NO is a short-lived free radical that plays a critical role in the regulation of cellular signaling. Mast cell (MC)-derived NO and exogenous NO regulate MC activities, including the inhibition of MC degranulation. At a molecular level, NO acts to modify protein structure and function through several mechanisms, including protein tyrosine nitration. To begin to elucidate the molecular mechanisms underlying the effects of NO in MCs, we investigated protein tyrosine nitration in human MC lines HMC-1 and LAD2 treated with the NO donor S-nitrosoglutathione. Using two-dimensional gel Western blot analysis with an anti-nitrotyrosine Ab, together with mass spectrometry, we identified aldolase A, an enzyme of the glycolytic pathway, as a target for tyrosine nitration in MCs. The nitration of aldolase A was associated with a reduction in the maximum velocity of aldolase in HMC-1 and LAD2. Nuclear magnetic resonance analysis showed that despite these changes in the activity of a critical enzyme in glycolysis, there was no significant change in total cellular ATP content, although the AMP/ATP ratio was altered. Elevated levels of lactate and pyruvate suggested that S-nitrosoglutathione treatment enhanced glycolysis. Reduced aldolase activity was associated with increased intracellular levels of its substrate, fructose 1,6-bisphosphate. Interestingly, fructose 1,6-bisphosphate inhibited IgE-mediated MC degranulation in LAD2 cells. Thus, for the first time we report evidence of protein tyrosine nitration in human MC lines and identify aldolase A as a prominent target. This posttranslational nitration of aldolase A may be an important pathway that regulates MC phenotype and function.

Mast cells (MCs) are critical effector cells of IgE-mediated allergic inflammation and are increasingly recognized for their roles in innate and adaptive immune responses (13). Reactive oxygen species and reactive nitrogen species are important in the regulation of MC secretion and in inflammatory conditions (4). Several studies demonstrated that MCs express NO synthases and can produce small amounts of NO (511). Moreover, NO from other sources, such as macrophages, can regulate MC activities (12). Endogenous and exogenous NO can inhibit MC degranulation; protease release; adherence to fibronectin; and leukotriene, cytokine, and chemokine production (4, 7). By contrast, NO can also upregulate MC properties, such as CD8 expression (13), cytotoxicity (14), and cyclooxygenase-2 expression (15). However, the mechanisms underlying the effects of NO on MCs are poorly understood.

At a molecular level, NO acts to modify protein structure and function through several mechanisms, including nitration of tyrosine and nitrosylation of cysteine residues and heme and thiol groups (16, 17). Tyrosine nitration is a marker of reactive nitrogen species and has been used as a biomarker in inflamed tissues. Protein tyrosine nitration can modify cellular signaling pathways, and nitration may be a pathogenic component in disease (18, 19). Protein conformational changes due to tyrosine nitration can generate novel antigenic epitopes, alter enzymatic activity, modulate metabolic pathways, and inhibit tyrosine phosphorylation by kinases (20). The targets for protein tyrosine nitration are diverse, and multiple functions can be affected by nitration of selected tyrosine residues on a protein (21, 22).

There are no reports of protein tyrosine nitration in MCs, although Ag/Ab-mediated activation of lung MCs from guinea pig induced peroxynitrite production, an important mediator of tyrosine nitration and, in turn, modulated the release of inflammatory mediators from MCs (23). Given the vast number of potential protein targets for this posttranslational modification by NO, we investigated protein tyrosine nitration in MCs using a proteomic approach. We identified that aldolase A, a critical enzyme in glycolysis, is one target for tyrosine nitration in human MC lines HMC-1 and LAD2. Nitration of aldolase inhibits its enzymatic activity and is associated with elevated levels of its substrate, fructose 1,6 bisphosphate (FBP), which can inhibit MC activation. Thus, the posttranslational tyrosine nitration of aldolase may be one of the mechanisms involved in NO-mediated regulation of MC function.

HMC-1, an immature human MC line derived from a patient with MC leukemia (a gift from J.H. Butterfield, Rochester, MN) was cultured in Iscove's medium with 5% heat-inactivated FBS, 2 mM l-glutamine, and 40 U/ml penicillin/streptomycin (all from Invitrogen, Burlington, Ontario, Canada). The cells were harvested at ~75% confluency (48 h) and were used if the passage number was <20. The seeding concentration between each passage was 1 × 105 cells/ml. LAD2, the growth factor-dependent human MC line (generously provided by Dr. A.S. Kirshenbaum and D.D. Metcalfe, National Institutes of Health, Bethesda, MD) (24), was cultured in StemPro-34 medium (Invitrogen) supplemented with 100 ng/ml recombinant human stem cell factor (PeproTech, Rocky Hill, NJ). LAD2 were cultured by hemidepletion of medium every 7 d and maintained at a cell concentration <5 × 105 cells/ml. LAD2 cells are used as an alternative to primary human MCs because they express IgERs and can be activated by IgE cross-linking (24).

Human cord blood-derived MCs (CBMCs) were cultured as described previously (25, 26). Briefly, placentae were obtained within 45 min of delivery in adherence with ethics approval from the University of Alberta and Capital Health Region and with informed patient consent at Royal Alexandra Hospital, Edmonton, Alberta, Canada. EDTA-treated umbilical cord blood was diluted with the same volume of 10 mM phosphate buffer (pH 7.4) containing 150 mM NaCl and layered over Histopaque-1077 (Sigma-Aldrich, Oakville, Ontario, Canada) at room temperature within 4 h of collection. The CBMC progenitor fraction was obtained by centrifugation at 1000 × g for 20 min at room temperature. The cells were washed twice with PBS and grown in tissue-culture flasks in AIM-V medium (Invitrogen) with 100 ng/ml recombinant human stem cell factor. Nonadherent cells were transferred to fresh culture flasks and grown for ≥8 wk. CD117 and IgER expression were confirmed by flow cytometry. All cells were maintained at 37°C in a humidified incubator at 5% CO2.

We chose an NO donor S-nitrosoglutathione (SNOG) (Calbiochem, San Diego, CA), whose decomposition rate is ~5% per hour in water at room temperature, and the half-life is ~80 h at 37°C (27). SNOG is a well-known NO donor that favors a cGMP-independent mechanism of action of NO in different systems (28), and 500 μM SNOG has been widely used in different studies on MCs and inhibits MC function (6, 2931). Moreover, SNOG has been used at up to 2 mM for 4 h to induce nitration in MCF-7 cells (32). Hence, our use of 500 μM SNOG over 4 h was designed to be consistent with physiological release of NO under some in vivo conditions. For the sham control, cells were treated with equivalent volumes of the vehicle (dH2O) in the culture media (v/v ratio of 1/40 of vehicle/media). The viability of the cells was documented before and after treatment with SNOG using trypan blue dye exclusion; the viability in all experiments was >95%.

MCs were centrifuged at 300 × g for 5 min at 4°C and washed three times with ice-cold PBS. Cell pellets were stored at −80°C until further use. The cell pellet was lysed by vortexing at high speed for 2 min in two-dimensional electrophoresis (2-DE) cell lysis buffer (9 M urea, 4% CHAPS, 50 mM DTT, 0.5% immobilized pH gradient [IPG] 3-10 ampholytes [Bio-Rad, Mississauga, Ontario, Canada], 10 μl Protease Arrest kit from Genotech [St. Louis, MO]) and incubated on ice for 60 min for protein denaturation and solubilization. Cell lysates were collected and stored at −80°C in 100-μl aliquots after centrifuging the homogenate at 17,530 × g for 20 min at 4°C. Protein concentration was determined using the Bradford reducing agent compatible/detergent compatible protocol. The 2D cleanup kit (Bio-Rad) was used according to the manufacturer’s instructions. The protein pellets were reconstituted in 2-DE rehydration buffer (8 M urea, 1.5% CHAPS, 0.2% DTT, 0.2% IPG 3–10 ampholytes), with a final concentration of 75 μg protein in 125 μl rehydration buffer, and loaded onto 7-cm isoelectric point (pI) 3–10 linear IPG strips (Bio-Rad). Isoelectric focusing (50 V 10 min, 250 V 30 min, 750 V 60 min, 8000 V 13000 Vh) was performed, and the strips were equilibrated for 10 min with equilibration buffer A (1% DTT, 6 M urea, 30% glycerol, 2% SDS, 50 mM Tris-HCl [pH 8.8], 0.0002% bromophenol blue), followed by 10 min with equilibration buffer B (2.5% iodoacetamide, 6 M urea, 30% glycerol, 2% SDS, 50 mM Tris-HCl [pH 8.8], 0.0002% bromophenol blue). Strips were loaded in Ready Gel Tris-HCl 10% gel (IPG) (Bio-Rad) and sealed with PROTEAN Plus overlay agarose (Bio-Rad). The proteins were separated at 130 V. Gels were stained with a PlusOne silver staining kit (GE Healthcare, Piscataway, NJ), and images were captured using Alphaimager 2200 (Alpha Innotech, San Leandro, CA).

Proteins were transferred to polyvinylidene difluoride (PVDF) membranes Immobilon-P with pore size 0.45 μm (Millipore, Bedford, MA), using the Transblot SD semidry transfer system (Bio-Rad) at 15 V for 90 min. Membranes were blocked for 60 min at room temperature with 5% nonfat milk (Bio-Rad) in Tris-buffered saline (pH 7.4) containing 0.05% Tween-20 for anti-nitrotyrosine Ab and with 5% BSA (Sigma-Aldrich) for anti-aldolase Ab. Rabbit polyclonal anti-nitrotyrosine Ab was purchased from Upstate (Millipore, Etobicoke, Ontario, Canada); goat polyclonal anti-rabbit muscle aldolase Ab was purchased from Chemicon (Billerica, MA), and secondary Abs were purchased from Jackson ImmunoResearch Laboratories (West Grove, PA). Rabbit IgG (Jackson ImmunoResearch Laboratories) and goat IgG (Chemicon) were used as isotype controls. Nitrotyrosine+ controls were purchased from Millipore. Membranes were incubated overnight with primary Abs (1:1,000 dilution) and with HRP-conjugated secondary Abs (1:10,000 dilution) and were visualized using ECL (Amersham Biosciences, Piscataway, NJ). Images were captured with high-performance chemiluminescence Hyperfilm (GE Healthcare).

