首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.

Background

Sevoflurane has been demonstrated to vasodilate the foeto-placental vasculature. We aimed to determine the contribution of modulation of potassium and calcium channel function to the vasodilatory effect of sevoflurane in isolated human chorionic plate arterial rings.

Methods

Quadruplicate ex vivo human chorionic plate arterial rings were used in all studies. Series 1 and 2 examined the role of the K+ channel in sevoflurane-mediated vasodilation. Separate experiments examined whether tetraethylammonium, which blocks large conductance calcium activated K+ (KCa++) channels (Series 1A+B) or glibenclamide, which blocks the ATP sensitive K+ (KATP) channel (Series 2), modulated sevoflurane-mediated vasodilation. Series 3 – 5 examined the role of the Ca++ channel in sevoflurane induced vasodilation. Separate experiments examined whether verapamil, which blocks the sarcolemmal voltage-operated Ca++ channel (Series 3), SK&F 96365 an inhibitor of sarcolemmal voltage-independent Ca++ channels (Series 4A+B), or ryanodine an inhibitor of the sarcoplasmic reticulum Ca++ channel (Series 5A+B), modulated sevoflurane-mediated vasodilation.

Results

Sevoflurane produced dose dependent vasodilatation of chorionic plate arterial rings in all studies. Prior blockade of the KCa++ and KATP channels augmented the vasodilator effects of sevoflurane. Furthermore, exposure of rings to sevoflurane in advance of TEA occluded the effects of TEA. Taken together, these findings suggest that sevoflurane blocks K+ channels. Blockade of the voltage-operated Ca++channels inhibited the vasodilator effects of sevoflurane. In contrast, blockade of the voltage-independent and sarcoplasmic reticulum Ca++channels did not alter sevoflurane vasodilation.

Conclusion

Sevoflurane appears to block chorionic arterial KCa++ and KATP channels. Sevoflurane also blocks voltage-operated calcium channels, and exerts a net vasodilatory effect in the in vitro foeto-placental circulation.  相似文献   

2.

Background

Oxidative stress increases the cytosolic content of calcium in the cytoplasm through a combination of effects on calcium pumps, exchangers, channels and binding proteins. In this study, oxidative stress was produced by exposure to tert-butyl hydroperoxide (tBHP); cell viability was assessed using a dye reduction assay; receptor binding was characterized using [3H]N-methylscopolamine ([3H]MS); and cytosolic and luminal endoplasmic reticulum (ER) calcium concentrations ([Ca2+]i and [Ca2+]L, respectively) were measured by fluorescent imaging.

Results

Activation of M3 muscarinic receptors induced a biphasic increase in [Ca2+]i: an initial, inositol trisphosphate (IP3)-mediated release of Ca2+ from endoplasmic reticulum (ER) stores followed by a sustained phase of Ca2+ entry (i.e., store-operated calcium entry; SOCE). Under non-cytotoxic conditions, tBHP increased resting [Ca2+]i; a 90 minute exposure to tBHP (0.5-10 mM ) increased [Ca2+]i from 26 to up to 127 nM and decreased [Ca2+]L by 55%. The initial response to 10 μM carbamylcholine was depressed by tBHP in the absence, but not the presence, of extracellular calcium. SOCE, however, was depressed in both the presence and absence of extracellular calcium. Acute exposure to tBHP did not block calcium influx through open SOCE channels. Activation of SOCE following thapsigargin-induced depletion of ER calcium was depressed by tBHP exposure. In calcium-free media, tBHP depressed both SOCE and the extent of thapsigargin-induced release of Ca2+ from the ER. M3 receptor binding parameters (ligand affinity, guanine nucleotide sensitivity, allosteric modulation) were not affected by exposure to tBHP.

Conclusions

Oxidative stress induced by tBHP affected several aspects of M3 receptor signaling pathway in CHO cells, including resting [Ca2+]i, [Ca2+]L, IP3 receptor mediated release of calcium from the ER, and calcium entry through the SOCE. tBHP had little effect on M3 receptor binding or G protein coupling. Thus, oxidative stress affects multiple aspects of calcium homeostasis and calcium dependent signaling.  相似文献   

3.
Verkhratsky  A.  Solovyova  N. 《Neurophysiology》2002,34(2-3):112-117
For many years, the endoplasmic reticulum (ER) was considered to be involved in rapid signalling events due to its ability to serve as a dynamic calcium store capable of accumulating large amounts of Ca2+ ions and of releasing them in response to physiological stimulation. Recent data significantly increased the importance of the ER as a signalling organelle, by demonstrating that the ER is associated with specific pathways regulating long-lasting adaptive processes and controlling cell survival. The ER lumen is enriched by enzymatic systems involved in protein synthesis and correcting post-translational folding of these proteins. The processes of post-translational protein processing are controlled by a class of specific enzymes known as chaperones, which in turn are regulated by the free Ca2+ concentration within the ER lumen ([Ca2+]L). At the same time, a high [Ca2+]L determines the ability of the ER to generate cytosolic Ca2+ signals. Thus, the ER is able to produce signals interacting within different temporal domains. Fast ER signals result from Ca2+ release via specific Ca2+-release channels and from rapid movements of Ca2+ ions within the ER lumen (calcium tunneling). Long-lasting signals involve Ca2+-dependent regulation of chaperones with subsequent changes in protein processing and synthesis. Any malfunctions in the ER Ca2+ homeostasis result in accumulation of unfolded proteins, which in turn activates several signalling systems aimed at appropriate compensatory responses or (in the case of severe ER dysregulation) in cellular pathology and death (ER stress responses). Thus, the Ca2+ ion emerges as a messenger molecule, which integrates various signals within the ER: fluctuations of the [Ca2+]L induced by signals originating at the level of the plasmalemma (i.e., Ca2+ entry or activation of the metabotropic receptors) regulate in turn protein synthesis and processing via generating secondary signalling events between the ER and the nucleus.  相似文献   

