Using ratiometric Ca2+ imaging and patch-clamp measurement of Ca2+ channel activity, we investigated Ca2+ signaling induced by vanadium compounds in Jurkat T lymphocytes and rat basophilic leukemia cells. In the presence of external Ca2+, vanadium compounds produced sustained or oscillatory Ca2+ elevations; in nominally Ca2+-free medium, a transient Ca2+ rise was generated. Vanadate-induced Ca2+ signaling was blocked by heparin, a competitive inhibitor of the 1,4,5-inositol trisphosphate (IP3) receptor, suggesting that Ca2+ influx is secondary to depletion of IP3-sensitive Ca2+ stores. In Jurkat T cells, vanadate also activated the Ca2+-dependent transcription factor, NF-AT. Intracellular dialysis with vanadate activated Ca2+ influx through Ca2+ release-activated Ca2+ (CRAC) channels with kinetics comparable to those of dialysis with IP3. Neither phosphatase inhibitors nor nonhydrolyzable nucleotide analogues modified CRAC channel activation. The action of vanadate, but not IP3, was prevented by the thiol-reducing agent DTT. In addition, the activation of CRAC channels by vanadate was mimicked by the thiol-oxidizing agent chloramine T. These results suggest that vanadate enhances Ca2+ signaling via thiol oxidation of a proximal element in the signal transduction cascade.

The role of Ca2+ signaling in the activation of T lymphocytes is well established (1, 2). Binding of the TCR to Ag associated with the MHC initiates a cascade of intracellular signaling that culminates in the production of cytokines and T cell proliferation. Among the proximal signaling steps that take place at the plasma membrane, tyrosine phosphorylation of substrates, including phospholipase C, generates two key second messengers, diacylglycerol, leading to the activation of protein kinase C, and 1,4,5-inositol trisphosphate (IP3),5 leading to the initiation of the Ca2+ signal. The Ca2+ signal is sustained by the activation of a specific type of store-operated Ca2+ channel in the plasma membrane called the Ca2+ release-activated Ca2+ (CRAC) channel (1). Ca2+ signaling must be oscillatory or sustained at the level of individual cells to activate gene transcription via the NF-AT pathway, leading to cytokine production (3, 4, 5).

The level of cellular tyrosine phosphorylation and subsequent activation events are controlled by a balance between phosphorylation and dephosphorylation reactions. Thus, compounds that alter the balance of kinase and phosphatase activity may profoundly influence cellular activation in the immune response. Sodium orthovanadate (Na3VO4) is a potent inhibitor of protein tyrosine phosphatase (PTPs) and other phosphoryl transfer enzymes (6). In the +5 oxidation state, vanadate adopts a trigonal bipyramidal configuration that mimics the phosphate ion (7). In this configuration, Na3VO4 competes with the γ-phosphate of ATP for binding to the active site of the Na+,K+-ATPase, thus inhibiting the enzyme (8, 9). Na3VO4 inhibits PTPs by a similar mechanism (reviewed in Refs. 10, 11, 12). Formation of a phosphate-thioester bond with a conserved cysteine at the active site of PTPs is critical to the dephosphorylation reaction. Na3VO4 reversibly displaces phosphate from the active site thiol (13). In addition, thiol oxidation may permanently inactivate PTPs during prolonged treatment of cells with Na3VO4 (14). Cysteine could be oxidized directly by Na3VO4 or indirectly following the intracellular formation of pervanadate. Pervandate is a complex mixture of peroxovanadium compounds formed in vitro by mixing H2O2 and Na3VO4 (reviewed in Ref. 10). Pervanadate produces an irreversible inhibition of PTPs that can be prevented by preincubation with thiol-reducing agents, suggesting that the effect of pervanadate is mediated by the oxidation of cysteine (15). Although vanadate and pervanadate have been widely used as general inhibitors of PTPs and clearly enhance tyrosine phosphorylation in immune cells (16, 17, 18, 19, 20), the effects of vanadium compounds on other cellular targets as a result of their thiol reactivity have not been fully explored.

The modulation of intracellular signaling by vanadium compounds takes on a greater significance because they are being considered as potential therapeutic agents for use in both insulin-dependent and noninsulin-dependent forms of diabetes (21, 22, 23, 24, 25). Their insulin-mimetic actions in vivo and in vitro correlate with their ability to inhibit PTPs and induce phosphorylation of the β subunit of the insulin receptor (26, 27, 28, 29, 30). Recent studies have shown that concentrations of vanadium compounds that increase glucose transport and normalize blood glucose levels (21, 31, 32) also produce tyrosine phosphorylation and Ca2+ mobilization in Jurkat T cells (16, 17, 33) and RBL cells (18). Thus, potential actions by these compounds on the immune system must also be considered.

Previous attempts to characterize the action of vanadate on cells of the immune system have yielded conflicting results. In Jurkat T cells, pervanadate was reported to induce Ca2+ influx independent of store release (17). However, in RBL cells pervanadate appeared to release Ca2+ from intracellular stores (18). In many cell types, Ca2+ influx is coupled to the depletion of intracellular Ca2+ stores by a process termed capacitative Ca2+ entry (34, 35). In lymphocytes and RBL cells store release leads to CRAC channel activation (36, 37, 38, 39, 40, 41, 42). To clarify the mechanisms of action of vanadate, we have investigated its effects on [Ca2+]i signaling in Jurkat T lymphocytes and RBL cells using single-cell imaging and electrophysiological techniques. Our experiments were designed to show whether the normal Ca2+ signaling mechanisms used by cells of the immune system are usurped by vanadate stimulation, whether Ca2+ channel activation is direct or secondary to release from Ca2+ stores, and whether nonphysiological mechanisms are evoked.

The human leukemia T cell line, Jurkat E6-1, and the rat basophilic leukemia cell line, RBL-2H3, were obtained from American Type Culture Collection (Manassas, VA). Jurkat-NZ cells containing a β-galactosidase reporter gene construct (lacZ) under the control of the NF-AT promoter were derived from the Jurkat subline NZDipA.1.5.22 and have been previously described and characterized (4). Jurkat E6-1 and Jurkat-NZ cells were grown in RPMI 1640 medium containing 10% heat-inactivated FBS, 10 mM HEPES, and 2 mM glutamine. Cells were cultured in 25-ml flasks (Costar, Cambridge, MA) at 37°C in 5% CO2 in a humidified incubator. RBL cells were maintained in Eagle’s MEM supplemented with 20% FCS and 2 mM glutamine. RBL cells were plated onto glass coverslips 1–2 days before use.

Thapsigargin (TG) and dipotassium oxodiperoxo (pyridine-2-carboxylato) vanadate (bpV(pic)) were obtained from Alexis Biochemicals (San Diego, CA). 5′-Adenlylimidodiphosphate tetralithium salt was obtained from Roche Molecular Biochemicals (Basel, Switzerland), and adenosine 5′-o-(3-thiotriphosphate) tetralithium salt (ATPγS) was obtained from Calbiochem (San Diego, CA). All other salts and reagents were obtained from Sigma-Aldrich (St. Louis, MO).