The specificity of anti-nitrotyrosine Ab was demonstrated by treating the PVDF membranes after transferring the protein with 100 mM sodium dithionite (Sigma-Aldrich) in 50 mM sodium borate buffer (pH 9) for 2 h at room temperature. Dithionite treatment reduces the nitro group into an amino group, thereby preventing the Ab from binding with the nitrated epitope. After dithionite treatment, membranes were washed three times with distilled water (5 min each wash), followed by blocking with Odyssey blocking buffer (LI-COR Biosciences, Lincoln, NE). Secondary Abs were goat anti-rabbit 800 (1:10,000) and goat anti-mouse 680 (1:10,000; both from LI-COR Biosciences). Blots were visualized with an Odyssey imager (LI-COR Biosciences) by scanning simultaneously at 700 and 800 nm. Odyssey software was used for molecular mass determination and quantitation of Western blots.

An immunoaffinity column was constructed by coupling cyanogen bromide-activated Sepharose 4B with polyclonal goat anti-aldolase Ab using standard protocols (33). HMC-1 and LAD2 cell lysates were passed through this column and washed extensively with wash buffer. The bound proteins were eluted using low-pH elution (glycine 2.5 pH). The eluted fractions containing enriched MC aldoase were neutralized with 2 M Tris buffer (pH 11.2), and the fractions were concentrated using Centricon centrifugal filter 10K devices (Millipore). The concentrated fractions were separated in a 10% SDS-PAGE gel and transferred to PVDF membrane for Western blot analysis.

In-gel tryptic digestion, peptide extraction, and mass spectrometry (MS) analysis were performed using standard protocols at the Mass Spectrometry Facility, Department of Chemistry, University of Alberta. Briefly, proteins in the gel were reduced with 5 mM DTT and carbamidomethylated with 10 mM iodoacetamide, followed by tryptic digestion overnight with 0.06 μg/μl modified bovine trypsin (Promega, Madison, WI) at 30°C. Five microliters of the resultant peptide digests was loaded onto a nanoAcquity UPLC system with peptide trap (180 μm × 20 mm, Symmetry C18 nanoAcquity column,) and a nano analytical column (75 μm × 100 mm, Atlantis dC18 nanoAcquity column, both from Waters, Milford, MA). Desalting on the peptide trap was achieved by flushing it with 1% acetonitrile, 0.1% formic acid at a flow rate of 10 μl/min for 3 min. Peptides were separated with a gradient of 2–95% solvent B (acetonitrile, 0.1% formic acid) over 35 min at a flow rate of 300 nl/min. The column was connected to a q-ToF premier (Waters) for electrospray ionization-tandem mass spectrometry (MS/MS) analysis. Obtained MS/MS data were analyzed through proteomic software PEAKS (Bioinformatics Solutions, Waterloo, Ontario, Canada). Database searches were done with the following settings: carbamidomethylation as fixed modification, and oxidation as the variable modification. Peptide identifications were further confirmed by examination of the scores and manual inspection of the original MS/MS spectra. Good spectra with significant numbers of matched high-intensity peaks were considered important for the identification.

Total RNA was extracted with an RNeasy Plus mini kit (Qiagen, Mississauga, Ontario, Canada), according to the manufacturer’s instructions, quantified by measuring OD at 260 nm, and assessed by applying to 1.2% formaldehyde-agarose gels. cDNA was synthesized using the SuperScript First-Strand Synthesis System for RT-PCR (Invitrogen). Five micrograms of total RNA from each sample was used as a template for the reverse transcription reaction. The RNA/primer mixture (5 μg total RNA, 0.5 μg oligonucleotide12–18 primers, and 0.5 mM 2′-deoxynucleoside 5′-triphosphate mixture in RNase-free water) was incubated for 5 min at 65°C and then on ice for 1 min. The reaction mixture (40 U RNase inhibitor in 20 mM Tris-HCl [pH 8.4], 50 mM KCl, 5 mM MgCl2, 10 mM DTT) was added and incubated at 42°C for 2 min. Then, 50 U SuperScript II reverse transcriptase was added, and the reaction was continued for 50 min at 42°C, 15 min at 70°C, and chilled to 4°C on ice. Before proceeding to PCR, 2 U RNase H was added and incubated for 20 min at 37°C. Twenty microliters cDNA was diluted to 100 μl for use in PCR amplification. Ten microliters cDNA was used for PCR with recombinant Taq DNA Polymerase (Invitrogen). Human kidney cDNA was a kind gift from Dr. Philip F. Halloran, Division of Nephrology, University of Alberta. PCR was carried out with iCycler (Bio-Rad). Isoform-specific primers were designed based on the published human aldolase sequences: aldolase A sense 5′-AGCACCATGCCCTACCAATA-3′, antisense 5′-ACGACACCACACACCACTGT-3′; human aldolase B sense 5′-GATCGTGGTGGGAATCAAGT-3′, antisense 5′-CGCTTCATAAAAGCCTCCTG-3′; human aldolase C sense 5′-AGCGTACACCCTCTGCACTT-3′, antisense 5′-CACGCCCATAGGAGAAGGTA-3′; and human β-actin sense 5′-GGCATCCTCACCCTGAAGTA-3′, antisense 5′-AGGGCATACCCCTCGTAGAT-3′. The conditions for PCR amplification were denaturing at 95°C for 30 s, annealing at 55°C for 30 s, and extension at 72°C for 1 min, 25 cycles. The PCR products were analyzed by 1% agarose gel electrophoresis with ethidium bromide staining and confirmed by automated sequencing using a CEQ2000XL DNA sequenator (Beckman Coulter, Mississauga, Ontario, Canada) at the Biochemistry Department, University of Alberta.

HMC-1 and LAD2 (2 × 106 cells) were treated with 50 μl water (sham) or 500 μM SNOG in 2 ml media for 4 h, and the cell pellets in 1 ml Tris buffer (pH 7.4) containing protease inhibitors were rapidly frozen in liquid N2. The frozen cell lysate was subjected to ultrasound treatment for 5 min in an ultrasonic bath (FS 30H, Fisher Scientific, Ottawa, Ontario, Canada). Three cycles of freeze–thawing were performed, followed by centrifugation at 17,530 × g for 15 min at 4°C. The supernatant of cell lysate was stored on ice. The hydrazine assay (34, 35) was used to measure total aldolase activity in 100 μl cell lysate using FBP as substrate and hydrazine sulfate as detection reagent for the 3-phosphoglyceraldehyde produced by the reaction. One enzyme unit is described as a change in absorbance of 1.00/min at 25°C (pH 7.5) and 240 nm wavelength. To calculate the maximum velocity (Vmax) and the Michaelis–Menten constant (Km), the enzymatic activity was measured using a range of substrate (FBP) concentrations (3.125–800 μM). Km and Vmax values from aldolase activity, in terms of change in OD values (ΔA240/2 × 105 cells/min), were calculated using the Michaelis–Menten equation in a nonlinear fitting (solver function in Microsoft Excel, Microsoft, Redmond, WA).

HMC-1 and LAD2 cells (1 × 107) were sham treated or treated with 500 μM SNOG for 4 h. The cells were washed with PBS, and the cell pellet was deproteinized using 100 μl 5% trichloroacetic acid for 15 min on ice and then neutralized with 12 μl 2 M Tris (pH 11.2). The cell extracts were centrifuged at 17,530 × g for 20 min, and the supernatants containing the metabolite pools were analyzed for their FBP content. The standard protocol for measuring FBP (34, 36) was used with minor modifications. Briefly, the sample was incubated with aldolase (rabbit muscle aldolase, Sigma-Aldrich), the enzyme that cleaves FBP into dihydroxyacetone phosphate (DAP) and d-glyceraldehyde 3-phosphate. DAP and d-glyceraldehyde 3-phosphate are interconverted by the enzyme triose-phosphate isomerase. Glycerol-3-phosphate dehydrogenase catalyzes the reduction of DAP by NADH. The kinetic measurement of depletion of NADH at 340 nm was measured for 5 min in a 48-well plate (Corning, Lowell, MA) containing 170 μl 400 mM triethanolamine buffer (pH 7.6), 100 μl FBP standards, 10 μl 4.2 mM β-NADH, 10 μl glycerol-3-phosphate dehydrogenase, and triose-phosphate isomerase to give the initial reading. Then, 10 μl rabbit muscle aldolase was added to the reaction mixture, followed by measurement at 340 nm every minute for 5 min. Rabbit muscle aldolase and FBP (Sigma-Aldrich) were used to obtain a standard curve, and the OD values were calibrated with this curve to quantitate the FBP levels in the HMC-1 and LAD2 cell lysate.

LAD2 cells (1 × 106) were seeded in a six-well plate and sensitized with 1 μg/ml human IgE (Chemicon) overnight at 37°C. Cells were pretreated for 4 h with 500 μM SNOG, 5 mM FBP, 5 mM fructose-1-phosphate (F1P), or 5 mM fructose-6-phosphate (F6P). Cells were washed and resuspended in HEPES-buffered Tyrode’s solution (pH 7.2), followed by stimulation with 10 μg/ml mouse anti-human IgE Ab (Dako, Mississauga, Ontario, Canada) for 30 min at 37°C. The cells were centrifuged, and the percentage of β-hexosaminidase (β-hex) released into the supernatant was calculated, as described previously (37).