4.
  • 1.1. Cadmium (Cd) and zinc (Zn) were inhibitory to calcium uptake by isolated gills of Fundulus heteroclitus in vitro. The metals appeared to act by displacing Ca2+ ions from protein carriers involved in facilitated diffusion.
  • 2.2. In saltwater fish, transport of calcium across the serosal membrane of gill chloride cells is partly energy dependent and is likely mediated by Ca2+-ATPase. However, much of the calcium transport through the gill epithelium appears to occur by passive processes.
  • 3.3. Cd (10−5M—10−3M) and Zn (10−7M—10−3 M) inhibited calcium uptake by isolated scale patches incubated in a physiological saline.
  • 4.4. Cyanide, oubain, and quercetin treatment of scale patches produced results similar to those of the Cd and Zn treatments suggesting that metal-induced inhibition of ATPases may be responsible for reduced calcium transport by scale osteoblasts.
  相似文献   

5.
  • 1.1. Ryanodine, an alkaloid used as an insecticide, has been shown to depress contraction while leaving excitation unaffected in mammalian hearts, an effect presumed to result from uncoupling of the transverse tubular system (TTS) from the sarcoplasmic reticulum (SR).
  • 2.2. The heart of the adult moth Hyalophora cecropia, a tissue known to have septate junctions between the TTS and SR and a Ca2+ -spike generating sarcolemma was used to further test this hypothesis.
  • 3.3. We first report the basic characteristics of the contractile response and demonstrate a negative force-frequency effect, a diminished calcium current (ICa2+) in the presence of acetylcholine and an enhanced ICa2+ with epinephrine.
  • 4.4. Ryanodine 10−8M added to this preparation slowed the inherent rhythm (interval 0.6–4 sec), depolarized the cells by 10–14 mV, reduced action-potential amplitude (from 66 to 52 mV), prolonged the plateau (from 80 to 280 msec), and decreased dV/dt from 4 to 2.8 V/sec.
  • 5.5. The magnitude of peak tension was not affected, but the time to peak tension was increased from 160 to 200 msec and the relaxation time was prolonged from 200 to 480 msec.
  • 6.6. The refractory period was increased, thereby preventing the heart from following increased rates of pacing by externally applied stimuli.
  • 7.7. We conclude that ryanodine interferes first with the sarcolemmal Ca2+-delivery system and then the SR calcium-sequestration system.
  相似文献   