Jurkat T cells were loaded with 3 μM fura-2-AM (Molecular Probes, Eugene, OR) for 30–40 min at room temperature (20–25°C). The cells were then washed three times with RPMI/10% FCS. RBL cells were loaded in medium containing 1 μM fura-2-AM and 2.5 mM probenecid for 30 min at 37°C in the incubator. These cells were then washed three times with Eagle’s MEM containing 2.5 mM probenecid and returned to the incubator for an additional 30 min to complete the hydrolysis of the fura-2 ester. After fura-2 loading, all cells were stored at room temperature in the dark until use. Before imaging, media were exchanged with mammalian Ringer containing: 160 mM NaCl, 4.5 mM KCl, 2 mM CaCl2, 1 mM MgCl2, 10 mM glucose, and 10 mM HEPES (pH 7.4; osmolality, 290–310 mOsm/kg). Illumination was provided by a xenon arc lamp (Carl Zeiss, New York, NY) and transmitted through a filter wheel unit (Lambda 10, Axon Instruments, Foster City, CA) containing 350- and 385-nm excitation filters. The filtered light was reflected by a 400-nm dichroic mirror through a ×63 oil immersion objective to illuminate cells. Emitted light above 480 nm was received by an intensified CCD camera (C2400, Hamamatsu Photonics, Bridgewater, NJ), and the video information was relayed to an image processing system (Videoprobe, ETM Systems, Petaluma, CA). Full field-of-view 8-bit images, averaged over 16 frames, were collected at 350- and 385-nm wavelengths. Digitally stored 350/385 ratios were constructed from background-corrected 350- and 385-nm images. Single-cell measurements of [Ca2+]i were calculated from the 350/385 ratios using the equation of Grynkiewicz et al. (43) and a Kd of 250 nM for fura-2. The minimum 350/385 ratio was measured in single cells after incubation for 10 min in Ca2+-free Ringer containing 2 mM EGTA. Maximum ratio values were obtained after perfusion with Ringer containing 10 mM Ca2+, 1 μM TG, and 10 μM ionomycin.

The sodium salt of low molecular mass heparin (average m.w., 3 kDa; Sigma, St. Louis, MO) was made up to a concentration of 25 mg/ml in a 100 mM KCl solution containing 0.1% tetramethylrhodamine dextran (Molecular Probes). The heparin solution was loaded into prepulled glass capillaries (Femtotip, Eppendorf, Hamburg, Germany). Cell microinjection was performed using a pressure injector (model 5246, Eppendorf) and micromanipulator (model 5171, Eppendorf) as previously described (44). Microinjections were performed 2 h before the calcium imaging experiments.

Membrane currents were measured in Jurkat T cells using the whole-cell configuration of the patch-clamp technique (45, 46). An EPC-9 amplifier (HEKA, Lambrecht, Germany) interfaced to a Macintosh Quadra 700 computer (Apple Computer, Cupertino, CA) was used for pulse application and data recording. Membrane voltages were corrected for liquid junction potentials, and current recordings were corrected for leak and capacitative currents. Patch pipettes were pulled from Accu-fill 90 Micropets (Becton Dickinson, Parsippany, NJ) using a P87 micropipette puller (Sutter Instruments, Novato, CA). Pipettes were coated with Sylgard (Dow Corning, Midland, MI) and heat polished to a final resistances of 2–5 MΩ. Patch-clamp experiments were performed at room temperature (20–25°C). Unless otherwise indicated the membrane currents were filtered at 1.5 kHz. Data analysis was performed using the program Pulse (HEKA, Lambrecht, Germany). During the whole-cell recordings, the membrane potential was held at 0 mV, and the Ca2+ current through CRAC channels was measured during 200-ms voltage ramps from −100 to +40 mV applied at 1-s intervals. The pipette solution contained 128 mM Cs+ aspartate, 12 mM bis(2-aminophenoxy)ethane-N,N,N′,N′-tetraacetate (BAPTA), 0.9 mM CaCl2, 3.16 mM MgCl2, and 10 mM HEPES; was titrated to pH 7.2 with CsOH; and had an estimated free Ca2+ concentration of 10 nM. The external solution had the following composition: 150 mM NaMeSO3, 20 mM CaCl2, 10 mM glucose, and 10 mM HEPES and was titrated to pH 7.4 with NaOH.

The expression of lacZ was measured using a fluorescence assay as previously described (47). In brief, Jurkat-NZ cells were seeded at 1 × 105 cells/well in 96-well plates containing culture medium alone or with addition of 1 μM TG or varying concentrations of bpV(pic), each condition in the presence or the absence of 50 nM PMA. The cells were activated in a humidified incubator at 37°C with 5% CO2 for 10 h, then lysed and incubated with 3 mM 4-methylumbelliferyl β-d-galactopyranoside (Molecular Probes). The fluorescence signal measured in a multiwell plate reader (CytoFluor Series 4000, Perspective Biosystems, Framingham, MA; 400 nm excitation, 505 nm emission) is proportional to lacZ activity driven by the NF-AT pathway.

Numerical values for single-cell [Ca2+]i traces were analyzed with Origin (Microcal, Northhampton, MA). Statistical analysis was performed on data sets using Excel v5.0 (Microsoft, Redmond, WA). Data are reported as the mean ± SD. Multisample hypotheses were tested using a single-factor ANOVA and performing a Dunnett’s test to determine the significance of differences from control values. Data were considered statistically different at p < 0.05.

Previous studies (17, 18) demonstrated that application of peroxovanadium compounds increased [Ca2+]i in both RBL cells and Jurkat T cells. However, different mechanisms have been postulated for the Ca2+ increase in each cell type. These differences may have been due to cell-specific effects of pervanadate or to different mixtures of peroxyvanadium compounds being used each study. To test the hypothesis that peroxyvanadium compounds act by a similar mechanism on T cells and mast cells we tested the effects of the cell-permeant peroxyvanadium compound, bpV(pic) on Jurkat T cells and RBL mast cells. When exposed to bpV(pic) (200 μM) in the presence of extracellular Ca2+, most Jurkat T cells (80 ± 10%) responded after a variable delay with an abrupt increase in [Ca2+]i. In the majority of responding cells the initial rise in [Ca2+]i was followed by slow oscillations with a 73 ± 4-s period and a peak [Ca2+]i near 700 nM (Fig. 1,A). The average Ca2+ response from all observed cells (Fig. 1,A, bottom graph) rose smoothly from the resting level of 66 ± 41 nM to a plateau of 488 ± 210 nM. This population response reflects the asynchronous Ca2+ oscillations in single cells and the fact that the average includes nonresponding cells. Our data confirm earlier observations that peroxovanadium compounds increase [Ca2+]i in Jurkat T cells (16, 17) and reveal at the single-cell level that bpV(pic) can produce Ca2+ oscillations analogous to those produced during Ag presentation (48). Extracellular Ca2+ is required to maintain the bpV(pic)-induced oscillations, because exchanging the bath solution with nominally Ca2+-free Ringer’s solution rapidly and reversibly terminated the oscillations (Fig. 1 A, application bar).

FIGURE 1.

The [Ca2+]i signaling induced by bpV(pic). [Ca2+]i was measured as described in Materials and Methods in Jurkat T cells (A) and RBL cells (B). Resting [Ca2+]i was determined in mammalian Ringer. At 300 s, 200 μM bpV(pic) was added to the Ringer (arrow). As illustrated (application bar), the Ca2+ concentration of the Ringer was varied between 2 mM (2 Ca2+) and nominally Ca2+ free (0 Ca2+). Traces for four representative cells (top four graphs) and the average of all cells within the field of view (bottom graph) are shown.

FIGURE 1.

