Rac2 interaction alters the GTPase activity of the Rac proteins.
Furthermore the biochemical composition and ultrastructural correlate
of this compartment remain to be determined, because we did not discern a distinct difference in intracellular locale at the resolution of
deconvolution immunofluorescence microscopy.
It is intriguing to speculate that the association of Rac2 with both
the NADPH oxidase complex and NOS2 may place superoxide anion and NO
production in close proximity to facilitate the formation of
peroxynitrite for microbicidal effects or, in contrast, to promote the
NO-mediated inactivation of the NADPH oxidase (26) to
limit damage to host tissues. We found that Rac2-overexpressing macrophages produced more superoxide anion than did controls under basal conditions, suggesting that Rac2 may be rate limiting for NADPH
oxidase activity. In addition the Rac2-overexpressing cells generated
little superoxide anion when NOS2 was induced by LPS+IFN- but high
amounts when NOS2 was inhibited with L-NMMA. These data suggest that the superoxide anion produced by the LPS+IFN-
-treated macrophages rapidly reacts with NOS2-generated NO to form
peroxynitrite. However, because the relative amounts and patterns of
proteins immunoreactive with two different anti-nitrotyrosine
antibodies, an index of peroxynitrite formation, did not remarkably
differ between the LPS+IFN-
-treated Raw-Rac2 and the Raw-Zeo cells, it remains possible that NO inactivates the NADPH oxidase in these cells to limit superoxide anion production, as has been reported in
other cell types (26).
The Rac proteins generally function as molecular switches, cycling between an active GTP-bound and an inactive GDP-bound state. In most examples, the active GTP-bound Rac binds to target effector proteins that relay cell signals. We found that recombinant Rac2 bound to NOS2 in vitro in the GDP-bound state and in the nominal absence of guanine nucleotides (Fig. 2). GTP-independent interactions of Rac isoforms in vitro have also been reported for other targets, such as Sra-1 (20), type I phosphatidylinositol-4-phosphate 5-kinase (45), and diacylglycerol kinase (44). Although we did not observe a consistent difference in the amount of in vitro translated NOS2 bound to GTP- or GDP-bound GST-Rac2 (Fig. 2), it remains possible that GTP-GDP cycling may influence the kinetics or stability of the interaction in vivo.
The fact that NOS2 interacted with both Rac1 and Rac2 in the GST pull-down assays (Fig. 1B) suggests that NOS2 associates with regions common to the Rac isoforms. This common sequence region would include all but the extreme COOH termini of the Rac proteins. The identification of the minimal region(s) within the NOS2 oxygenase domain to which Rac1 and Rac2 binds will be important, because it will not only provide insights into NOS2 bioregulation but also may help to identify other targets for the Rac isoforms and related Rho GTPases. Because the Rho GTPases participate in numerous cell processes, considerable effort has been made to identify molecular targets of these proteins. Several protein motifs recognized by Rho GTPases have been characterized in various target proteins, including the CRIB motif (7), REM-1 (36), and POSH (41), but these motifs are absent from the oxygenase domain of NOS2. Rac2 has been shown to bind in vitro to the COOH terminus of p67phox (11) and tetratricopeptide repeat motifs in the NH2-terminal region of p67phox (21), but there is no significant homology between p67phox and NOS2. Similarly, other known Rac targets, such as PAK (43), Wiskott-Aldrich syndrome protein (37), POR1 (46), IQGAP1 (25), and specifically Rac1-associated protein (20), also share no sequence homology with NOS2. Thus the Rac binding motif in NOS2, once defined, will be novel. Further definition of the protein domains of NOS2 and Rac2 that interact will require additional studies, as will determination of whether other members of the NOS and Rho GTPase families can associate with one another. Future studies to investigate the functional and regulatory aspects of NOS2-Rac relationships should provide fresh insights into the cell biology of not only phagocytic cells but also the numerous tissues and cell types in which these proteins are expressed in physiological and pathophysiological states.