6.
7.
Although the role of Ca2+ influx channels in oxidative stress signaling and cross-tolerance in plants is well established, little is known about the role of active Ca2+ efflux systems in this process. In our recent paper,17 we reported Potato Virus X (PVX)-induced acquired resistance to oxidative stress in Nicotiana benthamiana and showed the critical role of plasma membrane Ca2+/H+ exchangers in this process. The current study continues this research. Using biochemical and electrophysiological approaches, we reveal that both endomembrane P2A and P2B Ca2+-ATPases play significant roles in adaptive responses to oxidative stress by removing excessive Ca2+ from the cytosol, and that their functional expression is significantly altered in PVX-inoculated plants. These findings highlight the crucial role of Ca2+ efflux systems in acquired tolerance to oxidative stress and open up prospects for practical applications in agriculture, after in-depth comprehension of the fundamental mechanisms involved in common responses to environmental factors at the genomic, cellular and organismal levels.Key words: cytosolic calcium, reactive oxygen species, cross-tolerance, calcium pumpThe phenomenon of cross-tolerance to a variety of biotic and abiotic stresses is well-known.1,2 Some of the demonstrated examples include the correlation between oxidative stress tolerance and pathogen resistance.35 At the mechanistic level, changes in cytosolic Ca2+ levels [Ca2+]cyt, have long been implicated as a quintessential component of this process.6 The rise in [Ca2+]cyt is proven to be essential for the development of the oxidative burst required for triggering the activation of several plant defense reactions.7,8 The observed elevation in H2O2 level is believed to result from Ca2+-dependent activation of the NADPH oxidase,8 which then causes a further increase in [Ca2+]cyt via a positive feedback mechanism. This process is further accomplished by defense gene activation, phytoalexin synthesis and eventual cell death.9 Downstream from the stimulus-induced [Ca2+]cyt elevation, cells possess an array of proteins that can respond to a message. Such proteins include calmodulin (CaM),10 Ca2+-dependent protein kinases11 and CaM binding proteins.12 Of note is that when Ca2+ channels are blocked, biosynthesis of ROS is prevented.13While the role of Ca2+ influx channels in oxidative stress signaling and cross-tolerance in plants is well established, little is known about the involvement of active Ca2+ efflux systems in this process. In contrast, in animal systems the essential role of re-establishing [Ca2+]cyt to resting levels is widely reported. A sustained increase in [Ca2+]cyt in the alveolar macrophage is thought to be the consequence of membrane Ca2+-ATPase dysfunction.14 In endothelial cells, inhibition of the Ca2+/Na+ electroneutral exchanger of the mitochondria was named as one of the reasons for [Ca2+]cyt increases.15 A significant loss of the plasma membrane Ca2+-ATPase (PMCA) activity was reported in brain synapses in response to oxidative stress,16 suggesting that PMCA may be a downstream target of oxidative stress.In our recently published paper17 we reported the phenomenon of Potato Virus X (PVX)-induced acquired resistance to oxidative stress in Nicotiana benthamiana plants and showed the critical role of plasma membrane Ca2+/H+ exchangers in this process. Nonetheless, questions remain, is this transporter the only active Ca2+ efflux system involved in this process?In addition to Ca2+/H+ exchangers, active Ca2+ extrusion could also be achieved by Ca2+-ATPases. Two major types of Ca2+-ATPases that differ substantially in their pharmacology and sensitivity to CaM are known.18 Type P2A pumps (also called ER-type or ECA19,20) are predominantly ER-localized,19 although they are also present at other endomembranes (e.g., tonoplast and Golgi). Four members of this group have been identified in the Arabidopsis genome (named AtECAs 1 to 4).18,21 These pumps lack an N-terminal autoregulatory domain, are insensitive to CaM and suppressed by cyclopropiazonic acid (CPA).19 P2B (or ACA) pumps contain an autoinhibitory N-terminal domain that possesses a binding site for Ca2+-CaM.18 Ten members are known in Arabidopsis (termed AtACA1, 2, 4 and 7 to 13).21 Plant P2B pumps are located at the plasma membrane20 as well as in inner membranes such as tonoplast (e.g., ACA4), ER (e.g., ACA2) and plastids.18,19 These pumps probably constitute the basis for precise cytosolic Ca2+ regulation; as the Ca2+ concentration increases, CaM is activated and binds to the autoinhibitory domain of the Ca2+ pump. This results in the activation of the pump.In our recent study,17 we found no significant difference between the purified plasma membranes fractions isolated from control and UV-treated tobacco plants (with or without PVX