The [Ca2+]i signaling induced by bpV(pic). [Ca2+]i was measured as described in Materials and Methods in Jurkat T cells (A) and RBL cells (B). Resting [Ca2+]i was determined in mammalian Ringer. At 300 s, 200 μM bpV(pic) was added to the Ringer (arrow). As illustrated (application bar), the Ca2+ concentration of the Ringer was varied between 2 mM (2 Ca2+) and nominally Ca2+ free (0 Ca2+). Traces for four representative cells (top four graphs) and the average of all cells within the field of view (bottom graph) are shown.

Close modal

Treatment of RBL cells with pervanadate results in Ca2+ responses that vary with the peroxovanadium compounds formed in the mixture (18). Using bpV(pic), we characterized the effects of a single form of peroxovanadate on RBL cells. The average response of RBL cells to 200 μM bpV(pic) closely resembled that of the Jurkat T cells (Fig. 1,B). The [Ca2+]i began to increase 200 s after the application of bpV(pic) and gradually approached a plateau of 500 ± 170 nM after an additional 1200 s. The sustained Ca2+ plateau required Ca2+ influx, because [Ca2+]i returned to baseline during perfusion with nominally Ca2+-free Ringer’s solution (Fig. 1 B, application bar). Examination of the responses of individual cells revealed that RBL cells were less prone to oscillate than the Jurkat T cells and appeared to have a biphasic response to bpV(pic) consisting of an initial transient [Ca2+]i rise followed by a secondary [Ca2+]i increase to a plateau.

To separate the contributions of Ca2+ release and Ca2+ influx, we applied bpV(pic) in nominally Ca2+-free Ringer’s solution. For both Jurkat T and RBL cells, single-cell records show that application of bpV(pic) elicited an abrupt [Ca2+]i transient following a variable delay averaging 257 ± 177 s in Jurkat T cells and 579 ± 260 s in RBL cells (Fig. 2). In the averaged response (Fig. 2, bottom graphs), the Ca2+ transient appeared blunted due to the variation in response times between cells. In some experiments, the variation in latency resulted in the masking of initial [Ca2+]i transients in the averaged response. However, analysis of the single-cell responses clearly showed that bpV(pic) releases Ca2+ from intracellular stores. Ca2+ readdition to Ca2+-depleted cells caused [Ca2+]i to increase rapidly to a plateau level (750 ± 40 nM in Jurkat T cells and 530 ± 140 nM in RBL cells). Thus, bpV(pic) appears to release Ca2+ from intracellular stores and induce capacitative Ca2+ influx. To determine the extent of Ca2+ store depletion by bpV(pic), we probed for residual Ca2+ in the intracellular stores with TG, a specific inhibitor of the microsomal Ca2+-ATPase that depletes Ca2+ stores in the endoplasmic reticulum. When TG (1 μM) is applied alone in nominally Ca2+-free Ringer’s solution, it normally produces a rapid and complete emptying of the endoplasmic Ca2+ stores, resulting in a Ca2+ transient (49). When applied after perfusion with bpV(pic) in nominally Ca2+-free Ringer, TG could not elicit a Ca2+ transient (data not shown), indicating that bpV(pic) had completely depleted the TG-sensitive Ca2+ store.

FIGURE 2.

The bpV(pic) releases Ca2+ from intracellular stores. Jurkat T cells (A) or RBL cells (B) were stimulated with 200 μM bpV(pic) (arrow) in the absence of extracellular calcium. Extracellular Ca2+ was varied between 2 mM and nominally Ca2+ free (application bar). The [Ca2+]i is shown for four representative cells (top four graphs) and the average of all cells within the field of view (bottom graph).

FIGURE 2.

The bpV(pic) releases Ca2+ from intracellular stores. Jurkat T cells (A) or RBL cells (B) were stimulated with 200 μM bpV(pic) (arrow) in the absence of extracellular calcium. Extracellular Ca2+ was varied between 2 mM and nominally Ca2+ free (application bar). The [Ca2+]i is shown for four representative cells (top four graphs) and the average of all cells within the field of view (bottom graph).

Close modal

To assess the long term effects of vanadate treatment, we monitored transcriptional activation of the Ca2+-sensitive transcription factor, NF-AT. Jurkat-NZ cells contain a reporter construct in which the β-galactosidase gene (lacZ) is driven by a triple, tandem repeat of the NF-AT binding element found in the T cell IL-2 promoter. When Jurkat-NZ cells are activated, the translocation of NF-AT to the nucleus causes a dramatic increase in β-galactosidase production that can be assessed by fluorogenic substrates (4). Stimulation of Jurkat-NZ cells with PMA and TG caused a large increase in NF-AT activity, whereas PMA or TG alone had little effect (Fig. 3). Parallel experiments with bpV(pic) yielded a smaller, but still significant, increase in NF-AT activity over that in untreated cells regardless of the presence of PMA (Fig. 3). The ability of vanadate to induce NF-AT activity in the absence of PMA contrasts with the PMA requirement for gene expression triggered by TG or ionomycin (4, 50). These results suggest that vanadate may also act to induce PKC activity.

FIGURE 3.

The bpV(pic) activates NF-AT in Jurkat T cells. The NF-AT-mediated gene expression in Jurkat-NZ cells was measured in a multiwell fluorescence plate reader using a 4-methylumbelliferyl β-d-galactopyranoside assay as described in Materials and Methods. Fluorescence measurements were made on unstimulated cells (Control) or cells stimulated for10 h by the indicated treatment, with or without 50 nM PMA. The lacZ expression is plotted in arbitrary fluorescence units (a.u.); each bar represents the mean ± SD of six samples. Similar results were obtained in four repetitions.

FIGURE 3.

The bpV(pic) activates NF-AT in Jurkat T cells. The NF-AT-mediated gene expression in Jurkat-NZ cells was measured in a multiwell fluorescence plate reader using a 4-methylumbelliferyl β-d-galactopyranoside assay as described in Materials and Methods. Fluorescence measurements were made on unstimulated cells (Control) or cells stimulated for10 h by the indicated treatment, with or without 50 nM PMA. The lacZ expression is plotted in arbitrary fluorescence units (a.u.); each bar represents the mean ± SD of six samples. Similar results were obtained in four repetitions.

Close modal

If the effect of bpV(pic) on [Ca2+]i requires IP3-evoked Ca2+ release, blockade of the IP3 receptor (IP3R) with the competitive antagonist heparin (51) should prevent the action of bpV(pic). On the other hand, if vanadate directly activates Ca2+ influx, as suggested previously (17), heparin should be without effect. To test the efficacy of heparin microinjection, we made use of the calcium response in RBL cells produced by P2 purinergic receptor stimulation (52). RBL cells were stimulated with the P2 receptor agonist ATP (10 μM) in the presence and the absence of microinjected low m.w. heparin. Fig. 4,A shows that purinergic stimulation produces an abrupt rise in [Ca2+]i and several oscillations before [Ca2+]i returns to baseline levels. As expected, heparin-injected RBL cells (Fig. 4,B) did not respond to purinergic stimulation. Stimulation by bpV(pic) was inhibited by heparin injection (Fig. 5). Both the initial Ca2+ transient and the delayed Ca2+ plateau were completely blocked. In contrast, heparin preinjection did not inhibit Ca2+ responses evoked by TG or ionomycin (data not shown), demonstrating that heparin did not interfere directly with CRAC channel function. Microinjection with control dextrans had no effect on the Ca2+ response to bpV(pic). We conclude that both transient and sustained bpV(pic)-mediated Ca2+ signals require IP3-dependent release of Ca2+.