Biosynthesis of NOS2 is principally regulated at the level of gene transcription. The posttranslational control of NOS2 activity has received little investigative attention. Inferential evidence suggests that tyrosine phosphorylation of NOS2 in activated RAW 264.7 cells may increase NOS2 activity by a posttranslational mechanism. Exposure of activated RAW 264.7 cells to a tyrosine phosphatase inhibitor significantly increased the level of NOS2 tyrosine phosphorylation and NOS2 activity (32). The present report suggests a second example of a posttranslational mechanism that augments activity of the enzyme. The fact that NOS2 activity in the RAW-Rac2 cell homogenates, under conditions in which the concentrations of substrates and cofactors were fixed and saturating, was greater than that of RAW-Zeo cells suggests that Rac2 influences the activity of the enzyme via a direct mechanism rather than by increasing the availability of substrate or cofactors. This conclusion is supported by knowledge that intracellular levels of heme and tetrahydrobiopterin increase coincident with NOS2 induction in activated RAW 264.7 cells (1) and that further supplementation of these prosthetic groups or the substrate L-arginine does not increase NOS2 activity or dimerization in these cells (1). Further studies will be required to decipher the specific mechanism by which Rac2 enhances NOS2 activity.
![]() |
ACKNOWLEDGEMENTS |
---|
We thank Dr. J. David Lambeth, Emory University, for the gift of human Rac1 cDNA.
![]() |
FOOTNOTES |
---|
This work was supported by National Institutes of Health Grants DK-50745 and GM-20529 (to B. C. Kone) and the Department of Defense "DREAMS" Center. It was conducted during Dr. Kone's tenure as an Established Investigator of the American Heart Association.
Address for reprint requests and other correspondence: B. C. Kone, Div. of Renal Diseases and Hypertension, Depts. of Internal Medicine and of Integrative Biology, Pharmacology, and Physiology, The Univ. of Texas Medical School at Houston, 6431 Fannin, MSB 4.138, Houston, TX 77030 (E-mail: Bruce.C.Kone{at}uth.tmc.edu).
The costs of publication of this article were defrayed in part by the payment of page charges. The article must therefore be hereby marked "advertisement" in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.
Received 18 October 2000; accepted in final form 20 April 2001.
![]() |
REFERENCES |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
1.
Albakri, QA,
and
Stuehr DJ.
Intracellular assembly of inducible NO synthase is limited by nitric oxide-mediated changes in heme insertion and availability.
J Biol Chem
271:
5414-5421,
1996
2.
Bartel, PL,
and
Fields S.
Analyzing protein-protein interactions using two-hybrid system.
Methods Enzymol
254:
241-263,
1995[ISI][Medline].
3.
Bredt, DS,
Hwang PM,
Glatt CE,
Lowenstein C,
Reed RR,
and
Snyder SH.
Cloned and expressed nitric oxide synthase structurally resembles cytochrome P-450 reductase.
Nature
351:
714-718,
1991[ISI][Medline].
4.
Bredt, DS,
and
Snyder SH.
Nitric oxide: a physiologic messenger molecule.
Annu Rev Biochem
63:
175-195,
1994[ISI][Medline].
5.
Brenman, JE,
Chao DS,
Gee SH,
McGee AW,
Craven SE,
Santillano DR,
Wu Z,
Huang F,
Xia H,
Peters MF,
Froehner SC,
and
Bredt DS.
Interaction of nitric oxide synthase with the postsynaptic density protein PSD-95 and 1-syntrophin mediated by PDZ domains.
Cell
84:
757-767,
1996[ISI][Medline].
6.
Brenman, JE,
Chao DS,
Xia H,
Aldape K,
and
Bredt DS.
Nitric oxide synthase complexed with dystrophin and absent from skeletal muscle sarcolemma in Duchenne muscular dystrophy.