HMC-1 cells (1 × 108) were treated with sham or 500 μM SNOG for 4 h. Cells were centrifuged at 300 × g for 5 min at 4°C. One milliliter of supernatant was collected from each treatment and stored at −80°C until nuclear magnetic resonance (NMR) analysis. Cell pellets were washed with ice-cold PBS, and the metabolites from the cell pellets were extracted using a modified methanol/chloroform extraction procedure, as described by Bligh and Dyer (38). Briefly, cells were treated with 500 μl methanol/chloroform mixture (2:1) followed by rapid freezing in liquid N2. Three cycles of freeze–thawing were performed, followed by the addition of 100 μl cold methanol and 100 μl water. The cell lysates were centrifuged at 100 × g for 15 min, and 400 μl the top layer containing metabolite fractions was aliquoted, frozen, lyophilized, and stored at −80°C until NMR analysis. Cell-culture supernatants were filtered through 3 kDa cutoff filters. Cell lysate samples were prepared for NMR analysis by the addition of 185 μl dH2O and 65 μl internal standard containing 5 mM 2,2-dimethyl-2-silapentane-5-sulfonate and 0.2% NaN3 in 100% D2O. Filtered cell supernatants were prepared by the addition of 65 μl internal standard to 585 μl supernatant. Sample pH was adjusted to 6.8. One-dimensional NMR spectra of the extracted metabolite fractions were acquired using the first increment of the standard NOESY pulse sequence on a four-channel Varian Inova 800-MHz (cell lysate samples) or 600-MHz (cell supernatants) NMR spectrometer (Varian, Palo Alto, CA). All spectra were recorded at 25°C with a 12-parts per million sweep width, 1-s recycle delay, 100-ms mixing time, an acquisition time of 4 s, four dummy scans, and 32 transients, as previously described (39). [1H] decoupling of the water resonance was applied for 0.9 s of the recycle delay and during the 100-ms mixing time. All spectra were zero-filled to 128,000 data points and multiplied by an exponential weighting function corresponding to a line-broadening of 0.5 Hz (39).

HMC-1 and LAD2 cells (1 × 106) were treated with sham or 500 μM SNOG for 4 h in 6-well culture plates (Becton Dickinson, Mississauga, Ontario, Canada). Fifty microliters the media containing 5 × 103 cells were loaded in 96-well flat-bottom black plates (Corning) in triplicate. ATPlite 1step Luminescence ATP detection assay kit was used to quantitate ATP in the viable cells by luminescence measured using an I450 Micro β Scintillation and Luminescence detector (both from PerkinElmer, Woodbridge, Ontario, Canada).

Data were analyzed with the unpaired Student t test and the Tukey–Kramer multiple comparisons test (one-way ANOVA). A p value <0.05 was considered significant.

To investigate whether there was constitutive nitration among MC proteins, we performed Western blot with normally cultured, unstimulated HMC-1 and LAD2 cell lysates using anti-nitrotyrosine Ab. The anti-nitrotyrosine Ab detected many nitrated proteins (Fig. 1A), whereas dithionite treatment diminished the immunoreactivity, thereby confirming the existence of constitutive protein tyrosine nitration in MCs.

FIGURE 1.

Constitutive and SNOG-induced nitrotyrosine+ proteins of HMC-1 and LAD2 cells. A, Western blot using anti-nitrotyrosine Ab of HMC-1 proteins (lanes1, 4), LAD2 proteins (lanes2, 5) (100 μg/lane), and nitrotyrosine positive control (lanes3, 6). Dithionite treatment (lanes4–6) diminishes the nitrotyrosine immunoreactivity, establishing that there is constitutive nitration of MC proteins. Results are representative of three independent experiments. B, HMC-1 cells were treated (right panel) or not (left panel) with 500 μM SNOG for 4 h, and proteins were extracted for 2-DE. Western blots of 2-DE gels using anti-nitrotyrosine Ab show constitutively nitrated aldolase at pI 9.0, ~40 kDa (arrow, left panel) and prominent SNOG-induced nitration at pI 8.3, ~40 kDa (arrowhead, right panel). Results are representative of seven independent experiments. C, Western blots of 2-DE gels using anti-aldolase Ab. Multiple immunoreactive aldolase spots near molecular mass ~40 kDa with different pI (7.0–9.3) (arrow, left panel) and SNOG-induced nitration of aldolase at pI 8.3 (arrowhead, right panel) were detected in HMC-1 cells. Aldolase at pI 9.0 was found to be nitrated constitutively (compare arrows in left panels, A, B). Results are representative of seven independent experiments. D, LAD2 cells were treated (right panel) or not (left panel) with 500 μM SNOG for 4 h, and proteins were extracted for 2-DE. Western blots of 2-DE gels using NT and Aldo. Multiple immunoreactive aldolase spots near molecular mass ~40 kDa with different pI (7.0–9.3) and SNOG-induced nitration of aldolase at pI 8.3 were detected in LAD2 cells. Results are representative of four independent experiments. E, Western blots of affinity-purified aldolase of HMC-1 and LAD2 cells before and after dithionite treatment using NT and Aldo. The magnitude of dithionite reduction of nitrotyrosine reactivity of HMC-1 cells is represented as a bar graph (mean ± SE; densitometry from three independent experiments). *p < 0.05 compare with sham treatment. Aldo, anti-aldolase Ab; NT, anti-nitrotyrosine Ab.

FIGURE 1.

Constitutive and SNOG-induced nitrotyrosine+ proteins of HMC-1 and LAD2 cells. A, Western blot using anti-nitrotyrosine Ab of HMC-1 proteins (lanes1, 4), LAD2 proteins (lanes2, 5) (100 μg/lane), and nitrotyrosine positive control (lanes3, 6). Dithionite treatment (lanes4–6) diminishes the nitrotyrosine immunoreactivity, establishing that there is constitutive nitration of MC proteins. Results are representative of three independent experiments. B, HMC-1 cells were treated (right panel) or not (left panel) with 500 μM SNOG for 4 h, and proteins were extracted for 2-DE. Western blots of 2-DE gels using anti-nitrotyrosine Ab show constitutively nitrated aldolase at pI 9.0, ~40 kDa (arrow, left panel) and prominent SNOG-induced nitration at pI 8.3, ~40 kDa (arrowhead, right panel). Results are representative of seven independent experiments. C, Western blots of 2-DE gels using anti-aldolase Ab. Multiple immunoreactive aldolase spots near molecular mass ~40 kDa with different pI (7.0–9.3) (arrow, left panel) and SNOG-induced nitration of aldolase at pI 8.3 (arrowhead, right panel) were detected in HMC-1 cells. Aldolase at pI 9.0 was found to be nitrated constitutively (compare arrows in left panels, A, B). Results are representative of seven independent experiments. D, LAD2 cells were treated (right panel) or not (left panel) with 500 μM SNOG for 4 h, and proteins were extracted for 2-DE. Western blots of 2-DE gels using NT and Aldo. Multiple immunoreactive aldolase spots near molecular mass ~40 kDa with different pI (7.0–9.3) and SNOG-induced nitration of aldolase at pI 8.3 were detected in LAD2 cells. Results are representative of four independent experiments. E, Western blots of affinity-purified aldolase of HMC-1 and LAD2 cells before and after dithionite treatment using NT and Aldo. The magnitude of dithionite reduction of nitrotyrosine reactivity of HMC-1 cells is represented as a bar graph (mean ± SE; densitometry from three independent experiments). *p < 0.05 compare with sham treatment. Aldo, anti-aldolase Ab; NT, anti-nitrotyrosine Ab.

Close modal

In addition to constitutive protein tyrosine nitration in MC, we tested whether specific proteins would undergo nitration upon cell treatment with an NO donor. We treated HMC-1 cells with different doses and at different time periods (50, 250, and 500 μM for 0.5, 2, and 4 h for each dose) with SNOG, a slow-releasing NO donor. 2-DE Western blots with anti-nitrotyrosine Ab were used to detect changes in the pattern of nitrated proteins in SNOG-treated HMC-1 cells. Treatment with 500 μM of SNOG for 4 h induced a significant change in the nitration pattern (compare panels in Fig. 1B). Isotope controls using rabbit IgG did not identify any immunoreactivity (data not shown). SNOG-induced nitration was reproducibly evident on a protein at ~40 kDa and pI 8.3 (arrowhead in right panel, Fig. 1B). In additional studies, 500 μM SNOG was used for 4 h. The SNOG-induced nitrated spot at pI 8.3, ~40 kDa was matched with the corresponding spot in a silver-stained gel run in parallel, and the spot was removed under sterile conditions and submitted for spectroscopy (MS) identification of the protein.

MS analysis of the SNOG-induced nitrated spot (~40 kDa, pI 8.3) revealed that peptides matched human aldolase A. To ensure that the spot submitted to MS analysis and the protein selectively nitrated upon SNOG treatment were the same, we reprobed the membranes with anti-aldolase Ab and confirmed that the spot at pI 8.3, ~40 kDa was aldolase (arrowhead in right panel, Fig. 1C). We also documented multiple immunoreactive spots for aldolase near molecular mass ~40 kDa but at different pI, ranging from 7.0 to 9.0, in sham- and SNOG-treated HMC-1 cells (Fig. 1C). Aldolase at pI 9.0 matched the nitrated spot at pI 9.0 (compare and see arrows in left panel, Fig. 1B, 1C), indicating constitutive nitration of aldolase. SNOG-induced nitration was observed for aldolase A at pI 8.3 (arrowhead in right panel, Fig. 1B, 1C). Similar results were obtained when LAD2 cells were treated with SNOG. The region around 40 kDa in the 2-DE gel was highlighted to demonstrate that SNOG induces selective nitration of aldolase at pI 8.3 in LAD2 cells, as well as in HMC-1 cells (Fig. 1D). Moreover, in addition to aldolase, other spots were nitrated following SNOG treatment in LAD2 cells (data not shown). To further confirm constitutive nitration of aldolase, we enriched aldolase from HMC-1 or LAD2 cells using an aldolase affinity column and performed Western blot with anti-nitrotyrosine Ab (Fig. 1E). Dithionite treatment significantly reduced the nitro tyrosine reactivity of aldolase fractions, thereby confirming constitutive nitration of aldolase in these two MC lines. Similar results were observed when we preincubated nitrotyrosine Ab with free 3-nitrotyrosine (data not shown).