FIGURE 4.

Heparin blocks [Ca2+]i signaling elicited by purinergic stimulation. Before Ca2+ measurements, single RBL cells were microinjected with heparin (Heparin-injected) or were left uninjected (Control). The [Ca2+]i was measured in mammalian Ringer at rest and following the application of 4 μM ATP (arrow). For each condition, [Ca2+]i is shown for four representative cells (top four graphs) and the average of all cells within the field of view (bottom graph)

FIGURE 4.

Heparin blocks [Ca2+]i signaling elicited by purinergic stimulation. Before Ca2+ measurements, single RBL cells were microinjected with heparin (Heparin-injected) or were left uninjected (Control). The [Ca2+]i was measured in mammalian Ringer at rest and following the application of 4 μM ATP (arrow). For each condition, [Ca2+]i is shown for four representative cells (top four graphs) and the average of all cells within the field of view (bottom graph)

Close modal
FIGURE 5.

Heparin blocks [Ca]i signaling induced by bpV(pic). As described in Fig. 4, [Ca2+]i was measured in control and heparin-injected RBL cells in mammalian Ringer before and after the application of 200 μM bpV(pic) (arrow). For each condition, [Ca2+]i is shown for four representative cells (top four graphs) and the average of all cells within the field of view (bottom graph).

FIGURE 5.

Heparin blocks [Ca]i signaling induced by bpV(pic). As described in Fig. 4, [Ca2+]i was measured in control and heparin-injected RBL cells in mammalian Ringer before and after the application of 200 μM bpV(pic) (arrow). For each condition, [Ca2+]i is shown for four representative cells (top four graphs) and the average of all cells within the field of view (bottom graph).

Close modal

In Jurkat T cells, depletion of intracellular Ca2+ stores results in the opening of CRAC channels and Ca2+ influx. During passive store depletion by intracellular dialysis with heavily buffered low Ca2+ solutions, an initial lag period was followed by the development of inward Ca2+ current through CRAC channels (Fig. 6). These currents exhibit inward rectification, voltage-independent gating, and a reversal potential beyond +40 mV. Inclusion of bpV(pic) in the pipette solution accelerated the activation of the CRAC currents without altering their rectification or reversal potential (Fig. 6). Using the amplitude of the current at −80 mV, we compared the effects of bpV(pic) and vanadate with that of IP3 on the activation of CRAC currents (Fig. 7). When the channels were activated by passive store depletion alone, the current reached its maximum (−1.18 ± 0.6 pA/pF) within 260 s (t1/2 = 122 ± 48 s; Fig. 7,A and Table I). Addition of IP3 (10 μM) to the intracellular solution accelerated activation of the CRAC current, reducing the t1/2 of activation to 14 ± 7 s, but did not alter the final amplitude of the current (Fig. 7,B and Table I). Dialysis of the cell with solutions containing bpV(pic) or Na3VO4 (10 μM) also accelerated the activation of CRAC channels (Fig. 7, C and D) reducing the t1/2 to 14 ± 5 and 29 ± 18 s, respectively. Thus, the enhanced rate of activation of CRAC currents by Na3VO4 and bpV(pic) mimicked the well-studied effects of IP3 on CRAC currents.

FIGURE 6.

The bpV(pic) accelerates activation of CRAC channels in Jurkat T cells. Whole-cell patch clamp recordings of Ca2+ currents through CRAC channels were made during passive store depletion alone (A) or during intracellular dialysis with 10 μM bpV(pic) (B). Voltage clamp protocol: 200-ms ramps from −120 to +40 mV were taken from a holding potential of 0 mV at intervals of 1 s. Currents during ramps taken before CRAC current induction and after maximal activation are shown. Currents from the first five ramps obtained after beginning whole-cell recording were averaged, and the average was subtracted from the remaining records to eliminate leak currents. The time after establishing the whole-cell patch clamp configuration is indicated to the left of each current-voltage plot.

FIGURE 6.

The bpV(pic) accelerates activation of CRAC channels in Jurkat T cells. Whole-cell patch clamp recordings of Ca2+ currents through CRAC channels were made during passive store depletion alone (A) or during intracellular dialysis with 10 μM bpV(pic) (B). Voltage clamp protocol: 200-ms ramps from −120 to +40 mV were taken from a holding potential of 0 mV at intervals of 1 s. Currents during ramps taken before CRAC current induction and after maximal activation are shown. Currents from the first five ramps obtained after beginning whole-cell recording were averaged, and the average was subtracted from the remaining records to eliminate leak currents. The time after establishing the whole-cell patch clamp configuration is indicated to the left of each current-voltage plot.

Close modal
FIGURE 7.

Time course of CRAC current activation. Whole-cell recordings of Ca2+ currents through CRAC channels were made in Jurkat T cells during passive store depletion (A) or during intracellular dialysis with 10 μM IP3 (B), 10 μM Na3VO4 (C), or 10 μM bpV(pic) (D). Currents were recorded as described in Fig. 6, and the CRAC current density at −80 mV is shown from time that the whole-cell patch clamp configuration was established (t = 0).

FIGURE 7.

Time course of CRAC current activation. Whole-cell recordings of Ca2+ currents through CRAC channels were made in Jurkat T cells during passive store depletion (A) or during intracellular dialysis with 10 μM IP3 (B), 10 μM Na3VO4 (C), or 10 μM bpV(pic) (D). Currents were recorded as described in Fig. 6, and the CRAC current density at −80 mV is shown from time that the whole-cell patch clamp configuration was established (t = 0).

Close modal
Table I.

Effects of vanadium compounds, thiol-reactive agents and nucleotide substitution on CRAC channel activationa

Current Density (pA/pF)Activation t1/2 (s)
Control 1.18 ± 0.6 (55) 122 ± 48 (47) 
IP3 (10 μM) 1.12 ± 0.2 (9) 14 ± 7 (9)b 
bpV(pic) (10 μM) 0.89 ± 0.2 (9) 14 ± 5 (9)b 
VO4 (10 μM) 0.81 ± 0.1 (8) 29 ± 18 (8)b 
Calyculin A (0.1 μM) 0.92 ± 0.4 (8) 121 ± 38 (8) 
Okadaic acid (0.5 μM) 0.91 ± 0.6 (12) 111 ± 34 (7) 
AMP-PNP (5 mM) 0.91 ± 0.2 (7) 126 ± 27 (8) 
ATPγS (5 mM) 0.90 ± 0.4 (4) 125 ± 31 (4) 
V04 (10 μM)+ DTT (5 mM) 1.40 ± 0.4 (9) 119 ± 55 (9) 
IP3 (10 μM)+ DTT (5 mM) 0.97 ± 0.4 (5) 16 ± 9 (5)b 
Chloramine T (500 μM) 1.32 ± 0.1 (5) 36 ± 23 (5)b 
Current Density (pA/pF)Activation t1/2 (s)
Control 1.18 ± 0.6 (55) 122 ± 48 (47) 
IP3 (10 μM) 1.12 ± 0.2 (9) 14 ± 7 (9)b 
bpV(pic) (10 μM) 0.89 ± 0.2 (9) 14 ± 5 (9)b 
VO4 (10 μM) 0.81 ± 0.1 (8) 29 ± 18 (8)b 
Calyculin A (0.1 μM) 0.92 ± 0.4 (8) 121 ± 38 (8) 
Okadaic acid (0.5 μM) 0.91 ± 0.6 (12) 111 ± 34 (7) 
AMP-PNP (5 mM) 0.91 ± 0.2 (7) 126 ± 27 (8) 
ATPγS (5 mM) 0.90 ± 0.4 (4) 125 ± 31 (4) 
V04 (10 μM)+ DTT (5 mM) 1.40 ± 0.4 (9) 119 ± 55 (9) 
IP3 (10 μM)+ DTT (5 mM) 0.97 ± 0.4 (5) 16 ± 9 (5)b 
Chloramine T (500 μM) 1.32 ± 0.1 (5) 36 ± 23 (5)b 
a

During whole-cell recording, CRAC channel currents were measured in Jurkat T cells in response to 200-ms voltage ramps from −120 mV to 40 mV, applied at 1-s intervals from a holding potential of 0 mV. The indicated reagents were added by intracellular dialysis from the pipette. The amplitude of the current was determined at −80 mV, and the time to half maximal activation (t1/2) was measured. Currents were divided by membrane capacitance to calculate the current density. Data are presented as mean ± SD with number of cells studied.

b

Value significantly different from control; p < 0 01.