Cell
82:
743-752,
1995[ISI][Medline].
7.
Burbelo, PD,
Drechsel D,
and
Hall A.
A conserved binding motif defines numerous candidate target proteins for both Cdc42 and Rac GTPases.
J Biol Chem
270:
29071-29074,
1995
8.
DeLeo, FR,
and
Quinn MT.
Assembly of the phagocyte NADPH oxidase: molecular interaction of oxidase proteins.
J Leukoc Biol
60:
677-691,
1996[Abstract].
9.
Dunn, K,
and
Bacallao R.
Optical methods in renal research.
Semin Nephrol
18:
122-137,
1998[ISI][Medline].
10.
El Benna, J,
Ruedi JM,
and
Babior BM.
Cytosolic guanine nucleotide-binding protein Rac2 operates in vivo as a component of the neutrophil respiratory burst oxidase. Transfer of Rac2 and the cytosolic oxidase components p47phox and p67phox to the submembranous actin cytoskeleton during oxidase activation.
J Biol Chem
269:
6729-6734,
1994
11.
Faris, SL,
Rinckel LA,
Huang J,
Hong YR,
and
Kleinberg ME.
Phagocyte NADPH oxidase p67-phox possesses a novel carboxyl terminal binding site for the GTPases Rac2 and Cdc42.
Biochem Biophys Res Commun
247:
271-276,
1998[ISI][Medline].
12.
Fox, JE,
Reynolds CC,
and
Boyles JK.
Studying the platelet cytoskeleton in Triton X-100 lysates.
Methods Enzymol
215:
42-58,
1992[ISI][Medline].
13.
Gnanapandithen, K,
Chen Z,
Kau CL,
Gorczynski RM,
and
Marsden PA.
Cloning and characterization of murine endothelial constitutive nitric oxide synthase.
Biochim Biophys Acta
1308:
103-106,
1996[ISI][Medline].
14.
Gupta, AK,
Diaz RA,
Higham S,
and
Kone BC.
-MSH inhibits induction of C/EBP
-DNA binding activity and NOS2 gene transcription in macrophages.
Kidney Int
57:
2239-2248,
2000[ISI][Medline].
15.
Hall, A.
Rho-GTPases and the actin cytoskeleton.
Science
279:
509-514,
1998
16.
Hecker, M,
Walsh DT,
and
Vane JR.
Characterization of a microsomal calcium-dependent nitric oxide synthase in activated J774.2 monocyte/macrophages.
J Cardiovasc Pharmacol
20:
S139-S141,
1992[ISI][Medline].
17.
Hiki, K,
Yui Y,
Hattori R,
Eizawa H,
Kosuga K,
and
Kawai C.
Cytosolic and membrane-bound nitric oxide synthase.
Jpn J Pharmacol
56:
217-220,
1991[ISI][Medline].
18.
Jaffrey, SR,
and
Snyder SH.
PIN: an associated protein inhibitor of neuronal nitric oxide synthase.
Science
274:
774-777,
1996
19.
Ju, H,
Zou R,
Venema VJ,
and
Venema RC.
Direct interaction of endothelial nitric-oxide synthase and caveolin-1 inhibits synthase activity.
J Biol Chem
272:
18522-18525,
1997
20.
Kobayashi, K,
Kuroda S,
Fukata M,
Nakamura T,
Nagase T,
Nomura N,
Matsuura Y,
Yoshida-Kubomura N,
Iwamatsu A,
and
Kaibuchi K.
p140Sra-1 (specifically Rac1-associated protein) is a novel specific target for Rac1 small GTPase.
J Biol Chem
273:
291-295,
1998
21.
Koga, H,
Terasawa J,
Nunoi H,
Takeshige K,
Inagaki F,
and
Sumimoto H.
Tetratricopeptide repeat (TPR) motifs of p67phox participate in interaction with the small GTPase Rac and activation of the phagocyte NADPH oxidase.