Because there were multiple pI forms of aldolase in MCs, we defined the different isoforms of aldolase in MCs. To investigate the mRNA expression of the isoforms of aldolase in MCs, we used RT-PCR with isoform-specific primers for aldolase A, B, and C with cDNA of HMC-1 cells, LAD2 cells, and CBMCs. cDNA from nontumor areas of kidney from a patient who underwent nephrectomy was used as a positive control for all three isoforms of aldolase. We identified mRNA for two isoforms, aldolase A and C, in HMC-1 cells, LAD2 cells, and CBMCs, whereas mRNA for all three isoforms was found in human kidney (Fig. 2). The products were sequenced and confirmed. Similarly, the affinity-purified fractions of aldolase from HMC-1 cells were analyzed using MS. We found peptides for aldolase A (Supplemental Table I) and aldolase C (Supplemental Table II).

FIGURE 2.

Expression of aldolase A and C isoforms in different MC lines. RT-PCR using cDNA of human kidney, HMC-1 and LAD2 cells and CBMCs using aldolase isoform (aldolase A, aldolase B, aldolase C)-specific and β-actin primers. All PCR products were confirmed by sequencing, and the results are representative of five independent experiments.

FIGURE 2.

Expression of aldolase A and C isoforms in different MC lines. RT-PCR using cDNA of human kidney, HMC-1 and LAD2 cells and CBMCs using aldolase isoform (aldolase A, aldolase B, aldolase C)-specific and β-actin primers. All PCR products were confirmed by sequencing, and the results are representative of five independent experiments.

Close modal

To determine whether nitration of aldolase affects the activity of the enzyme, we measured the Km and Vmax of aldolase from HMC-1 and LAD2 cell homogenates before and after treatment with SNOG. SNOG induced a reduction in the Km and the Vmax of aldolase in HMC-1 cells (Fig. 3A, 3B) and a reduction in the Vmax of aldolase in LAD2 cells (Fig. 3D). The raw data representing the aldolase activity in terms of change in OD values (ΔA240/2 × 105 cells/min) between sham- and SNOG-treated HMC-1 and LAD2 cells are shown in Supplemental Fig. 1. Because inhibition of the activity of an enzyme can result in accumulation of the substrate, we measured levels of FBP, the substrate for aldolase A, before and after treatment with SNOG. SNOG treatment significantly increased the intracellular FBP levels by 613% in HMC-1 cells and by 61% in LAD2 cells (Fig. 4).

FIGURE 3.

SNOG-induced reduction in the total aldolase enzymatic activity of MCs. HMC-1 and LAD2 cells (2 × 106) were treated with 50 μl of water or 500 μM SNOG for 4 h, and the total aldolase activity in the cell lysates was measured. To calculate the Vmax and the Km, 3.125–800 μM (serial 2-fold dilution) of substrate was used. SNOG treatment reduced the Km and Vmax of aldolase in HMC-1 cells (A, B), but it only reduced the Vmax in LAD2 cells (C, D). Results are from seven (HMC-1) and five (LAD2) independent experiments (mean ± SE). *p < 0.05.

FIGURE 3.

SNOG-induced reduction in the total aldolase enzymatic activity of MCs. HMC-1 and LAD2 cells (2 × 106) were treated with 50 μl of water or 500 μM SNOG for 4 h, and the total aldolase activity in the cell lysates was measured. To calculate the Vmax and the Km, 3.125–800 μM (serial 2-fold dilution) of substrate was used. SNOG treatment reduced the Km and Vmax of aldolase in HMC-1 cells (A, B), but it only reduced the Vmax in LAD2 cells (C, D). Results are from seven (HMC-1) and five (LAD2) independent experiments (mean ± SE). *p < 0.05.

Close modal
FIGURE 4.

SNOG-induced elevation of FBP in MCs. HMC-1 cells (A) or LAD2 cells (B) (1 × 107) were treated with 250 μl of water or 500 μM SNOG in 10 ml media for 4 h, and the total cell pellet were deproteinized using 5% TCA followed by neutralization by 2 M Tris. The FBP concentration was estimated in the metabolite fractions. Results are from five (A) or eight (B) independent experiments (mean ± SE). *p < 0.05.

FIGURE 4.

SNOG-induced elevation of FBP in MCs. HMC-1 cells (A) or LAD2 cells (B) (1 × 107) were treated with 250 μl of water or 500 μM SNOG in 10 ml media for 4 h, and the total cell pellet were deproteinized using 5% TCA followed by neutralization by 2 M Tris. The FBP concentration was estimated in the metabolite fractions. Results are from five (A) or eight (B) independent experiments (mean ± SE). *p < 0.05.

Close modal

Because aldolase is a crucial enzyme in the glycolytic pathway, we studied whether inhibition of aldolase A activity might alter ATP levels in MCs. We used [1H] NMR to begin to define components of the metabolome (set of <1500-Da metabolites) (39, 40) of HMC-1 cells treated with sham or SNOG. Surprisingly, in contrast to the expectation from the inhibition of aldolase activity, ATP levels in cell extracts were similar in SNOG- and sham-treated MCs (Fig. 5A), although the AMP/ATP ratio was increased following SNOG treatment (Fig. 5B). Interestingly, pyruvate and lactate, measures of cellular glycolytic activity, were significantly elevated in culture media of SNOG-treated MCs (Fig. 5C, 5D). Similar results with HMC-1 and LAD2 cells were obtained using a different assay system (ATPlite 1step ATP kit; Supplemental Fig. 2).

FIGURE 5.

SNOG-induced glycolytic intermediates change in HMC-1 cells. HMC-1 cells (1 × 108) were treated with sham (2.5 ml water in 100 ml media) or with 500 μM SNOG for 4 h, and glycolytic metabolites were analyzed using NMR. SNOG did not increase the ATP level (A), but it significantly increased the AMP/ATP ratio (B) and the pyruvate (C) and lactate (D) levels (enhanced glycolysis). Error bars represent SE from three independent experiments. *p < 0.05.

FIGURE 5.

SNOG-induced glycolytic intermediates change in HMC-1 cells. HMC-1 cells (1 × 108) were treated with sham (2.5 ml water in 100 ml media) or with 500 μM SNOG for 4 h, and glycolytic metabolites were analyzed using NMR. SNOG did not increase the ATP level (A), but it significantly increased the AMP/ATP ratio (B) and the pyruvate (C) and lactate (D) levels (enhanced glycolysis). Error bars represent SE from three independent experiments. *p < 0.05.

Close modal

To simulate the effects of increased intracellular FBP in MCs and to study the effects of FBP on MC function, we used an in vitro IgE-dependent degranulation assay of MCs. Because exogenously applied FBP was shown to enter intracellular compartments poorly (41), we used 5 mM exogenous FBP in our model, as other investigators used previously (42). Indeed, when we tested intracellular levels of FBP following treatment with exogenous FBP (100 nmol to 100 μmol), only a few nanomoles of FBP were available in the intracellular compartment of LAD2 cells (Supplemental Fig 3).

IgE/anti-IgE stimulation induced release of 21.8 ± 2.1% of the total cellular β-hex from LAD2 cells, whereas pretreatment with SNOG or FBP significantly inhibited this degranulation of LAD2 by 59 or 48%, respectively (Fig. 6). However, F1P and F6P, as negative controls for FBP, did not inhibit MC degranulation.

FIGURE 6.

Effects of exogenous FBP on degranulation of LAD2 cells. LAD2 cells were sensitized with 1 μg/ml human IgE overnight at 37°C and then pretreated with 500 μM SNOG, 5 mM FBP, 5 mM F1P, or 5 mM F6P for 4 h at 37°C. The cells were washed and resuspended in HEPES-buffered Tyrode’s solution and stimulated with 10 μg/ml mouse anti-human IgE Ab for 30 min. The cells were centrifuged, and the percent release of β-hex into the supernatant was calculated. Results are expressed as mean ± SE from six independent experiments performed in triplicate. Data were analyzed using one-way ANOVA, followed by the Tukey–Kramer multiple comparisons test. *p < 0.05.

FIGURE 6.

Effects of exogenous FBP on degranulation of LAD2 cells. LAD2 cells were sensitized with 1 μg/ml human IgE overnight at 37°C and then pretreated with 500 μM SNOG, 5 mM FBP, 5 mM F1P, or 5 mM F6P for 4 h at 37°C. The cells were washed and resuspended in HEPES-buffered Tyrode’s solution and stimulated with 10 μg/ml mouse anti-human IgE Ab for 30 min. The cells were centrifuged, and the percent release of β-hex into the supernatant was calculated. Results are expressed as mean ± SE from six independent experiments performed in triplicate. Data were analyzed using one-way ANOVA, followed by the Tukey–Kramer multiple comparisons test. *p < 0.05.

Close modal

Endogenous and exogenous NO have been widely reported to modulate several MC functions (43). One common pathway involved is that NO activates soluble guanylate cyclase and increases intracellular cGMP, which, in turn, regulates numerous physiological events in the cell (44). However, many studies indicated that non-cGMP–mediated pathways are also important in the effects of NO on MC function (6, 12, 31, 45). Earlier, we suggested that S-nitrosylation of calpain was responsible for the NO-mediated inhibition of MC adhesion (31). In the current study, we provide the first evidence that protein tyrosine nitration plays a role in NO-mediated inhibitory effects on MC function.

Using Western blot with an anti-nitrotyrosine Ab, we identified constitutive protein tyrosine nitration in MCs, thereby providing evidence that this posttranslational modification of proteins occurs in a variety of MCs (Fig. 1). Our observation is supported by evidence that constitutive nitration occurs in other leukocytes, such as macrophage cell lines (46), eosinophils (47, 48), and neutrophils (49). It is interesting that a recent study claims that there are no nitrotyrosine+ proteins in MCs on lung sections of cystic fibrosis patients (48). Unfortunately, isotype controls were not reported in that study, and there was no information about optimization of sensitivity of the immunohistochemistry to detect nitrotyrosine reactivity in MCs or the numbers of MCs assessed on the tissues.

In addition to the constitutive nitration of multiple MC proteins, we identified SNOG-induced changes in protein nitration (Fig. 1B). It is interesting that aldolase at pI 9.0 was nitrated constitutively, whereas NO-mediated nitration of aldolase A was associated with an 8.3 pI form (Fig. 1B, 1C). The presence of multiple pI forms (~7.0–9.3) of aldolase in MCs is consistent with earlier findings in rat diaphragm and mouse sperm cells (50, 51). Aldolase can also be phosphorylated (52) and regulated by different posttranslational modifications.