The actions of Na3VO4 and bpV(pic) on CRAC channels could result from increased tyrosine phosphorylation secondary to phosphatase inhibition. We used nucleotide replacement to test whether the effects of NaVO4 could be duplicated by changes in the level of cellular phosphorylation. ATPγS is an ATP analogue that is readily used by kinases (53, 54, 55, 56). When the thiophosphoryl group from ATPγS is transferred to a protein, the resulting bond is resistant to the action of phosphatases. Therefore, nucleotide substitution of ATPγS for ATP should mimic the effects of inhibiting intracellular phosphatases. Intracellular dialysis with 5 mM ATPγS did not alter the amplitude or rate of activation of CRAC currents (Table I). Furthermore, intracellular dialysis with adenlylimidodiphosphate tetralithium salt (5 mM), a nonhydrolyzable analogue of ATP that should reduce protein phosphorylation, did not reduce the rate of CRAC channel activation. In addition, the serine phosphatase inhibitors okadaic acid and calyculin did not increase the rate of CRAC channel activation (Table I). Thus, treatments that shift the phosphorylation status of cellular proteins did not duplicate the effects of vanadate on CRAC channels.

To determine whether the effects of vanadate on CRAC currents are mediated by interaction with cysteine residues, we tested whether the action of Na3VO4 could be prevented by the thiol-reducing agent DTT. Fig. 8, A–C, shows that DTT (5 mM; in pipette and bath) prevented the activation of CRAC currents by Na3VO4 (10 μM). Comparison of rates of activation shows that DTT can prevent activation of CRAC currents by Na3VO4, but not IP3 (Table I). To test whether thiol oxidation alone could enhance the activation of CRAC channels, we included chloramine T, an agent that selectively oxidizes methionine and cysteine residues (57), in the intracellular solution. The rate of activation of CRAC channels was enhanced by chloramine T with respect to activation during passive store depletion (Fig. 8, C and D, and Table I). These data demonstrate that intracellular dialysis with vanadate or chloramine T produces an increased rate of activation of CRAC currents. Shifting the intracellular redox state to a more reducing environment eliminates the activity of both compounds, suggesting that the effects of vanadate and chloramine T are mediated by thiol oxidation. We further evaluated the hypothesis that chloramine T and vanadate act via the same mechanism. Fig. 9,A shows that chloramine T, like pbV(pic) (Fig. 5), produces a gradual rise in [Ca2+]i from baseline to micromolar levels that was inhibited by preinjection with heparin. Thus, Ca2+ signaling induced by either vanadate or chloramine T requires IP3-dependent intracellular Ca2+ release.

FIGURE 8.

Dithiothreitol blocks and chloramine T mimics vanadate-induced activation of CRAC currents. Whole-cell patch-clamp recordings of the CRAC currents in Jurkat T cells were measured at −80 mV during passive store depletion (A), during intracellular dialysis with 10 μM Na3VO4 (B), during intracellular dialysis with 10 μM Na3VO4 with 1 mM DTT in the pipette and bath solutions (C), or during intracellular dialysis with 500 μM chloramine T (D). Currents were recorded as described in Fig. 6 and are shown from time that the whole-cell patch clamp configuration was established (t = 0).

FIGURE 8.

Dithiothreitol blocks and chloramine T mimics vanadate-induced activation of CRAC currents. Whole-cell patch-clamp recordings of the CRAC currents in Jurkat T cells were measured at −80 mV during passive store depletion (A), during intracellular dialysis with 10 μM Na3VO4 (B), during intracellular dialysis with 10 μM Na3VO4 with 1 mM DTT in the pipette and bath solutions (C), or during intracellular dialysis with 500 μM chloramine T (D). Currents were recorded as described in Fig. 6 and are shown from time that the whole-cell patch clamp configuration was established (t = 0).

Close modal
FIGURE 9.

Chloramine T and bpV(pic) induce calcium signaling in RBL 2H3 cells by similar mechanisms. As shown in Fig. 5 for bpV(pic), heparin blocks [Ca2+]i signaling elicited by chloramine T. Before Ca2+ measurements, single RBL cells were microinjected with heparin (Heparin-injected) or were left uninjected (Control). The [Ca2+]i was measured in mammalian Ringer’s solutin and following the application of 50 μM chloramine T (arrow). For each condition, [Ca2+]i is shown for four representative cells (top four graphs) and the average of all cells within the field of view (bottom graph).

FIGURE 9.

Chloramine T and bpV(pic) induce calcium signaling in RBL 2H3 cells by similar mechanisms. As shown in Fig. 5 for bpV(pic), heparin blocks [Ca2+]i signaling elicited by chloramine T. Before Ca2+ measurements, single RBL cells were microinjected with heparin (Heparin-injected) or were left uninjected (Control). The [Ca2+]i was measured in mammalian Ringer’s solutin and following the application of 50 μM chloramine T (arrow). For each condition, [Ca2+]i is shown for four representative cells (top four graphs) and the average of all cells within the field of view (bottom graph).

Close modal

The data presented above demonstrate that treatment with the peroxovanadium compound, bpV(pic), induces Ca2+ release from intracellular stores and Ca2+ entry in Jurkat T lymphocytes and RBL cells. Single-cell [Ca2+]i imaging revealed oscillations as well as sustained [Ca2+]i signals when extracellular Ca2+ was present. In the absence of extracellular Ca2+, Ca2+ release transients evoked by vanadate led to complete depletion of the TG-sensitive store. Heparin, which blocks the association of IP3 with its receptor, blocked bpV(pic)-induced Ca2+ signaling. Remarkably, intracellular dialysis with bpV(pic) or Na3VO4 activated CRAC channels as rapidly as dialysis with IP3. CRAC channel activation by vanadium compounds was prevented by incubation with the thiol-reducing agent DTT and was mimicked by the thiol oxidant chloramine T, suggesting that a redox reaction contributes to their mechanism of action. Ca2+ signaling induced by bpV(pic) resulted in downstream effects on gene expression by activating the transcription factor, NF-AT, without a requirement for costimulation by addition of PMA.