J Biol Chem
274:
25051-25060,
1999
22.
Kone, BC.
Nitric oxide in renal health and disease.
Am J Kidney Dis
30:
311-333,
1997[ISI][Medline].
23.
Kone, BC,
and
Higham S.
Nitric oxide inhibits transcription of the Na+-K+-ATPase 1-subunit gene in an MTAL cell line.
Am J Physiol Renal Physiol
276:
F614-F621,
1999
24.
Kone, BC,
and
Higham SC.
A novel N-terminal splice variant of the rat H+-K+-ATPase 2 subunit. Cloning, functional expression, and renal adaptive response to chronic hypokalemia.
J Biol Chem
273:
2543-2552,
1998
25.
Kuroda, S,
Fukata M,
Kobayashi K,
Nakafuku M,
Nomura N,
Iwamatsu A,
and
Kaibuchi K.
Identification of IQGAP as a putative target for the small GTPases, Cdc42 and Rac1.
J Biol Chem
271:
23363-23367,
1996
26.
Lee, C,
Miura K,
Liu X,
and
Zweier JL.
Biphasic regulation of leukocyte superoxide generation by nitric oxide and peroxynitrite.
J Biol Chem
275:
38965-38972,
2000
27.
MacMicking, J,
Xie QW,
and
Nathan C.
Nitric oxide and macrophage function.
Annu Rev Immunol
15:
323-350,
1997[ISI][Medline].
28.
McNally, JG,
Karpova T,
Cooper J,
and
Conchello JA.
Three-dimensional imaging by deconvolution microscopy.
Methods
19:
373-385,
1999[ISI][Medline].
29.
Mohaupt, M,
Madsen KM,
Wilcox CS,
and
Kone BC.
Molecular characterization and localization of three nitric oxide synthase isoforms in rat kidney.
In: Biology of Nitric Oxide, , edited by Moncada S,
Feelisch M,
Busse R,
and Higgs EA.. New York: Portland, 1994, pt. 4, p. 137-141.
30.
Mohaupt, MG,
Schwobel J,
Elzie JL,
Kannan GS,
and
Kone BC.
Cytokines activate inducible nitric oxide synthase gene transcription in inner medullary collecting duct cells.
Am J Physiol Renal Fluid Electrolyte Physiol
268:
F770-F777,
1995
31.
Moilanen, E,
Moilanen T,
Knowles R,
Charles I,
Kadoya Y,
al-Saffar N,
Revell PA,
and
Moncada S.
Nitric oxide synthase is expressed in human macrophages during foreign body inflammation.
Am J Pathol
150:
881-887,
1997[Abstract].
32.
Pan, J,
Burgher KL,
Szczepanik AM,
and
Ringheim GE.
Tyrosine phosphorylation of inducible nitric oxide synthase: implications for potential post-translational regulation.
Biochem J
314:
889-894,
1996[ISI][Medline].
33.
Patterson, C,
Ruef J,
Madamanchi NR,
Barry-Lane P,
Hu Z,
Horaist C,
Ballinger CA,
Brasier AR,
Bode C,
and
Runge MS.
Stimulation of a vascular smooth muscle cell NAD(P)H oxidase by thrombin. Evidence that p47(phox) may participate in forming this oxidase in vitro and in vivo.
J Biol Chem
274:
19814-19822,
1999
34.
Ratovitski, EA,
Alam MR,
Quick RA,
McMillan A,
Bao C,
Kozlovsky C,
Hand TA,
Johnson RC,
Mains RE,
Eipper BA,
and
Lowenstein CJ.
Kalirin inhibition of inducible nitric-oxide synthase.
J Biol Chem
274:
993-999,
1999
35.
Ratovitski, EA,
Bao C,
Quick RA,
McMillan A,
Kozlovsky C,
and
Lowenstein CJ.
An inducible nitric-oxide synthase (NOS)-associated protein inhibits NOS dimerization and activity.