Interestingly, aldolase A can be found in the cytoplasm (53), mitochondria (54), and the heterochromatin region of nucleus (55), and in vivo nitration can be seen in many intracellular compartments (56), including nuclear histones (57). Tyrosine nitration of a protein can be mediated by multiple mechanisms, including ONOO (22). In MCs, ONOO can be generated by the reaction of NO and MC-produced O2, which could, in turn, nitrate aldolase. The half-life of ONOO is only a few seconds; thus, ONOO reactivity may act in the local microenvironment on a limited spectrum of proteins. In our experiment, we decided to use a slow-releasing NO donor SNOG (decomposition rate is ~5% per hour in water at room temperature, and the half-life is ~80 h at 37°C), which would likely combine with physiologically relevant, MC-derived O2 and, thus, generate a more physiological nitration event. In addition to ONOO, there are other mechanisms of nitration (58); however, the precise mechanisms underlying the nitration of aldolase are beyond the scope of the current study. Another consideration is that NO might facilitate selective nitration of aldolase A in different intracellular compartments of HMC-1 cells and modulate compartment-specific functions of aldolase (see below).

There are three isoforms of mammalian aldolase: A, B, and C (59). Aldolase A and C are preferentially involved in the glycolytic cycle and predominantly expressed in muscle and brain, respectively, whereas aldolase B is reported to be preferentially involved in gluconeogenesis and expressed in liver (59). However, the expression of these isoforms is not restricted to specific tissues, (e.g., all three isoforms are expressed in kidney) (60, 61). Tyrosine nitration of aldolase A, B, and C was reported in a wide range of tissues and under different inflammatory conditions (50, 6265). To define specific tyrosine residues that are nitrated in aldolase A or in aldolase C, MS identification of nitration-specific sites requires large amounts of protein; we are working toward enriching the nitrated aldolase fractions with specific techniques (66).

Aldolase A is the predominant isoform expressed in lymphocytes (67). Aldolase C expression was reported in platelets, but it was not detectable in erythrocytes and lymphocytes (68). It is interesting that MCs express aldolase C, the isoform predominantly expressed in neuronal tissues. The potential significance of aldolase C expression in MCs is multiple. For example, if there is inactivation of one isoform of the enzyme, the other isoform, in this case aldolase C, might act as a backup enzyme in the glycolytic cycle. Interestingly, aldolase C was reported to be involved in a stress-response pathway for lung epithelial cell function during hypoxia (69). Moreover, subsets of cerebellar Purkinje cells expressing aldolase C were resistant to excitotoxic insult (70), and aldolase C small interfering RNA transfection resulted in the death of these subsets of Purkinje cells. Because MCs play a major role in TGF-β1–mediated excitotoxicity (71), expression of aldolase C may be of functional significance in MC homeostasis.

It is intriguing that aldolase A has been identified as a target of tyrosine nitration in several other cells and tissues, including diaphragm (50), human skin fibroblasts (35), quadriceps muscle of patients with severe chronic obstructive pulmonary disease (72), and in lung tissue in a model of asthma (65). However, this is the first report in human MCs. The discovery of aldolase as one of the targets for nitration in MCs might be due to the abundance of this protein, the extent of nitration of tyrosines in aldolase, and the sensitivity of the methods used. Interestingly, other proteins were nitrated (e.g., pI 7.0, molecular mass 25 kDa) upon SNOG treatment in LAD2 cells (data not shown). We are currently working toward identifying these targets using MS.

An earlier in vitro study using rabbit muscle aldolase treated with ONOO demonstrated four tyrosine residues (Y222, Y243, Y342, and Y363) that were targets for tyrosine nitration (35). Although, we do not know which tyrosine residues in MC aldolase A are nitrated following SNOG treatment, NO treatment reduced total activity of MC aldolase (Fig. 3), as previously shown using rabbit muscle aldolase (35). Because it is difficult to separate the two isoforms of aldolase A and C from the whole-cell lysate, we studied total aldolase activity, as reported earlier in different tissues and under different conditions (55, 73). The tyrosine at the carboxy terminal end of the aldolase A (i.e., tyrosine 363) was critical for the enzymatic activity; nitration of this residue significantly reduced the Vmax of the enzyme activity (35). Interestingly, our observation that NO induced a reduction in the Km and Vmax of HMC-1 aldolase is consistent with earlier studies in which substitution of tyrosine 363 with serine in human aldolase A also reduced enzyme Km and Vmax (74). SNOG reduced the Km of HMC-1 aldolase, whereas there was no change in the Km of LAD2 cells (Fig. 3). Moreover, there was a 10-fold difference between the Km values of HMC-1 and LAD2 cells (Fig. 3). Thus, there may be cell-type–specific differences in the effects of nitration (i.e., LAD2 is a mature MC line, whereas HMC-1 is an immature cell line). Moreover, the expression and intracellular distribution of aldolase isoforms A and C may be different between these two cell lines, and the tyrosine nitration sites may differ as well.

Given the reduction in the enzymatic activity of aldolase, we postulated that this would increase the levels of its substrate FBP. Indeed, we found that the intracellular FBP in HMC-1 cell extracts significantly increased following SNOG treatment (Fig. 4A). Similarly, a statistically significant increase in FBP levels following SNOG treatment was documented in LAD2 cells (Fig. 4B). It is interesting to note that the SNOG-induced FBP levels also differed between HMC-1 and LAD2 cells, as did the Km values of these two MC lines (see above). Thus, the difference in magnitude of Km values, unaltered Km values for LAD2 cells compared with the reduction in Km of HMC-1 cells after SNOG treatment, and the variable intracellular FBP levels strongly suggest that the outcome of nitration might be different between HMC-1 and LAD2 cells. Further detailed study is necessary to understand our observations. Interestingly, a mutation of aldolase A (Glu206Lys) was reported in a patient with an inherited metabolic myopathy and hemolysis and was associated with increased FBP levels (182.9% of control) in the patient’s RBCs (75). Moreover, NO-induced inhibition of GAPDH, the enzyme downstream of aldolase in the glycolytic cycle, and a corresponding increase in FBP levels in rat intestinal tissues were reported (76). Thus, increased FBP levels upon inactivation of aldolase A may have physiological relevance.

To assess the effects of aldolase nitration on the metabolic pathways of MCs, we analyzed metabolites using NMR. ATP levels in cell extracts were similar in sham- and SNOG-treated HMC-1 cells (Fig. 5A). It is interesting that cellular ATP levels remained similar, despite a reduction in the activity of aldolase, thereby suggesting restoration of ATP levels through homeostatic mechanisms. AMP-activated protein kinase was reported to act as a fuel gauge in mammalian cells. It can enhance cellular ATP levels by switching off energy-using pathways and switching on energy-generating pathways (77, 78). AMP-activated protein kinase activation occurs when there is an increase in AMP levels and a decrease in ATP levels (increased AMP/ATP ratio) (77). We predicted that this regulatory pathway also exists in MCs; therefore, we measured the AMP/ATP ratio before and after SNOG treatment. Indeed, the AMP/ATP ratio increased following SNOG treatment in HMC-1 cells (Fig. 5B), providing evidence for homeostatic regulation of ATP levels in our model. In addition to in HMC-1 cells, the ATP levels in LAD2 cells were unaltered after SNOG treatment (Supplemental Fig. 2). The lower ATP levels detected using NMR (Fig. 5A) compared with live cell-based ATP assay (Supplemental Fig. 2) might be due to the loss of ATP during the preparation of samples for NMR (see 1Materials and Methods).

Measurement of pyruvate, the end product of glycolysis, and lactate, the end product of anaerobic glycolysis, in the media of cultured cells provides a clear understanding of the metabolic status of the cells (79). Interestingly, pyruvate and lactate were significantly elevated in culture media of NO-treated MCs (Fig. 5C, 5D), indicating enhanced glycolysis, as previously shown in astrocytes treated with NO (80). NO downregulates mitochondrial energy production, but the cells maintain energy production by regulating glycolysis in astrocytes (81). It seems paradoxical that, although aldolase enzymatic activity is depressed (Fig. 3), glycolysis seems to be enhanced. In support of our observations, studies of skeletal muscle showed that elevated FBP is associated with increased glycolysis (81) and FBP activation of 6-phosphofructo-1-kinase, a key rate-limiting enzyme in glycolysis (81). It is interesting that some cells can respond to NO by activating glycolysis (astrocytes), whereas others do not (neurons) (81). Our results suggest that SNOG fundamentally alters MC metabolism to keep energy production at a homeostatic level.

FBP has several activities, including protection in ischemia/reperfusion injury and hypoxia (76, 82) and suppression of T cell proliferation (83). FBP has been used as a component of protective solutions in the transport of donor organs, such as liver (84). Other investigators have filed a patent on FBP as a therapeutic for asthma (85), and a recent in vivo study in rats demonstrated that FBP has an inhibitory role on MC degranulation and histamine release (86). We hypothesized that increased FBP levels, following aldolase A nitration in MCs, might inhibit some MC functions. Indeed, FBP significantly inhibited the β-hex release by LAD2 cells (Fig. 6), whereas similar concentrations of F1P or F6P did not have significant effects, supporting the specific effects of FBP on MC degranulation. The effects of NO donors on human MC activation depends on their NO release kinetics rather than the amount of NO released (87). Hence, our data on the effects of SNOG pretreatment followed by MC activation in an NO-free environment supports NO-mediated posttranslational modification, in this case nitration of aldolase. Moreover, the slow NO-releasing donor SNOG has been widely reported to mediate its action through non-cGMP pathways (28, 87).