inoculation) either in the Ca2+-ATPase activity or in the Ca2+-ATPase expression level and its ability to bind CaM. This suggests that the plasma membrane P2B type pumps (the only pump type known to be expressed at the plasma membrane) play no major role in removing excess Ca2+ from the cytosol under oxidative stress conditions. This led to an obvious question: what about endomembrane Ca2+-ATPases?To address this issue, microsomal membrane fractions were isolated from tobacco leaves in a manner previously described for plasma membrane fractions17 (Fig. 1A). Western blot and CaM overlay assays were then made to investigate the role of endomembrane P2B Ca2+-ATPases in our reported phenomena of acquired resistance. The results show that the expression of the P2B Ca2+ pumps in PVX-inoculated plants is significantly higher than in control plants (Fig. 1B), correlating well with the CaM overlay assay (Fig. 1C). As no difference was observed for the P2B Ca2+-ATPase expression levels in the plasma membranes,17 the observed difference in the microsomal fractions of PVX-infected plants must be due to an increased expression of endomembrane P2B Ca2+-ATPases. Given the fact that Ca2+ pumps have a high affinity for calcium, the observed increase in endomembrane P2B-type Ca2+-ATPases expression in PVX-inoculated plants may be advantageous for more efficient Ca2+ removal from the cytosol into internal organelles.Open in a separate windowFigure 1Expression of P2B Ca2+ in purified microsomal fractions from tobacco leaves. Measurements were undertaken C = mock controls; C-UV = mock controls treated with UV-light; PVX = PVX infected plants; PVX-UV = PVX inoculated plants treated with UV-light. (A) Coomassie Brilliant Blue-stained gel; (B) Protein blot immunostained with a non isoform-specific polyclonal antibody for P2B Ca2+-ATPases; (C) CaM overlay assay.To decipher the possible role of P2A Ca2+-ATPases in acquired resistance, a series of electrophysiological experiments were conducted using inhibitors of P2A-type Ca2+-ATPases, such as thapsigargin (TG)22 and cyclopiazonic acid (CPA).23 Ion-selective Ca2+ microelectrodes were prepared as described elsewhere in reference 24 and 25, and net Ca2+ fluxes were measured from tobacco mesophyll tissue following previously described protocols.17 Leaf pre-treatment for 2 h in either of these inhibitors dramatically suppressed the net Ca2+ efflux measured from tobacco mesophyll cells 2 h after UV light exposure (Fig. 2). Given the specificity of TG and CPA inhibitors for P2A-type Ca2+-ATPases, these results strongly support a hypothesis that both endomembrane P2A and P2B Ca2+-ATPases play significant roles in plant adaptive responses to oxidative stress. This is achieved by removing excess Ca2+ from the cytosol.Open in a separate windowFigure 2Effect of known Ca2+-ATPase blockers on light-induced Ca2+ flux kinetics after 20 min of UV-C treatment. Leaf mesophyll segments were pre-treated in either 5 µM TG (thapsigargin) or 50 µM CPA (cyclopiazonic acid) for 1–1.5 h prior to exposure to UV-C light. Net Ca2+ fluxes were measured 2 h after the end of UV treatment. These were compared with two controls: (1) no pre-treatment/no UV exposure (closed circles) and (2) no pre-treatment/20 min UV exposure (open squares). Mean ± SE (n = 4 to 7).Combining these results with our previously reported observations in reference 17, the following model is proposed (Fig. 3). Oxidative stress (such as UV) causes increased ROS production in leaf chloroplasts, leading to the elevated [Ca2+]cyt. Several Ca2+ efflux systems are involved in restoring basal cytosolic Ca2+ levels. Two of these, the plasma membrane Ca2+/H+ exchanger17 and endomembrane P2A and P2B Ca2+-ATPases (as reported in this study) are upregulated in PVX inoculated plants and contribute to the improved tolerance to oxidative stress. Overall, these findings highlight the potential role of Ca2+ efflux systems in virus-induced tolerance to oxidative stress in plants. This is consistent with our previous reports on the important role of Ca2+ efflux systems in biotic stress tolerance26 and brings forth possibilities for genetic engineering of more tolerant plants by targeting expression and regulation of active Ca2+ efflux systems at either the plasma or endomembranes.Open in a separate windowFigure 3The proposed model of oxidative stress signaling and the role of Ca2+-efflux systems in acquired resistance and plant adaptation to oxidative stress.Overall, a better adaptation of virus-infected plants to a short wave UV irradiation as compared to uninfected controls may suggest that infection triggers common defense mechanisms that could be efficient against secondary unrelated stresses. This observation may lead to the development of novel strategies to protect plants against complex environmental stress conditions.  相似文献   