J Biol Chem
274:
30250-30257,
1999
36.
Reid, T,
Furuyashiki T,
Ishizaki T,
Watanabe G,
Watanabe N,
Fujisawa K,
Morii N,
Madaule P,
and
Narumiya S.
Rhotekin, a new putative target for Rho bearing homology to a serine/threonine kinase, PKN, and rhophilin in the rho-binding domain.
J Biol Chem
271:
13556-13560,
1996
37.
Rudolph, MG,
Bayer P,
Abo A,
Kuhlmann J,
Vetter IR,
and
Wittinghofer A.
The Cdc42/Rac interactive binding region motif of the Wiskott-Aldrich syndrome protein (WASP) is necessary but not sufficient for tight binding to Cdc42 and structure formation.
J Biol Chem
273:
18067-18076,
1998
38.
Schmidt, HH,
Warner TD,
Nakane M,
Forstermann U,
and
Murad F.
Regulation and subcellular location of nitrogen oxide synthases in RAW264.7 macrophages.
Mol Pharmacol
41:
615-624,
1992[Abstract].
39.
Spaargaren, M,
and
Bischoff JR.
Identification of the guanine nucleotide dissociation stimulator for Ral as a putative effector molecule of R-ras, H-ras, K-ras, and Rap.
Proc Natl Acad Sci USA
91:
12609-12613,
1994
40.
Stuehr, DJ.
Structure-function aspects in the nitric oxide synthases.
Annu Rev Pharmacol Toxicol
37:
339-359,
1997[ISI][Medline].
41.
Tapon, N,
Nagata K,
Lamarche N,
and
Hall A.
A new rac target POSH is an SH3-containing scaffold protein involved in the JNK and NF-B signalling pathways.
Embo J
17:
1395-1404,
1998
42.
Thomas, D,
Patterson SD,
and
Bradshaw RA.
Src homologous and collagen (Shc) protein binds to F-actin and translocates to the cytoskeleton upon nerve growth factor stimulation in PC12 cells.
J Biol Chem
270:
28924-28931,
1995
43.
Thompson, G,
Chalk PA,
and
Lowe PN.
Interaction of PAK with Rac: determination of a minimum binding domain on PAK (Abstract).
Biochem Soc Trans
25:
509S,
1997[Medline].
44.
Tolias, KF,
Couvillon AD,
Cantley LC,
and
Carpenter CL.
Characterization of a Rac1- and RhoGDI-associated lipid kinase signaling complex.
Mol Cell Biol
18:
762-770,
1998
45.
Tolias, KF,
Hartwig JH,
Ishihara H,
Shibasaki Y,
Cantley LC,
and
Carpenter CL.
Type I phosphatidylinositol-4-phosphate 5-kinase mediates Rac-dependent actin assembly.
Curr Biol
10:
153-156,
2000[ISI][Medline].
46.
Van Aelst, L,
Joneson T,
and
Bar-Sagi D.
Identification of a novel Rac1-interacting protein involved in membrane ruffling.
Embo J
15:
3778-3786,
1996[Abstract].
47.
Van Bergen en Henegouwen, PM,
den Hartigh JC,
Romeyn P,
Verkleij AJ,
and
Boonstra J.
The epidermal growth factor receptor is associated with actin filaments.
Exp Cell Res
199:
90-97,
1992[ISI][Medline].
48.
Vodovotz, Y,
Russell D,
Xie QW,
Bogdan C,
and
Nathan C.
Vesicle membrane association of nitric oxide synthase in primary mouse macrophages.
J Immunol
154:
2914-2925,
1995
49.
Wheeler, MA,
Smith SD,
Garcia-Cardena G,
Nathan CF,
Weiss RM,
and
Sessa WC.
Bacterial infection induces nitric oxide synthase in human neutrophils.
J Clin Invest
99:
110-116,
1997