Where does vanadate act to promote Ca2+ signaling? In principle, vanadate could produce Ca2+ influx directly by activating Ca2+ channels in the plasma membrane or indirectly by releasing Ca2+ from intracellular stores. The CRAC activation by vanadate was extremely rapid (Table I), suggesting that vanadate bypasses the membrane-delimited, rate-limiting steps in Ca2+ influx activation following TCR engagement. However, rapid kinetics alone do not indicate that CRAC channels are directly affected by vanadate. Our data clearly demonstrate that vanadate elicits the release of Ca2+ from intracellular stores in both RBL and Jurkat T cells (Fig. 2). Furthermore, heparin inhibited vanadate-induced Ca2+ entry in RBL cells (Fig. 5), indicating that Ca2+ entry is secondary to intracellular Ca2+ store depletion and requires the binding of IP3 to its receptor in the endoplasmic reticulum. Finally, currents activated by vanadate were indistinguishable in every characteristic from those activated by dialysis with IP3 or by passive Ca2+ store depletion (Fig. 6). Taken together, our results eliminate the possibility that vanadate produces Ca2+ influx by direct activation of CRAC channels or any other Ca2+ influx mechanism. These data are consistent with earlier studies showing pervanadate-induced Ca2+ entry secondary to store release in single RBL cells (18), but conflict with results in Jurkat T cells suggesting that pervanadate directly activates membrane Ca2+ channels (17). The latter conclusion was drawn when a rise in [Ca2+]i was not observed in the absence of extracellular Ca2+. However, that study was performed on cell suspensions using a spectrophotometer in which the average Ca2+ response from many cells was sampled simultaneously. Our results using video imaging of individual cells show that Ca2+ transients occur asynchronously following treatment with bpV(pic). Small transients in the absence of extracellular Ca2+ would be missed in population studies (see, for example, Fig. 2 B). Therefore, we conclude that vanadate induces Ca2+ influx secondarily to promoting IP3 production and Ca2+ store release, consistent with previous results demonstrating increased PLCγ activity and higher cytosolic IP3 levels in intact cells (16, 18).

How does vanadate act chemically inside the cell to promote Ca2+ signaling? Vanadate and bpV(pic) are multifunctional reagents that can interact with a number of cellular components in two primary modes of action: 1) mimicking the transition state complex formed by the phosphate ion during phosphoryl transfer reactions; and 2) directly oxidizing cysteine thiols (6). Although Na3VO4 does not shift the total cellular redox state in lymphocytes, which would appear to emphasize the importance of the first mechanism (19), measurements of the total cellular redox state principally assess the status of nonprotein thiols; thus, the selective oxidation of active site cysteines may not be reflected in those experiments. In fact, a number of studies demonstrate that thiol oxidation plays an important role in the mechanism by which vanadium compounds inhibit PTPs. First, inactivation of PTP1B by vanadium compounds is prevented, but not reversed, by the reducing agent DTT (15). Second, pervanadate directly oxidizes the active site cysteine to cysteic acid in PTP1B (15). Third, Na3VO4 and pervanadate oxidize cysteine sulfhydryls (58). Other agents that cause thiol oxidation, such as thimerosal, have been reported to enhance Ca2+ release from intracellular stores through direct interaction with the IP3R (59, 60, 61). Thimerosal was shown to increase both the single-channel conductance and the mean open time of the reconstituted Ca2+ release channel (62). In addition, thimerosal-induced Ca2+ influx has been linked to the release of Ca2+ from intracellular stores and the activation of CRAC channels in the plasma membrane of RBL-2H3 cells (63). Our data show that chloramine T mimics and DTT reverses the activation of CRAC channels by Na3VO4 and bpV(pic) (Figs. 8 and 9). Thus, our results favor oxidation as vanadate’s mode of action. The primary target must be more proximal than the IP3 receptor, because heparin blocks bpV(pic)-induced Ca2+ signaling. Furthermore, vanadate-induced gene expression did not require the presence of phorbol ester (Fig. 3), suggesting a direct action of vanadate on PLCγ to generate adequate diacylglycerol for PKC stimulation. Additional effects that might potentiate activity of the IP3R activity are not excluded.

Of the myriad of cellular actions demonstrated with vanadium compounds, their insulin-mimetic effects have garnered the most clinical interest. Micromolar concentrations of vanadate and peroxovanadium compounds stimulate hexose uptake, glucose oxidation, and lipogenesis in vivo and in vitro (26, 27, 28, 31, 64). Clinical trials demonstrating that sodium metavanadate and vanadyl sulfate improve insulin sensitivity and fasting blood glucose levels have led to suggestions for use of these agents in adjunctive therapy in diabetes (22, 23, 24, 25). The clinically applied forms of vanadate have been shown to interconvert within the cell to the +5 oxidation state and to pervandate depending upon the intracellular redox state (reviewed in Refs. 10, 11 , and 65). Thus, the effects of vanadate on Ca2+ signaling may contribute to the insulin-mimetic properties of these compounds. Unfortunately, beyond their beneficial metabolic effects, long term mitogenic and potentially tumorigenic effects of vanadium compounds must be considered (reviewed in Ref. 60). When applied to lymphocytes at clinically relevant concentrations, vanadium compounds mimic receptor-mediated activation. Our data show that bpV(pic) activates NF-AT-dependent gene expression in Jurkat T cells. Previous reports have demonstrated that peroxyvanadate compounds activate c-Jun and c-Fos and induce nuclear translocation of NF-κB in lymphocytes (17, 19). Recently, peroxovanadium compounds were shown to stimulate HIV-1 production in latently infected cell lines in conjunction with T cell activation (20). Thus, studies of the mechanisms by which vanadium compounds interact with lymphocytes may have relevance to its proposed clinical application and yield new approaches to other diseases of the immune system.

We appreciate the expert technical assistance of Dr. Luette Forrest during this project. The use of the Optical Biology Shared Resource of the Chao Family Comprehensive Cancer Center is gratefully acknowledged.

1

This work was supported by National Institutes of Health Grants NS14609 and GM41514 (to M.D.C.) and 5T32CA09054 (from the National Cancer Institute, to C.M.F.), and by fellowships from the Alexander von Humboldt Foundation (to C.E. and H.R.).

5

Abbreviations used in this paper: IP3, 1,4,5-inositol trisphosphate; bpV(pic), dipotassium oxodiperoxo (pyridine-2-carboxylato) vanadate (V); [Ca2+]i, intracellular free calcium concentration; CRAC channel, Ca2+ release-activated Ca2+ channel; IP3R, IP3 receptor; PTK, protein tyrosine kinase; PTP, protein tyrosine phosphatase; RBL cell, rat basophilic leukemia mast cell; TG, thapsigargin; fura-2-AM, fura-2-acetoxymethyl ester.