Phospholipase C (PLC) cleaves phosphatidylinositol 4,5-bisphosphate to diacylglycerol and inositol 1,4,5-triphosphate (IP3), a critical intracellular messenger. The PLCγ signaling pathway plays a pivotal role in MC degranulation (88). Interestingly, some effects of FBP in other cells depend on PLC signaling (82). Moreover, FBP shares a binding site on aldolase A with IP3 (89) and significantly inhibits IP3 binding to aldolase A (89, 90). This binding also inhibits aldolase A enzymatic activity and influences release/repartition of aldolase A from binding to cytoskeletal elements (90). The posttranslational modification of tyrosine nitration of aldolase A may alter its relative affinity for FBP and/or IP3. Thus, it is possible that NO acts through excess FBP that modifies IP3 and PLC signaling cascades critical in MC secretion in IgE-dependent responses. Interestingly, aldolase can bind to many intracellular proteins, such as actin and ryanodine receptor (91, 92). Aldolase nitration might alter some of these interactions and, in turn, regulate MC function. Further studies with mature MCs may enhance our knowledge of these pathways.

In conclusion, we have provided evidence that protein tyrosine nitration regulates MC function. We identified that aldolase A, a critical enzyme in the glycolytic pathway, is a target for nitration. Because of the importance of aldolase A in cellular metabolism and homeostasis, further dissection of the relevance of our observations is necessary to more fully elucidate the role of tyrosine nitration in the regulation of MC function in allergic and other diseases.

We thank Randy Whittal and Jing Zheng for help with MS analysis. We also thank Chris St. Laurent and Katherine Morris for critical comments and help with the preparation of the manuscript.

Disclosures The authors have no financial conflicts of interest.

th

This work was supported by Canadian Institutes of Health Research Grant MT-7034 (to A.D.B.). Y.S. was funded by The Lung Association Alberta and North West Territory and by a 75 Anniversary Graduate Student award from the Faculty of Medicine and Dentistry, University of Alberta. T.C.M. is a recipient of fellowships from the Korea Research Foundation (KRF-2005-214-C00111), the Alberta Heritage Foundation for Medical Research, and the Canadian Institute of Health Research/Canadian Lung Association/GlaxoSmithKline.

The online version of this article contains supplemental material.

Abbreviations used in this paper:

2-DE

two-dimensional electrophoresis

Aldo

anti-aldolase Ab

β-hex

β-hexosaminidase

CBMC

cord blood derived-mast cell

DAP

dihydroxyacetone phosphate

F1P

fructose-1-phosphate

F6P

fructose-6-phosphate

FBP

fructose 1,6-bisphosphate

IPG

immobilized pH gradient

IP3

inositol 1,4,5-triphosphate

Km

Michaelis–Menten constant

MC

mast cell

MS

mass spectrometry

MS/MS

tandem mass spectrometry

NMR

nuclear magnetic resonance

NT

anti-nitrotyrosine Ab

pI

isoelectric point

PLC

phospholipase C

PVDF

polyvinylidene difluoride

SNOG

S-nitrosoglutathione

Vmax

maximum velocity.