8.
9.
Calcium entry through voltage-gated calcium channels has widespread cellular effects upon a host of physiological processes including neuronal excitability, muscle excitation-contraction coupling, and secretion. Using single particle analysis methods, we have determined the first three-dimensional structure, at 23 Å resolution, for a member of the low voltage-activated voltage-gated calcium channel family, CaV3.1, a T-type channel. CaV3.1 has dimensions of ∼115 × 85 × 95 Å, composed of two distinct segments. The cytoplasmic densities form a vestibule below the transmembrane domain with the C terminus, unambiguously identified by the presence of a His tag being ∼65 Å long and curling around the base of the structure. The cytoplasmic assembly has a large exposed surface area that may serve as a signaling hub with the C terminus acting as a “fishing rod” to bind regulatory proteins. We have also determined a three-dimensional structure, at a resolution of 25 Å, for the monomeric form of the cardiac L-type voltage-gated calcium (high voltage-activated) channel with accessory proteins β and α2δ bound to the ion channel polypeptide CaV1.2. Comparison with the skeletal muscle isoform finds a good match particularly with respect to the conformation, size, and shape of the domain identified as that formed by α2. Furthermore, modeling of the CaV3.1 structure (analogous to CaV1.2 at these resolutions) into the heteromeric L-type voltage-gated calcium channel complex volume reveals multiple interaction sites for β-CaV1.2 binding and for the first time identifies the size and organization of the α2δ polypeptides.To date, five different types of voltage-gated calcium channels (VGCCs)4 have been identified, L, N, P/Q, R, and T, and classified according to their physiological and pharmacological characteristics (13). On the basis of their electrophysiological properties, VGCCs can be divided into two classes: high voltage-activated (HVA) and low voltage-activated (LVA). T-type Ca2+ channels form the LVA family and are characterized by their low threshold of activation, small single channel conductance, slow deactivation, and a low sensitivity to classical blockers of HVA channels (46). T-type channels have a central role regulating, for example, cardiac pacemaking of sinoatrial node cells and tonic firing patterns in neurons (5). Three T-type channel isoforms have been identified and cloned: CaV3.1, CaV3.2, and CaV3.3, with each isoform possessing several splice variants showing distinct functional properties (reviewed in Ref. 7).Each VGCC is composed of a pore-forming polypeptide termed the CaV α1-subunit, with 10 mammalian α1 isoforms identified, divided into three subfamilies: CaV1–3 (8). Housed within the CaV α1 subunit are the calcium pore, voltage-sensing apparatus, drug binding sites, and numerous structural determinants required for binding auxiliary subunits and other regulatory proteins. Analysis of the amino acid sequences and predicted secondary structure of the T-type channels suggests a similar topology to the HVA CaV α1 subunits and K+ and Na+ channels, implying that they are evolutionarily related (5).HVA channels are heteromeric complexes formed by the CaV α1 polypeptide, with several accessory subunits non-covalently bound. For example, the cardiac L-type voltage-gated calcium channel is formed by the CaV1.2 subunit in association with the soluble β-polypeptide localized to the intracellular side of the plasma membrane and a largely extracellular α2δ subunit (9, 10). The β-subunit has a role in regulating activation, inactivation, and voltage dependence as well as targeting of CaV1.2 to the plasma membrane. The crystal structure of the core region of the β-polypeptide in complex with a peptide corresponding to the interacting region of the CaV1.2 (AID) has been described (11, 12), revealing that it is comprised of two domains, a type 3 Src homology (SH3) domain and a guanylate kinase-like domain. CaV1.2 is associated, on the extracellular side of the membrane, with the α2δ subunit, the product of post-translational cleavage of a single gene comprised of a glycosylated extracellular α2 domain linked by disulfide bonds to the transmembrane δ polypeptide, which is also mainly extracellular and glycosylated. Four isoforms of α2δ have been identified (13, 14). The role of the α2δ protein is not as well understood as that of the β-subunit, but it has been shown to increase the current amplitude and have effects on inactivation (15). An additional membrane-spanning auxiliary subunit, γ, was initially thought to be unique to the skeletal muscle LTCC; however, recent studies have identified neuronal isoforms, although it remains unclear whether they have any role as calcium channel subunits (16, 17).We report here the purification of a recombinant CaV3.1, leading to the calculation of the first three-dimensional structure for a member of the LVA channel family. CaV3.1 is formed by two distinct segments, which we have been able to assign to the transmembrane and cytoplasmic domains. We have identified the C-terminal domain that forms a tail that winds around the base of the structure, providing insights as to how this channel may be regulated through the binding of accessory/regulatory proteins and/or long range conformational movements. Furthermore, we have been able to utilize this new three-dimensional structure to probe the assembly of the polypeptides forming the cardiac LTCCs after having also calculated a novel three-dimensional structure for the monomeric form of the channel purified from bovine heart. At the resolutions described here, CaV3.1 can be considered analogous to CaV1.2; see Fig. 1A for a sequence alignment overview. No mandatory auxiliary subunits for the T-type, LVA, channels have been identified. However, studies have shown that co-expression of CaV3.1 with α2δ subunits led to a 2-fold increase in T-type-mediated currents (18), and thus this model may also provide an insight as to how accessory proteins may associate with CaV3.1 to exert regulatory effects.Open in a separate windowFIGURE 1.Characterization, purification and image analysis of the T-type voltage gated calcium channel Cav3.1. A, schematic overview of a sequence comparison of voltage-gated calcium Ca2+ channel subunits CaV1.2 and CaV3.1. Sequence alignment was carried out using ClustalW (37). Solid regions indicate aligned sequences (black blocks correspond to the transmembrane helices); extracellular loops comprising <10 amino acids are not depicted. B, lane 1, silver-stained 10% SDS-PAGE of purified recombinant CaV3.1 revealing a single polypeptide band at ∼250 kDa. Lane 2, identification of the purified Cav3.1 by Western blotting (anti-CaV3.1, Santa Cruz Biotechnology sc-25690). Lane 3, Western blot of the purified Cav3.1 using an anti-His (Santa Cruz Biotechnology) antibody revealing a single protein product (recombinant Cav3.1 with a C-terminal His tag) at ∼250 kDa. C, field of negatively stained (2% w/v uranyl acetate) recombinant CaV3.1 showing white protein particles presenting a range of orientations ∼85–115 Å in size. The asterisk indicates a small area of aggregation that is easily distinguishable from the single CaV3.1 complexes. The black arrows indicate square-shaped particles with a side length of ∼100 Å. D, column I, examples of reference-free class averages obtained by alignment of the raw data that are representative of the range of multiple orientations sampled (box size 230 × 230 Å). Column II, corresponding back projections of the final three-dimensional volume illustrate that the structural features are consistent with those shown in the class averages. E, Fourier shell correlation plot indicating at a correlation of 0.5 that the three-dimensional CaV3.1 structure is at a resolution of 23 Å.  相似文献   