1
Lewis, R. S., M. D. Cahalan.
1995
. Potassium and calcium channels in lymphocytes.
Annu. Rev. Immunol.
13
:
623
2
Crabtree, G. R., N. A. Clipstone.
1994
. Signal transmission between the plasma membrane and nucleus of T lymphocytes.
Annu. Rev. Biochem.
63
:
1045
3
Negulescu, P. A., N. Shastri, M. D. Cahalan.
1994
. Intracellular calcium dependence of gene expression in single T lymphocytes.
Proc. Natl. Acad. Sci. USA
91
:
2873
4
Fanger, C. M., M. Hoth, G. R. Crabtree, R. S. Lewis.
1995
. Characterization of T cell mutants with defects in capacitative calcium entry: genetic evidence for the physiological roles of CRAC channels.
J. Cell Biol.
131
:
655
5
Timmerman, L. A., N. A. Clipstone, S. N. Ho, J. P. Northrop, G. R. Crabtree.
1996
. Rapid shuttling of NF-AT in discrimination of Ca2+ signals and immunosuppression.
Nature
383
:
837
6
Nechay, B. R..
1984
. Mechanisms of action of vanadium.
Annu. Rev. Pharmacol. Toxicol.
24
:
501
7
Crans, D. C., A. S. Tracey.
1998
. The chemistry of vanadium in aqueous and nonaqueous solution. A. S. Tracey, and D. C. Crans, eds.
Vanadium Compounds: Chemistry, Biochemistry, and Therapeutic Applications
2
American Chemical Society, Washington, DC.
8
Cantley, L. C. J., M. D. Resh, G. Guidotti.
1978
. Vanadate inhibits the red cell Na+,K+ ATPase from the cytoplasmic side.
Nature
272
:
552
9
Cantley, L. C. J., L. G. Cantley, L. Josephson.
1978
. A characterization of vanadate interactions with the Na,K-ATPase: mechanistic and regulatory implications.
J. Biol. Chem.
253
:
7361
10
Shaver, A., N. J. Ng, D. A. Hall, B. I. Posner.
1995
. The chemistry of peroxovanadium compounds relevant to insulin mimesis.
Mol. Cell. Biochem.
153
:
5
11
Crans, D. C., M. Mahroof-Tahir, A. D. Keramidas.
1995
. Vanadium chemistry and biochemistry of relevance for use of vanadium compounds as antidiabetic agents.
Mol. Cell. Biochem.
153
:
17
12
Morinville, A., D. Maysinger, A. Shaver.
1998
. From vanadis to atropos: vanadium compounds as pharmacological tools in cell death signalling.
Trends Pharmacol. Sci.
19
:
452
13
Tracey, A. S., M. J. Gresser.
1990
. Vandates as phosphate analogs in biochemistry. N. D. Chasteen, ed.
Vanadium in Biological Systems: Physiology and Biochemistry
63
Kluwer Academic, Dordrecht.
14
Pugazhenthi, S., F. Tanha, B. Dahl, R.L. Khandelwal.
1996
. Inhibition of a Src homology 2 domain containing protein tyrosine phosphatase by vanadate in the primary culture of hepatocytes.
Arch. Biochem. Biophys.
335
:
273
15
Huyer, G., S. Liu, J. Kelly, J. Moffat, P. Payette, B. Kennedy, G. Tsaprailis, M. J. Gresser, C. Ramachandran.
1997
. Mechanism of inhibition of protein-tyrosine phosphatases by vanadate and pervanadate.
J. Biol. Chem.
272
:
843
16
Secrist, J. P., L. A. Burns, L. Karnitz, G. A. Koretzky, R. T. Abraham.
1993
. Stimulatory effects of the protein tyrosine phosphatase inhibitor, pervanadate, on T-cell activation events.
J. Biol. Chem.
268
:
5886
17
Imbert, V., J. F. Peyron, F. Farahi, B. Mari, P. Auberger, B. Rossi.
1994
. Induction of tyrosine phosphorylation and T-cell activation by vanadate peroxide, an inhibitor of protein tyrosine phosphatases.
Biochem. J.
297
:
163
18
Teshima, R., H. Ikebuchi, M. Nakanishi, J. Sawada.
1994
. Stimulatory effect of pervanadate on calcium signals and histamine secretion of RBL-2H3 cells.
Biochem. J.
302
:
867
19
Krejsa, C. M., S. G. Nadler, J. M. Esselstyn, T. J. Kavanagh, J. A. Ledbetter, G. L. Schieven.
1997
. Role of oxidative stress in the action of vanadium phosphotyrosine phosphatase inhibitors: redox independent activation of NF-κB.
J. Biol. Chem.
272
:
11541
20
Barbeau, B., R. Bernier, N. Dumais, G. Briand, M. Olivier, R. Faure, B. I. Posner, M. Tremblay.
1997
. Activation of HIV-1 long terminal repeat transcription and virus replication via NF-κB-dependent and -independent pathways by potent phosphotyrosine phosphatase inhibitors, the peroxovanadium compounds.
J. Biol. Chem.
272
:
12968
21
Goldfine, A. B., D. C. Simonson, F. Folli, M. E. Patti, C. R. Kahn.
1995
. In vivo and in vitro studies of vanadate in human and rodent diabetes mellitus.
Mol. Cell. Biochem.
153
:
217
22
Goldfine, A. B., D. C. Simonson, F. Folli, M. E. Patti, C. R. Kahn.
1995
. Metabolic effects of sodium metavanadate in humans with insulin-dependent and noninsulin-dependent diabetes mellitus in vivo and in vitro studies.
J. Clin. Endocrinol. Metab.
80
:
3311
23
Cohen, N., M. Halberstam, P. Shlimovich, C. J. Chang, H. Shamoon, L. Rossetti.
1995
. Oral vanadyl sulfate improves hepatic and peripheral insulin sensitivity in patients with non-insulin-dependent diabetes mellitus.
J. Clin. Invest.
95
:
2501
24
Halberstam, M., N. Cohen, P. Shlimovich, L. Rossetti, H. Shamoon.
1996
. Oral vanadyl sulfate improves insulin sensitivity in NIDDM but not in obese nondiabetic subjects.
Diabetes
45
:
659
25
Boden, G., X. Chen, J. Ruiz, G. D. van Rossum, S. Turco.
1996
. Effects of vanadyl sulfate on carbohydrate and lipid metabolism in patients with non-insulin-dependent diabetes mellitus.
Metabolism
45
:
1130
26
Kadota, S., I. G. Fantus, G. Deragon, H. J. Guyda, B. Hersh, B. I. Posner.
1987
. Peroxides of vanadium: a novel and potent insulin-mimetic agent which activates the insulin receptor kinase.
Biochem. Biophys. Res. Commun.
147
:
259
27
Fantus, I. G., S. Kadota, G. Deragon, B. Foster, B. I. Posner.
1989
. Pervanadate [peroxides of vanadate] mimics insulin action in rat adipocytes via activation of the insulin receptor tyrosine kinase.
Biochemistry
28
:
8864
28
Bevan, A. P., J. W. Burgess, J. F. Yale, P. G. Drake, D. Lachance, G. Baquiran, A. Shaver, B. I. Posner.
1995
. In vivo insulin mimetic effects of pV compounds: role for tissue targeting in determining potency.
Am. J. Physiol.
268
:
E60
29
Bevan, A. P., P. G. Drake, J. F. Yale, A. Shaver, B. I. Posner.
1995
. Peroxovanadium compounds: biological actions and mechanism of insulin-mimesis.
Mol. Cell. Biochem.
153
:
49
30
Drake, P. G., B. I. Posner.
1998
. Insulin receptor-associated protein tyrosine phosphatases: role in insulin action.
Mol. Cell. Biochem.
182
:
79
31
Eriksson, J. W., P. Leonnroth, B. I. Posner, A. Shaver, C. Wesslau, U. P. Smith.
1996
. A stable peroxovanadium compound with insulin-like action in human fat cells.