1
Moon
T. C.
,
St Laurent
C. D.
,
Morris
K. E.
,
Marcet
C.
,
Yoshimura
T.
,
Sekar
Y.
,
Befus
A. D.
.
2010
.
Advances in mast cell biology: new understanding of heterogeneity and function.
Mucosal Immunol.
3
:
111
128
.
2
Marshall
J. S.
2004
.
Mast-cell responses to pathogens.
Nat. Rev. Immunol.
4
:
787
799
.
3
Galli
S. J.
,
Nakae
S.
,
Tsai
M.
.
2005
.
Mast cells in the development of adaptive immune responses.
Nat. Immunol.
6
:
135
142
.
4
Swindle
E. J.
,
Metcalfe
D. D.
.
2007
.
The role of reactive oxygen species and nitric oxide in mast cell-dependent inflammatory processes.
Immunol. Rev.
217
:
186
205
.
5
Bidri
M.
,
Féger
F.
,
Varadaradjalou
S.
,
Ben Hamouda
N.
,
Guillosson
J. J.
,
Arock
M.
.
2001
.
Mast cells as a source and target for nitric oxide.
Int. Immunopharmacol.
1
:
1543
1558
.
6
Gilchrist
M.
,
McCauley
S. D.
,
Befus
A. D.
.
2004
.
Expression, localization, and regulation of NOS in human mast cell lines: effects on leukotriene production.
Blood
104
:
462
469
.
7
McCauley
S. D.
,
Gilchrist
M.
,
Befus
A. D.
.
2005
.
Nitric oxide: a major determinant of mast cell phenotype and function.
Mem. Inst. Oswaldo Cruz
100
(
Suppl. 1
):
11
14
.
8
Forsythe
P.
,
Gilchrist
M.
,
Kulka
M.
,
Befus
A. D.
.
2001
.
Mast cells and nitric oxide: control of production, mechanisms of response.
Int. Immunopharmacol.
1
:
1525
1541
.
9
Gilchrist
M.
,
Savoie
M.
,
Nohara
O.
,
Wills
F. L.
,
Wallace
J. L.
,
Befus
A. D.
.
2002
.
Nitric oxide synthase and nitric oxide production in in vivo-derived mast cells.
J. Leukoc. Biol.
71
:
618
624
.
10
Gilchrist
M.
,
Befus
A. D.
.
2008
.
Interferon-gamma regulates chemokine expression and release in the human mast cell line HMC1: role of nitric oxide.
Immunology
123
:
209
217
.
11
Yip
K. H.
,
Huang
Y.
,
Waye
M. M.
,
Lau
H. Y.
.
2008
.
Induction of nitric oxide synthases in primary human cultured mast cells by IgE and proinflammatory cytokines.
Int. Immunopharmacol.
8
:
764
768
.
12
Coleman
J. W.
2002
.
Nitric oxide: a regulator of mast cell activation and mast cell-mediated inflammation.
Clin. Exp. Immunol.
129
:
4
10
.
13
Nohara
O.
,
Kulka
M.
,
Déry
R. E.
,
Wills
F. L.
,
Hirji
N. S.
,
Gilchrist
M.
,
Befus
A. D.
.
2001
.
Regulation of CD8 expression in mast cells by exogenous or endogenous nitric oxide.
J. Immunol.
167
:
5935
5939
.
14
Bissonnette
E. Y.
,
Hogaboam
C. M.
,
Wallace
J. L.
,
Befus
A. D.
.
1991
.
Potentiation of tumor necrosis factor-alpha-mediated cytotoxicity of mast cells by their production of nitric oxide.
J. Immunol.
147
:
3060
3065
.
15
Moon
T. C.
,
Befus
A. D.
.
2008
.
Exogenous nitric oxide regulates cyclooxygenase-2 expression and prostaglandin D(2) generation through p38 MAPK in mouse bone marrow-derived mast cells.
Free Radic. Biol. Med.
45
:
780
788
.
16
Ischiropoulos
H.
,
Gow
A.
.
2005
.
Pathophysiological functions of nitric oxide-mediated protein modifications.
Toxicology
208
:
299
303
.
17
Gow
A. J.
,
Farkouh
C. R.
,
Munson
D. A.
,
Posencheg
M. A.
,
Ischiropoulos
H.
.
2004
.
Biological significance of nitric oxide-mediated protein modifications.
Am. J. Physiol. Lung Cell. Mol. Physiol.
287
:
L262
L268
.
18
Koeck
T.
,
Stuehr
D. J.
,
Aulak
K. S.
.
2005
.
Mitochondria and regulated tyrosine nitration.
Biochem. Soc. Trans.
33
:
1399
1403
.
19
Szabó
C.
,
Ischiropoulos
H.
,
Radi
R.
.
2007
.
Peroxynitrite: biochemistry, pathophysiology and development of therapeutics.
Nat. Rev. Drug Discov.
6
:
662
680
.
20
Schopfer
F. J.
,
Baker
P. R.
,
Freeman
B. A.
.
2003
.
NO-dependent protein nitration: a cell signaling event or an oxidative inflammatory response?
Trends Biochem. Sci.
28
:
646
654
.
21
Souza
J. M.
,
Peluffo
G.
,
Radi
R.
.
2008
.
Protein tyrosine nitration—functional alteration or just a biomarker?
Free Radic. Biol. Med.
45
:
357
366
.
22
Abello
N.
,
Kerstjens
H. A.
,
Postma
D. S.
,
Bischoff
R.
.
2009
.
Protein tyrosine nitration: selectivity, physicochemical and biological consequences, denitration, and proteomics methods for the identification of tyrosine-nitrated proteins.
J. Proteome Res.
8
:
3222
3238
.
23
Kim
J. Y.
,
Lee
K. H.
,
Lee
B. K.
,
Ro
J. Y.
.
2005
.
Peroxynitrite modulates release of inflammatory mediators from guinea pig lung mast cells activated by antigen-antibody reaction.
Int. Arch. Allergy Immunol.
137
:
104
114
.
24
Kirshenbaum
A. S.
,
Akin
C.
,
Wu
Y.
,
Rottem
M.
,
Goff
J. P.
,
Beaven
M. A.
,
Rao
V. K.
,
Metcalfe
D. D.
.
2003
.
Characterization of novel stem cell factor responsive human mast cell lines LAD 1 and 2 established from a patient with mast cell sarcoma/leukemia; activation following aggregation of FcepsilonRI or FcgammaRI.
Leuk. Res.
27
:
677
682
.
25
Moon
T. C.
,
Lee
E.
,
Baek
S. H.
,
Murakami
M.
,
Kudo
I.
,
Kim
N. S.
,
Lee
J. M.
,
Min
H. K.
,
Kambe
N.
,
Chang
H. W.
.
2003
.
Degranulation and cytokine expression in human cord blood-derived mast cells cultured in serum-free medium with recombinant human stem cell factor.
Mol. Cells
16
:
154
160
.
26
Kambe
N.
,
Kambe
M.
,
Chang
H. W.
,
Matsui
A.
,
Min
H. K.
,
Hussein
M.
,
Oskerizian
C. A.
,
Kochan
J.
,
Irani
A. A.
,
Schwartz
L. B.
.
2000
.
An improved procedure for the development of human mast cells from dispersed fetal liver cells in serum-free culture medium.
J. Immunol. Methods
240
:
101
110
.
27
Askew
S. C.
,
Butler
A. R.
,
Flitney
F. W.
,
Kemp
G. D.
,
Megson
I. L.
.
1995
.
Chemical mechanisms underlying the vasodilator and platelet anti-aggregating properties of S-nitroso-N-acetyl-DL-penicillamine and S-nitrosoglutathione.
Bioorg. Med. Chem.
3
:
1
9
.
28
Gordge
M. P.
,
Hothersall
J. S.
,
Noronha-Dutra
A. A.
.
1998
.
Evidence for a cyclic GMP-independent mechanism in the anti-platelet action of S-nitrosoglutathione.
Br. J. Pharmacol.
124
:
141
148
.
29
Koranteng
R. D.
,
Dearman
R. J.
,
Kimber
I.
,
Coleman
J. W.
.
2000
.
Phenotypic variation in mast cell responsiveness to the inhibitory action of nitric oxide.
Inflamm. Res.
49
:
240
246
.
30
Davis
B. J.
,
Flanagan
B. F.
,
Gilfillan
A. M.
,
Metcalfe
D. D.
,
Coleman
J. W.
.
2004
.
Nitric oxide inhibits IgE-dependent cytokine production and Fos and Jun activation in mast cells.
J. Immunol.
173
:
6914
6920
.
31
Forsythe
P.
,
Befus
A. D.
.
2003
.
Inhibition of calpain is a component of nitric oxide-induced down-regulation of human mast cell adhesion.
J. Immunol.
170
:
287
293
.
32
Chazotte-Aubert
L.
,
Hainaut
P.
,
Ohshima
H.
.
2000
.
Nitric oxide nitrates tyrosine residues of tumor-suppressor p53 protein in MCF-7 cells.
Biochem. Biophys. Res. Commun.
267
:
609
613
.
33
Springer
T. A.
1995
.
Immunoaffinity chromatography.
In
Current Protocols in Molecular Biology.
Ausubel
F. M.
,
Brent
R.
,
Kingston
R. E.
,
Moore
D. D.
,
Seidman
J. G.
,
Smith
J. A.
,
Struhl
K.
,
John Wiley & Sons, Inc.
,
Hoboken, NJ
,
Vol. 2.
10.11.1
10.11.6
.
34
Michal
G.
1984
.
D-Fructose 1,6-Bisphosphate, Dihydroxyacetone Phosphate and D-Glyceraldehyde 3-Phosphate.
Academic Press
,
New York
.
35
Koeck
T.
,
Levison
B.
,
Hazen
S. L.
,
Crabb
J. W.
,
Stuehr
D. J.
,
Aulak
K. S.
.
2004
.
Tyrosine nitration impairs mammalian aldolase A activity.
Mol. Cell. Proteomics
3
:
548
557
.
36
Xu
K.
,
Stringer
J. L.
.
2008
.
Pharmacokinetics of fructose-1,6-diphosphate after intraperitoneal and oral administration to adult rats.
Pharmacol. Res.
57
:
234
238
.
37
Hohman
R. J.
,
Dreskin
S. C.
.
2001
.
Measuring degranulation of mast cells.
Curr. Protoc. Immunol.
Chapter 7:Unit 7.26
.
38
Bligh
E. G.
,
Dyer
W. J.
.
1959
.
A rapid method of total lipid extraction and purification.
Can. J. Biochem. Physiol.
37
:
911
917
.
39
Weljie
A. M.
,
Newton
J.
,
Mercier
P.
,
Carlson
E.
,
Slupsky
C. M.
.
2006
.
Targeted profiling: quantitative analysis of 1H NMR metabolomics data.
Anal. Chem.
78
:
4430
4442
.
40
Wishart
D. S.
,
Tzur
D.
,
Knox
C.
,
Eisner
R.
,
Guo
A. C.
,
Young
N.
,
Cheng
D.
,
Jewell
K.
,
Arndt
D.
,
Sawhney
S.
, et al
.
2007
.
HMDB: the Human Metabolome Database.
Nucleic Acids Res.
35
(
Database issue
):
D521
D526
.
41
Hardin
C. D.
,
Roberts
T. M.
.
1994
.
Metabolism of exogenously applied fructose 1,6-bisphosphate in hypoxic vascular smooth muscle.
Am. J. Physiol.
267
:
H2325
H2332
.
42
Ehringer
W. D.
,
Niu
W.
,
Chiang
B.
,
Wang
O. L.
,
Gordon
L.
,
Chien
S.
.
2000
.
Membrane permeability of fructose-1,6-diphosphate in lipid vesicles and endothelial cells.
Mol. Cell. Biochem.
210
:
35
45
.
43
Sekar
Y.
,
Moon
T. C.
,
Muñoz
S.
,
Befus
A. D.
.
2005
.
Role of nitric oxide in mast cells: controversies, current knowledge, and future applications.
Immunol. Res.
33
:
223
239
.
44
Krumenacker
J. S.
,
Hanafy
K. A.
,
Murad
F.
.
2004
.
Regulation of nitric oxide and soluble guanylyl cyclase.
Brain Res. Bull.
62
:
505
515
.
45
Deschoolmeester
M. L.
,
Eastmond
N. C.
,
Dearman
R. J.
,
Kimber
I.
,
Basketter
D. A.
,
Coleman
J. W.
.
1999
.
Reciprocal effects of interleukin-4 and interferon-gamma on immunoglobulin E-mediated mast cell degranulation: a role for nitric oxide but not peroxynitrite or cyclic guanosine monophosphate.
Immunology
96
:
138
144
.
46
Palazzolo-Ballance
A. M.
,
Suquet
C.
,
Hurst
J. K.
.
2007
.
Pathways for intracellular generation of oxidants and tyrosine nitration by a macrophage cell line.
Biochemistry
46
:
7536
7548
.
47
Takemoto
K.
,
Ogino
K.
,
Wang
D. H.
,
Takigawa
T.
,
Kurosawa
C. M.
,
Kamyabashi
Y.
,
Hibino
Y.
,
Hitomi
Y.
,
Ichimura
H.
.
2007
.
Biochemical characterization of reactive nitrogen species by eosinophil peroxidase in tyrosine nitration.
Acta Med. Okayama
61
:
17
30
.
48
Ulrich
M.
,
Petre
A.
,
Youhnovski
N.
,
Promm
F.
,
Schirle
M.
,
Schumm
M.
,
Pero
R. S.
,
Doyle
A.
,
Checkel
J.
,
Kita
H.
, et al