10.
As a stable analog for ADP-sensitive phosphorylated intermediate of sarcoplasmic reticulum Ca2+-ATPase E1PCa2·Mg, a complex of E1Ca2·BeFx, was successfully developed by addition of beryllium fluoride and Mg2+ to the Ca2+-bound state, E1Ca2. In E1Ca2·BeFx, most probably E1Ca2·BeF3, two Ca2+ are occluded at high affinity transport sites, its formation required Mg2+ binding at the catalytic site, and ADP decomposed it to E1Ca2, as in E1PCa2·Mg. Organization of cytoplasmic domains in E1Ca2·BeFx was revealed to be intermediate between those in E1Ca2·AlF4 ADP (transition state of E1PCa2 formation) and E2·BeF3·(ADP-insensitive phosphorylated intermediate E2P·Mg). Trinitrophenyl-AMP (TNP-AMP) formed a very fluorescent (superfluorescent) complex with E1Ca2·BeFx in contrast to no superfluorescence of TNP-AMP bound to E1Ca2·AlFx. E1Ca2·BeFx with bound TNP-AMP slowly decayed to E1Ca2, being distinct from the superfluorescent complex of TNP-AMP with E2·BeF3, which was stable. Tryptophan fluorescence revealed that the transmembrane structure of E1Ca2·BeFx mimics E1PCa2·Mg, and between those of E1Ca2·AlF4·ADP and E2·BeF3. E1Ca2·BeFx at low 50–100 μm Ca2+ was converted slowly to E2·BeF3 releasing Ca2+, mimicking E1PCa2·Mg → E2P·Mg + 2Ca2+. Ca2+ replacement of Mg2+ at the catalytic site at approximately millimolar high Ca2+ decomposed E1Ca2·BeFx to E1Ca2. Notably, E1Ca2·BeFx was perfectly stabilized for at least 12 days by 0.7 mm lumenal Ca2+ with 15 mm Mg2+. Also, stable E1Ca2·BeFx was produced from E2·BeF3 at 0.7 mm lumenal Ca2+ by binding two Ca2+ to lumenally oriented low affinity transport sites, as mimicking the reverse conversion E2P· Mg + 2Ca2+E1PCa2·Mg.Sarcoplasmic reticulum Ca2+-ATPase (SERCA1a),2 a representative member of the P-type ion transporting ATPases, catalyze Ca2+ transport coupled with ATP hydrolysis (Fig. 1) (19). The enzyme forms phosphorylated intermediates from ATP or Pi in the presence of Mg2+ (1013). In the transport cycle, the enzyme is first activated by cooperative binding of two Ca2+ ions at high affinity transport sites (E2 to E1Ca2, steps 1–2) (14) and autophosphorylated at Asp351 with MgATP to form the ADP-sensitive phosphoenzyme (E1P, step 3), which reacts with ADP to regenerate ATP in the reverse reaction. Upon this E1P formation, the two bound Ca2+ are occluded in the transport sites (E1PCa2). Subsequent isomeric transition to the ADP-insensitive form (E2PCa2), i.e. loss of ADP sensitivity at the catalytic site, results in rearrangement of the Ca2+ binding sites to deocclude Ca2+, reduce the affinity, and open the lumenal gate, thus releasing Ca2+ into the lumen (E2P, steps 4–5). Finally Asp351-acylphosphate in E2P is hydrolyzed to form the Ca2+-unbound inactive E2 state (steps 6 and 7). Mg2+ bound at the catalytic site is required as a physiological catalytic cofactor in phosphorylation and dephosphorylation and thus for the transport cycle. The cycle is totally reversible, e.g. E2P can be formed from Pi in the presence of Mg2+ and absence of Ca2+, and subsequent Ca2+ binding at lumenally oriented low affinity transport sites of E2P reverses the Ca2+-releasing step and produces E1PCa2, which is then decomposed to E1Ca2 by ADP.Open in a separate windowFIGURE 1.Ca2+ transport cycle of Ca2+-ATPase.Various intermediate structural states in the transport cycle were fixed as their structural analogs produced by appropriate ligands such as AMP-PCP (non-hydrolyzable ATP analog) or metal fluoride compounds (phosphate analogs), and their crystal structures were solved at the atomic level (1522). The three cytoplasmic domains, N, P, and A, largely move and change their organization state during the transport cycle, and the changes are coupled with changes in the transport sites. Most remarkably, in the change from E1Ca2·AlF4·ADP (the transition state for E1PCa2 formation, E1PCa2·ADP·Mg) to E2·BeF3 (the ground state E2P·Mg) (2325), the A domain largely rotates by more than 90° approximately parallel to the membrane plane and associates with the P domain, thereby destroying the Ca2+ binding sites, and opening the lumenal gate, thus releasing Ca2+ into the lumen (see Fig. 2). E1PCa2·Ca·AMP-PN formed by CaAMP-PNP without Mg2+ is nearly the same as E1Ca2·AlF4·ADP and E1Ca2·CaAMP-PCP in their crystal structures (17, 18, 22).Open in a separate windowFIGURE 2.Structure of SERCA1a and its change during processing of phosphorylated intermediate. E1Ca2·AlF4·ADP (the transition state analog for phosphorylation E1PCa2·ADP·Mg) and E2·BeF3 (the ground state E2P analog (25)) were obtained from the Protein Data Bank (PDB accession code 1T5T (17) and 2ZBE (21), respectively). Cytoplasmic domains N (nucleotide binding), P (phosphorylation), and A (actuator), and 10 transmembrane helices (M1–M10) are indicated. The arrows on the domains, M1′ and M2 (Tyr122) in E1Ca2·AlF4·ADP, indicate their approximate motions predicted for E1PCa2·ADP·MgE2P·Mg. The phosphorylation site Asp351, TGES184 of the A domain, Arg198 (tryptic T2 site) on the Val200 loop (DPR198AV200NQD) of the A domain, and Thr242 (proteinase K site) on the A/M3-linker are shown. Seven hydrophobic residues gather in the E2P state to form the Tyr122-hydrophobic cluster (Y122-HC); Tyr122/Leu119 on the top part of M2, Ile179/Leu180/Ile232 of the A domain, and Val705/Val726 of the P domain. The overall structure of E1Ca2·AlF4·ADP is virtually the same as those of E1Ca2·CaAMP-PCP and E1PCa2·Ca·AMP-PN (17, 18, 22).Despite these atomic structures, yet unsolved is the structure of E1PCa2·Mg, the genuine physiological intermediate E1PCa2 with bound Mg2+ at the catalytic site without the nucleotide. Its stable structural analog has yet to be developed. E1PCa2·Mg is the major intermediate accumulating almost exclusively at steady state under physiological conditions. Its rate-limiting isomerization results in Ca2+ deocclusion/release producing E2P·Mg as a key event for Ca2+ transport. In E1Ca2·CaAMP-PCP, E1Ca2·AlF4·ADP, and E1PCa2·Ca·AMP-PN, the N and P domains are cross-linked and strongly stabilized by the bound nucleotide and/or Ca2+ at the catalytic site, thus they are crystallized (17, 18, 22). Kinetically, E1PCa2·Ca formed with CaATP is markedly stabilized due to Ca2+ binding at the catalytic Mg2+ site, and its isomerization to E2P is strongly retarded in contrast to E1PCa2·Mg (26, 27). Thus, the bound Ca2+ at the catalytic Mg2+ site likely produces a significantly different structural state from that with bound Mg2+.Therefore, it is now essential to develop a genuine E1PCa2·Mg analog without bound nucleotide and thereby gain further insight into the structural mechanism in the Ca2+ transport process. It is also crucial to further clarify the structural importance of Mg2+ as the physiological catalytic cation. In this study, we successfully developed the complex E1Ca2·BeFx, most probably E1Ca2·BeF3, as the E1PCa2·Mg analog by adding beryllium fluoride (BeFx) to the E1Ca2 state without any nucleotides. For its formation, Mg2+ binding at the catalytic site was required and Ca2+ substitution for Mg2+ was absolutely unfavorable, revealing a likely structural reason for its preference as the physiological cofactor. In E1Ca2·BeF3, two Ca2+ ions bound at the high affinity transport sites are occluded. It was also produced from E2·BeF3 by lumenal Ca2+ binding at the lumenally oriented low affinity transport sites, mimicking E2P·Mg + 2Ca2+E1PCa2·Mg. All properties of the newly developed E1Ca2·BeF3 fulfilled the requirements as the E1PCa2·Mg analog, and hence we were able to uncover the hitherto unknown nature of E1PCa2·Mg as well as structural events occurring in the phosphorylation and isomerization processes. Also, we successfully found the conditions that perfectly stabilize the E1Ca2·BeF3 complex.  相似文献   