Diabetologia
39
:
235
32
Etcheverry, S. B., D. C. Crans, A. D. Keramidas, A. M. Cortizo.
1997
. Insulin-mimetic action of vanadium compounds on osteoblast-like cells in culture.
Arch. Biochem. Biophys.
338
:
7
33
Imbert, V., D. Farahifar, P. Auberger, D. Mary, B. Rossi, J. F. Peyron.
1996
. Stimulation of the T-cell antigen receptor-CD3 complex signaling pathway by the tyrosine phosphatase inhibitor pervanadate is mediated by inhibition of CD45: evidence for two interconnected Lck/Fyn- or zap-70-dependent signaling pathways.
J. Inflamm.
46
:
65
34
Putney, J. J..
1986
. A model for receptor-regulated calcium entry.
Cell Calcium
7
:
1
35
Putney, J. J..
1990
. Capacitative calcium entry revisited.
Cell Calcium
11
:
611
36
Lewis, R. S., M. D. Cahalan.
1989
. Mitogen-induced oscillations of cytosolic Ca2+ and transmembrane Ca2+ current in human leukemic T cells.
Cell Regul.
1
:
99
37
Lewis, R. S., M. D. Cahalan.
1990
. Ion channels and signal transduction in lymphocytes.
Annu. Rev. Physiol.
52
:
415
38
Hoth, M., R. Penner.
1992
. Depletion of intracellular calcium stores activates a calcium current in mast cells.
Nature
355
:
353
39
Zweifach, A., R. S. Lewis.
1993
. Mitogen-regulated Ca2+ current of T lymphocytes is activated by depletion of intracellular Ca2+ stores.
Proc. Natl. Acad. Sci. USA
90
:
6295
40
Hoth, M., R. Penner.
1993
. Calcium release-activated calcium current in rat mast cells.
J. Physiol.
465
:
359
41
Premack, B. A., T. V. McDonald, P. Gardner.
1994
. Activation of Ca2+ current in Jurkat T cells following the depletion of Ca2+ stores by microsomal Ca2+-ATPase inhibitors.
J. Immunol.
152
:
5226
42
Zhang, L., M. A. McCloskey.
1995
. Immunoglobulin E receptor-activated calcium conductance in rat mast cells.
J. Physiol.
483
:
59
43
Grynkiewicz, G., M. Poenie, R.Y. Tsien.
1985
. A new generation of Ca2+ indicators with greatly improved fluorescence properties.
J. Biol. Chem.
260
:
3440
44
Fanger, C. M., S. Ghanshani, N. J. Logsdon, H. Rauer, K. Kalman, J. Zhou, K. Beckingham, K. G. Chandy, M. D. Cahalan, J. Aiyar.
1999
. Calmodulin mediates calcium-dependent activation of the intermediate conductance KCa channel, IKCa1.
J. Biol. Chem.
274
:
5754
45
Hamill, O. P., A. Marty, E. Neher, B. Sakmann, F. J. Sigworth.
1981
. Improved patch-clamp techniques for high-resolution current recording from cells and cell-free membrane patches.
Pflügers Arch. Eur. J. Physiol.
391
:
85
46
Cahalan, M. D., K. G. Chandy, T. E. DeCoursey, S. Gupta.
1985
. A voltage-gated potassium channel in human T lymphocytes.
J. Physiol.
358
:
197
47
Ehring, G. R., H. H. Kerschbaum, C. Eder, A. L. Neben, C. M. Fanger, R. M. Khoury, P. A. Negulescu, M. D. Cahalan.
1998
. A nongenomic mechanism for progesterone-mediated immunosuppression: inhibition of K+ channels, Ca2+ signaling, and gene expression in T lymphocytes.
J. Exp. Med.
188
:
1593
48
Negulescu, P. A., T. B. Krasieva, A. Khan, H. H. Kerschbaum, M. D. Cahalan.
1996
. Polarity of T cell shape, motility, and sensitivity to antigen.
Immunity
4
:
421
49
Thastrup, O., A. P. Dawson, O. Scharff, B. Foder, P. J. Cullen, B. K. Drobak, P. J. Bjerrum, S. B. Christensen, M. R. Hanley.
1989
. Thapsigargin, a novel molecular probe for studying intracellular calcium release and storage.
Agents Actions
27
:
17
50
Fiering, S., J. P. Northrop, G. P. Nolan, P. S. Mattila, G. R. Crabtree, L. A. Herzenberg.
1990
. Single cell assay of a transcription factor reveals a threshold in transcription activated by signals emanating from the T-cell antigen receptor.
Genes Dev
4
:
1823
51
Ghosh, T. K., P. S. Eis, J. M. Mullaney, C. L. Ebert, D. L. Gill.
1988
. Competitive, reversible, and potent antagonism of inositol 1,4,5-trisphosphate-activated calcium release by heparin.
J. Biol. Chem.
263
:
11075
52
Osipchuk, Y. V., M. D. Cahalan.
1992
. Cell-to-cell spread of calcium signals mediated by ATP receptors in mast cells.
Nature
359
:
241
53
Gadsby, D. C., A. C. Nairn.
1999
. Control of CFTR channel gating by phosphorylation and nucleotide hydrolysis.
Physiol. Rev.
79
:
S77
54
Kuhnae, M. R., Z. Zhao, J. Rowles, B. E. Lavan, S. H. Shen, E. H. Fischer, G. E. Lienhard.
1994
. Dephosphorylation of insulin receptor substrate 1 by the tyrosine phosphatase PTP2C.
J. Biol. Chem.
269
:
15833
55
Wagner, P. D., N. D. Vu.
1989
. Thiophosphorylation causes Ca2+-independent norepinephrine secretion from permeabilized PC12 cells.
J. Biol. Chem.
264
:
19614
56
Yount, R. G..
1975
. ATP analogs.
Adv. Enzymol.
43
:
1
57
Shechter, Y., Y. Burstein, A. Patchornik.
1975
. Selective oxidation of methionine residues in proteins.
Biochemistry
14
:
4497
58
Mikalsen, S. O., O. Kaalhus.
1998
. Properties of pervanadate and permolybdate: connexin43, phosphatase inhibition, and thiol reactivity as model systems.
J. Biol. Chem.
273
:
10036
59
Bird, G. S., G. M. Burgess, J. J. Putney.
1993
. Sulfhydryl reagents and cAMP-dependent kinase increase the sensitivity of the inositol 1,4,5-trisphosphate receptor in hepatocytes.
J. Biol. Chem.
268
:
17917
60
Pintado, E., D. Baquero-Leonis, M. Conde, F. Sobrino.
1995
. Effect of thimerosal and other sulfhydryl reagents on calcium permeability in thymus lymphocytes.
Biochem. Pharmacol.
49
:
227
61
Kaplin, A. I., C. D. Ferris, S. M. Voglmaier, S. H. Snyder.
1994
. Purified reconstituted inositol 1,4,5-trisphosphate receptors: thiol reagents act directly on receptor protein.
J. Biol. Chem.
269
:
28972
62
Thrower, E. C., H. Duclohier, E. J. Lea, G. Molle, A. P. Dawson.
1996
. The inositol 1,4,5-trisphosphate-gated Ca2+ channel: effect of the protein thiol reagent thimerosal on channel activity.
Biochem. J.
318
:
61
63
Parekh, A. B., R. Penner.
1995
. Activation of store-operated calcium influx at resting InsP3 levels by sensitization of the InsP3 receptor in rat basophilic leukaemia cells.
J. Physiol.
489
:
377
64
Leonnroth, P., J. W. Eriksson, B. I. Posner, U. Smith.
1993
. Peroxovanadate but not vanadate exerts insulin-like effects in human adipocytes.
Diabetologia
36
:
113
65
Stern, A., X. Yin, S. S. Tsang, A. Davison, J. Moon.
1993
. Vanadium as a modulator of cellular regulatory cascades and oncogene expression.
Biochem. Cell Biol.
71
:
103