.
2008
.
Post-translational tyrosine nitration of eosinophil granule toxins mediated by eosinophil peroxidase.
J. Biol. Chem.
283
:
28629
28640
.
49
Klink
M.
,
Bednarska
K.
,
Jastrzembska
K.
,
Banasik
M.
,
Sulowska
Z.
.
2007
.
Signal transduction pathways affected by nitric oxide donors during neutrophil functional response in vitro.
Inflamm. Res.
56
:
282
290
.
50
Vassilakopoulos
T.
,
Govindaraju
K.
,
Parthenis
D.
,
Eidelman
D. H.
,
Watanabe
Y.
,
Hussain
S. N.
.
2007
.
Nitric oxide production in the ventilatory muscles in response to acute resistive loading.
Am. J. Physiol. Lung Cell. Mol. Physiol.
292
:
L1013
L1022
.
51
Dukes
A. A.
,
Van Laar
V. S.
,
Cascio
M.
,
Hastings
T. G.
.
2008
.
Changes in endoplasmic reticulum stress proteins and aldolase A in cells exposed to dopamine.
J. Neurochem.
106
:
333
346
.
52
Arcelay
E.
,
Salicioni
A. M.
,
Wertheimer
E.
,
Visconti
P. E.
.
2008
.
Identification of proteins undergoing tyrosine phosphorylation during mouse sperm capacitation.
Int. J. Dev. Biol.
52
:
463
472
.
53
Schindler
R.
,
Weichselsdorfer
E.
,
Wagner
O.
,
Bereiter-Hahn
J.
.
2001
.
Aldolase-localization in cultured cells: cell-type and substrate-specific regulation of cytoskeletal associations.
Biochem. Cell Biol.
79
:
719
728
.
54
MacDonnell
P. C.
,
Greengard
O.
.
1974
.
Enzymes in intracellular organelles of adult and developing rat brain.
Arch. Biochem. Biophys.
163
:
644
655
.
55
Mamczur
P.
,
Dzugaj
A.
.
2004
.
Nuclear localization of aldolase A in pig cardiomyocytes.
Histol. Histopathol.
19
:
753
758
.
56
Heijnen
H. F.
,
van Donselaar
E.
,
Slot
J. W.
,
Fries
D. M.
,
Blachard-Fillion
B.
,
Hodara
R.
,
Lightfoot
R.
,
Polydoro
M.
,
Spielberg
D.
,
Thomson
L.
, et al
.
2006
.
Subcellular localization of tyrosine-nitrated proteins is dictated by reactive oxygen species generating enzymes and by proximity to nitric oxide synthase.
Free Radic. Biol. Med.
40
:
1903
1913
.
57
Haqqani
A. S.
,
Kelly
J. F.
,
Birnboim
H. C.
.
2002
.
Selective nitration of histone tyrosine residues in vivo in mutatect tumors.
J. Biol. Chem.
277
:
3614
3621
.
58
Miller
S.
,
Ross-Inta
C.
,
Giulivi
C.
.
2007
.
Kinetic and proteomic analyses of S-nitrosoglutathione-treated hexokinase A: consequences for cancer energy metabolism.
Amino Acids
32
:
593
602
.
59
Lebherz
H. G.
,
Rutter
W. J.
.
1969
.
Distribution of fructose diphosphate aldolase variants in biological systems.
Biochemistry
8
:
109
121
.
60
Yañez
A. J.
,
Ludwig
H. C.
,
Bertinat
R.
,
Spichiger
C.
,
Gatica
R.
,
Berlien
G.
,
Leon
O.
,
Brito
M.
,
Concha
I. I.
,
Slebe
J. C.
.
2005
.
Different involvement for aldolase isoenzymes in kidney glucose metabolism: aldolase B but not aldolase A colocalizes and forms a complex with FBPase.
J. Cell. Physiol.
202
:
743
753
.
61
Takashi
M.
,
Haimoto
H.
,
Koshikawa
T.
,
Kato
K.
.
1990
.
Expression of aldolase C isozyme in renal cell carcinoma.
Am. J. Clin. Pathol.
93
:
631
636
.
62
Miyagi
M.
,
Sakaguchi
H.
,
Darrow
R. M.
,
Yan
L.
,
West
K. A.
,
Aulak
K. S.
,
Stuehr
D. J.
,
Hollyfield
J. G.
,
Organisciak
D. T.
,
Crabb
J. W.
.
2002
.
Evidence that light modulates protein nitration in rat retina.
Mol. Cell. Proteomics
1
:
293
303
.
63
Kanski
J.
,
Behring
A.
,
Pelling
J.
,
Schöneich
C.
.
2005
.
Proteomic identification of 3-nitrotyrosine-containing rat cardiac proteins: effects of biological aging.
Am. J. Physiol. Heart Circ. Physiol.
288
:
H371
H381
.
64
Aulak
K. S.
,
Miyagi
M.
,
Yan
L.
,
West
K. A.
,
Massillon
D.
,
Crabb
J. W.
,
Stuehr
D. J.
.
2001
.
Proteomic method identifies proteins nitrated in vivo during inflammatory challenge.
Proc. Natl. Acad. Sci. USA
98
:
12056
12061
.
65
Ghosh
S.
,
Janocha
A. J.
,
Aronica
M. A.
,
Swaidani
S.
,
Comhair
S. A.
,
Xu
W.
,
Zheng
L.
,
Kaveti
S.
,
Kinter
M.
,
Hazen
S. L.
,
Erzurum
S. C.
.
2006
.
Nitrotyrosine proteome survey in asthma identifies oxidative mechanism of catalase inactivation.
J. Immunol.
176
:
5587
5597
.
66
Parastatidis
I.
,
Thomson
L.
,
Burke
A.
,
Chernysh
I.
,
Nagaswami
C.
,
Visser
J.
,
Stamer
S.
,
Liebler
D. C.
,
Koliakos
G.
,
Heijnen
H. F.
, et al
.
2008
.
Fibrinogen beta-chain tyrosine nitration is a prothrombotic risk factor.
J. Biol. Chem.
283
:
33846
33853
.
67
Steinhagen-Thiessen
E.
,
Hilz
H.
.
1979
.
Aldolase activity and cross-reacting material in lymphocytes of aged individuals.
Gerontology
25
:
132
135
.
68
Day
I. N.
,
Thompson
R. J.
.
1984
.
Levels of immunoreactive aldolase C, creatine kinase-BB, neuronal and non-neuronal enolase, and 14-3-3 protein in circulating human blood cells.
Clin. Chim. Acta
136
:
219
228
.
69
Jean
J. C.
,
Rich
C. B.
,
Joyce-Brady
M.
.
2006
.
Hypoxia results in an HIF-1-dependent induction of brain-specific aldolase C in lung epithelial cells.
Am. J. Physiol. Lung Cell. Mol. Physiol.
291
:
L950
L956
.
70
Slemmer
J. E.
,
Haasdijk
E. D.
,
Engel
D. C.
,
Plesnila
N.
,
Weber
J. T.
.
2007
.
Aldolase C-positive cerebellar Purkinje cells are resistant to delayed death after cerebral trauma and AMPA-mediated excitotoxicity.
Eur. J. Neurosci.
26
:
649
656
.
71
Mesplès
B.
,
Fontaine
R. H.
,
Lelièvre
V.
,
Launay
J. M.
,
Gressens
P.
.
2005
.
Neuronal TGF-beta1 mediates IL-9/mast cell interaction and exacerbates excitotoxicity in newborn mice.
Neurobiol. Dis.
18
:
193
205
.
72
Barreiro
E.
,
Rabinovich
R.
,
Marin-Corral
J.
,
Barberà
J. A.
,
Gea
J.
,
Roca
J.
.
2009
.
Chronic endurance exercise induces quadriceps nitrosative stress in patients with severe COPD.
Thorax
64
:
13
19
.
73
Madej
J. A.
,
Sobiech
K. A.
,
Klimentowski
S.
.
1989
.
In vitro studies into some parameters of protein and carbohydrate metabolism in lymphocytes infected with bovine leucosis virus.
Arch. Exp. Veterinarmed.
43
:
923
927
.
74
Takahashi
I.
,
Takasaki
Y.
,
Hori
K.
.
1989
.
Site-directed mutagenesis of human aldolase isozymes: the role of Cys-72 and Cys-338 residues of aldolase A and of the carboxy-terminal Tyr residues of aldolases A and B.
J. Biochem.
105
:
281
286
.
75
Kreuder
J.
,
Borkhardt
A.
,
Repp
R.
,
Pekrun
A.
,
Göttsche
B.
,
Gottschalk
U.
,
Reichmann
H.
,
Schachenmayr
W.
,
Schlegel
K.
,
Lampert
F.
.
1996
.
Brief report: inherited metabolic myopathy and hemolysis due to a mutation in aldolase A.
N. Engl. J. Med.
334
:
1100
1104
.
76
Sola
A.
,
Roselló-Catafau
J.
,
Gelpí
E.
,
Hotter
G.
.
2001
.
Fructose-1,6-biphosphate in rat intestinal preconditioning: involvement of nitric oxide.
Gut
48
:
168
175
.
77
Hardie
D. G.
,
Carling
D.
.
1997
.
The AMP-activated protein kinase—fuel gauge of the mammalian cell?
Eur. J. Biochem.
246
:
259
273
.
78
Carling
D.
2004
.
AMPK.
Curr. Biol.
14
:
R220
.
79
Sonnewald
U.
,
Müller
T. B.
,
Westergaard
N.
,
Unsgård
G.
,
Petersen
S. B.
,
Schousboe
A.
.
1994
.
NMR spectroscopic study of cell cultures of astrocytes and neurons exposed to hypoxia: compartmentation of astrocyte metabolism.
Neurochem. Int.
24
:
473
483
.
80
Almeida
A.
,
Almeida
J.
,
Bolaños
J. P.
,
Moncada
S.
.
2001
.
Different responses of astrocytes and neurons to nitric oxide: the role of glycolytically generated ATP in astrocyte protection.
Proc. Natl. Acad. Sci. USA
98
:
15294
15299
.
81
Almeida
A.
,
Moncada
S.
,
Bolaños
J. P.
.
2004
.
Nitric oxide switches on glycolysis through the AMP protein kinase and 6-phosphofructo-2-kinase pathway.
Nat. Cell Biol.
6
:
45
51
.
82
Donohoe
P. H.
,
Fahlman
C. S.
,
Bickler
P. E.
,
Vexler
Z. S.
,
Gregory
G. A.
.
2001
.
Neuroprotection and intracellular Ca2+ modulation with fructose-1,6-bisphosphate during in vitro hypoxia-ischemia involves phospholipase C-dependent signaling.
Brain Res.
917
:
158
166
.
83
Cohly
H.
,
Jenkins
J.
,
Skelton
T.
,
Meydrech
E.
,
Markov
A. K.
.
2004
.
Fructose-1,6-diphosphate suppresses T-lymphocyte proliferation, promotes apoptosis and inhibits interleukins-1, 6, beta-actin mRNAs, and transcription factors expression.
Immunol. Invest.
33
:
407
421
.
84
Moresco
R. N.
,
Santos
R. C.
,
Alves Filho
J. C.
,
Cunha
A. A.
,
Dos Reis
C.
,
Reichel
C. L.
,
De Oliveira
J. R.
.
2004
.
Protective effect of fructose-1,6-bisphosphate in the cold storage solution for liver preservation in rat hepatic transplantation.
Transplant. Proc.
36
:
1261
1264
.
85
Markov
A. K.
Treatment of asthma with fructose-1,6-diphosphate.
U.S. Patent 5858985 1999 Jan 12
86
Cuesta
E.
,
Boada
J.
,
Calafell
R.
,
Perales
J. C.
,
Roig
T.
,
Bermudez
J.
.
2006
.
Fructose 1,6-bisphosphate prevented endotoxemia, macrophage activation, and liver injury induced by D-galactosamine in rats.
Crit. Care Med.
34
:
807
814
.
87
Yip
K.
,
Leung
F.
,
Huang
Y.
,
Lau
H.
.
2009
.
Inhibition of anti-IgE mediated human mast cell activation by NO donors is dependent on their NO release kinetics.
Br. J. Pharmacol.
156
:
1279
1286
.
88
Gilfillan
A. M.
,
Tkaczyk
C.
.
2006
.
Integrated signalling pathways for mast-cell activation.
Nat. Rev. Immunol.
6
:
218
230
.
89
Baron
C. B.
,
Ozaki
S.
,
Watanabe
Y.
,
Hirata
M.
,
LaBelle
E. F.
,
Coburn
R. F.
.
1995
.
Inositol 1,4,5-trisphosphate binding to porcine tracheal smooth muscle aldolase.
J. Biol. Chem.
270
:
20459
20465
.
90
Koppitz
B.
,
Vogel
F.
,
Mayr
G. W.
.
1986
.
Mammalian aldolases are isomer-selective high-affinity inositol polyphosphate binders.
Eur. J. Biochem.
161
:
421
433
.
91
Waingeh
V. F.
,
Gustafson
C. D.
,
Kozliak
E. I.
,
Lowe
S. L.
,
Knull
H. R.
,
Thomasson
K. A.
.
2006
.
Glycolytic enzyme interactions with yeast and skeletal muscle F-actin.
Biophys. J.
90
:
1371
1384
.
92
Seo
I. R.
,
Moh
S. H.
,
Lee
E. H.
,
Meissner
G.
,
Kim
do H.
.
2006
.
Aldolase potentiates DIDS activation of the ryanodine receptor in rabbit skeletal sarcoplasmic reticulum.
Biochem. J.
399
:
325
333
.