11.
  • 1.1. Blood Na+ and Cl−1 levels in Crangon crangon and Carcinus maenas were not significantly affected during Cu/Zn (0.2 5mg·l−1) or hypoxic (pwO2 = 40 torr) exposure at both 13.5 and 27.0%. However decreases in blood ion levels were evident in heavy metal/hypoxia combinations in low salinity media.
  • 2.2. In Carcinus blood Ca2+ regulation was not affected by heavy metal or hypoxic exposure, however, combinations resulted in salinity-dependent increases in blood Ca2+ levels.
  相似文献   

12.
Endothelial dysfunction causes an imbalance in endothelial NO and O2 production rates and increased peroxynitrite formation. Peroxynitrite and its decomposition products cause multiple deleterious effects including tyrosine nitration of proteins, superoxide dismutase (SOD) inactivation, and tissue damage. Studies have shown that peroxynitrite formation during endothelial dysfunction is strongly dependent on the NO and O2 production rates. Previous experimental and modeling studies examining the role of NO and O2 production imbalance on peroxynitrite formation showed different results in biological and synthetic systems. However, there is a lack of quantitative information about the formation and biological relevance of peroxynitrite under oxidative, nitroxidative, and nitrosative stress conditions in the microcirculation. We developed a computational biotransport model to examine the role of endothelial NO and O2 production on the complex biochemical NO and O2 interactions in the microcirculation. We also modeled the effect of variability in SOD expression and activity during oxidative stress. The results showed that peroxynitrite concentration increased with increase in either O2 to NO or NO to O2 production rate ratio (QO2/QNO or QNO/QO2, respectively). The peroxynitrite concentrations were similar for both production rate ratios, indicating that peroxynitrite-related nitroxidative and nitrosative stresses may be similar in endothelial dysfunction or inducible NO synthase (iNOS)-induced NO production. The endothelial peroxynitrite concentration increased with increase in both QO2/QNO and QNO/QO2 ratios at SOD concentrations of 0.1–100 μM. The absence of SOD may not mitigate the extent of peroxynitrite-mediated toxicity, as we predicted an insignificant increase in peroxynitrite levels beyond QO2/QNO and QNO/QO2 ratios of 1. The results support the experimental observations of biological systems and show that peroxynitrite formation increases with increase in either NO or O2 production, and excess NO production from iNOS or from NO donors during oxidative stress conditions does not reduce the extent of peroxynitrite mediated toxicity.  相似文献   

13.
Neutrophil-like HL-60 cells reacted to N -formyl- -Methionyl- -Leucyl- -P henylalanine (f MLP) with a rise in the intracellular calcium concentration ([Ca2]i), NADPH oxidase activation, and increased superoxide anion (O2) production. [Ca2+]imobilization and superoxide production were largely dependent on extracellular calcium (Ca2+]e) and a capacitative calcium entry. The monomeric G-protein, Rac-1, regulates NADPH oxidase activity. We tested the effect of removal of Ca2+]eon Rac-1 plasma membrane sequestration and activation of NADPH oxidase using immunodetection and a double labelling fluorescent method. Results showed that Rac-1 activation is mediated via a pertussis toxin (PTX)-sensitive heteromeric G-protein pathway, and that Rac-1 membrane sequestration was preceded by [Ca2+]imobilization following entry of Ca2+e. Therefore, we propose that O2production is dependent on activation of PTX-sensitive G-proteins and sequestration of Rac-1 in the plasma membrane, following entry of Ca2+e.  相似文献   

14.
  • 1.1. The intracellular concentration of ionized calcium was measured with double-barrelled ion-sensitive microelectrodes under short-circuit conditions in the isolated outer mantle epithelium of Anodonta cygnea.
  • 2.2. When the outside baths contained 1 mmol/1 Ca2+ the average intracellular Ca2+ was 5.42 ± 0.64 mmol/1(N = 41) while the equilibrium concentration estimated from the intracellular potential measured in the same cells was 5.51 ± 0.33 mmol/l.
  • 3.3. Bilateral removal of calcium from the external baths induced a fast fall in the intracellular concentration of this ion by almost three orders of magnitude. This effect was similar to that obtained by removing calcium from the bath on the basolateral side.
  • 4.4. Removal of calcium from the bath in contact with the apical side of the preparation had little effect on intracellular calcium.
  相似文献   

15.

Background/Aims

Resveratrol has been demonstrated to be protective in the cardiovascular system. The aim of this study was to assess the effects of resveratrol on hydrogen peroxide (H2O2)-induced increase in late sodium current (I Na.L) which augmented the reverse Na+-Ca2+ exchanger current (I NCX), and the diastolic intracellular Ca2+ concentration in ventricular myocytes.

Methods

I Na.L, I NCX, L-type Ca2+ current (I Ca.L) and intracellular Ca2+ properties were determined using whole-cell patch-clamp techniques and dual-excitation fluorescence photomultiplier system (IonOptix), respectively, in rabbit ventricular myocytes.

Results

Resveratrol (10, 20, 40 and 80 µM) decreased I Na.L in myocytes both in the absence and presence of H2O2 (300 µM) in a concentration dependent manner. Ranolazine (3–9 µM) and tetrodotoxin (TTX, 4 µM), I Na.L inhibitors, decreased I Na.L in cardiomyocytes in the presence of 300 µM H2O2. H2O2 (300 µM) increased the reverse I NCX and this increase was significantly attenuated by either 20 µM resveratrol or 4 µM ranolazine or 4 µM TTX. In addition, 10 µM resveratrol and 2 µM TTX significantly depressed the increase by 150 µM H2O2 of the diastolic intracellular Ca2+ fura-2 fluorescence intensity (FFI), fura-fluorescence intensity change (△FFI), maximal velocity of intracellular Ca2+ transient rise and decay. As expected, 2 µM TTX had no effect on I Ca.L.

Conclusion

Resveratrol protects the cardiomyocytes by inhibiting the H2O2-induced augmentation of I Na.L.and may contribute to the reduction of ischemia-induced lethal arrhythmias.  相似文献   

16.
The acetoxy-functionalized bis(imidazolyl)borate [B(ImN-Me)2(OC(O)Me)Me] (=LOAc) is synthesized by the reaction of the alkoxy precursor [B(ImN-Me)2(OPri)Me] (=LOiPr) with acetic acid. In the presence of weak Brønstead acid, migration of nickel-bound acetate anion to the boron center giving LOAc occurs. The boron-acetoxy linkage survives upon the treatment of the nickel complexes with OH, although the acetoxy group on LOAc does not coordinate to the nickel center.  相似文献   

17.
Manganese(II) complexes, Mn2L13(ClO4)4, MnL1(H2O)2(ClO4)2, MnL2(H2O)2(ClO4)2, and {(μ-Cl)MnL2(PF6)}2 based on N,N′-bis(2-pyridinylmethylene) ethanediamine (L1) and N,N′-bis(2-pyridinylmethylene) propanediamine (L2) ligands have been prepared and characterized. The single crystal X-ray diffraction analysis of Mn2L23(ClO4)4 shows that each of the two Mn(II) ion centers with a Mn-Mn distance of 7.15 Å are coordinated by one ligand while a common third ligand bridges the metal centers. Solid-state magnetic susceptibility measurements as well as DFT calculations confirm that each of the manganese centers is high-spin S = 5/2. The electronic structure obtained shows no orbital overlap between the Mn(II) centers indicating that the observed weak antiferromagentism is a result of through space interactions between the two Mn(II) centers. Under different reaction conditions, L1 and Mn(II) yielded a one-dimensional polymer, MnL1(H2O)2(ClO4)2. Ligand L2 when reacted with manganese(II) perchlorate gives contrarily to L1 mononuclear MnL2(H2O)2(ClO4)2 complex. The analysis of the structural properties of the MnL2(H2O)2(ClO4)2 lead to the design of dinuclear complex {(μ-Cl)MnL2(PF6)} where two chlorine atoms were utilized as bridging moieties. This complex has a rhomboidal Mn2Cl2 core with a Mn-Mn distance of 3.726 Å. At room temperature {(μ-Cl)MnL2(PF6)} is ferromagnetic with observed μeff = 4.04 μB per Mn(II) ion. With cooling, μeff grows reaching 4.81 μB per Mn(II) ion at 8 K, and then undergoes ferromagnetic-to-antiferromagnetic phase transition.  相似文献   

18.
Activation of the human red cell calcium ATPase by calcium pretreatment   总被引:1,自引:0,他引:1  
Some kinetic parameters of the human red cell Ca2+-ATPase were studied on calmodulin-free membrane fragments following preincubation at 37°C. After 30 min treatment with EGTA(1 mm) plus dithioerythritol (1 mm), a V max of about 0.4 μmol Pi/mg × hr and a K s of 0.3 μm Ca2+ were found. When Mg2+ (10 mm) or Ca2+(10 μm) were also added during preincubation, V maxbut not Kwas altered. Ca2+ was more effective than Mg2+, thus increasing V max to about 1.3 μmol Pi/mg × hr. The presence of both Ca2+ and Mg2+ during pretreatment decreasedKto 0.15 μm, while having no apparent effect on V max. Conversely, addition of ATP (2 mm) with either Ca2+ or Ca2+ plus Mg2+increased Vmax without affecting K. Preincubation with Ca2+ for periods longer than 30 min further increased Vmaxand reduced Kto levels as low as found with calmodulin treatment. The Ca2+ activation was not prevented by adding proteinase inhibitors (iodoacetamide, 10 mm; leupeptin, 200 μm; pepstatinA, 100 μm; phenylmethanesulfonyl fluoride, 100 μm). The electrophoretic pattern of membranes preincubated with or without Mg2+, Ca2+ or Ca2+ plus Mg2+ did not differ significantly from each other. Moreover, immunodetection of Ca2+-ATPase by means of polyclonal antibodiesrevealed no mobility change after the various treatments. The above stimulation was not altered by neomycin (200 μm), washing with EGTA (5 mm) or by both incubating and washing with delipidized serum albumin (1 mg/ml), or omitting dithioerythritol from the preincubation medium. On the other hand, the activation elicited by Ca2+ plus ATP in the presence of Mg2+ was reduced 25–30% by acridine orange (100 μm), compound 48/80 (100 μm) or leupeptin (200 μm) but not by dithio-bis-nitrobenzoic acid (1 mm). The fluorescence depolarization of 1,6-diphenyl-and l-(4-trimethylammonium phenyl)-6-phenyl 1,3,5-hexatriene incorporated into membrane fragments was not affected after preincubating under the different conditions. The results show that proteolysis, fatty acid production, an increased phospholipid metabolism or alteration of membrane fluidity are not involved in the Ca2+ effect. Ca2+ preincubation may stimulate the Ca2+-ATPase activity by stabilizing or promoting the E1 conformation.  相似文献   

19.
Ion translocation by the sarcoplasmic reticulum Ca2+-ATPase depends on large movements of the A-domain, but the driving forces have yet to be defined. The A-domain is connected to the ion-binding membranous part of the protein through linker regions. We have determined the functional consequences of changing the length of the linker between the A-domain and transmembrane helix M3 (“A-M3 linker”) by insertion and deletion mutagenesis at two sites. It was feasible to insert as many as 41 residues (polyglycine and glycine-proline loops) in the flexible region of the linker without loss of the ability to react with Ca2+ and ATP and to form the phosphorylated Ca2E1P intermediate, but the rate of the energy-transducing conformational transition to E2P was reduced by >80%. Insertion of a smaller number of residues gave effects gradually increasing with the length of the insertion. Deletion of two residues at the same site, but not replacement with glycine, gave a similar reduction as the longest insertion. Insertion of one or three residues in another part of the A-M3 linker that forms an α-helix (“A3 helix”) in E2/E2P conformations had even more profound effects on the ability of the enzyme to form E2P. These results demonstrate the importance of the length of the A-M3 linker and of the position and integrity of the A3 helix for stabilization of E2P and suggest that, during the normal enzyme cycle, strain of the A-M3 linker could contribute to destabilize the Ca2E1P state and thereby to drive the transition to E2P.The sarco(endo)plasmic reticulum Ca2+-ATPase (SERCA)2 is a membrane-bound ion pump that transports Ca2+ against a steep concentration gradient, utilizing the energy derived from ATP hydrolysis (13). It belongs to the family of P-type ATPases, in which the γ-phosphoryl group of ATP is transferred to a conserved aspartic acid residue during the reaction cycle. Both phospho and dephospho forms of the enzyme undergo transitions between so-called E1 and E2 conformations (Scheme 1). The E1 and E1P states display specificity for reaction with ATP and ADP, respectively (“kinase activity”), whereas E2P and E2 react with water and Pi instead of nucleotide (“phosphatase activity”). The E1 dephosphoenzyme of the Ca2+-ATPase binds two Ca2+ ions with high affinity from the cytoplasmic side, thereby triggering the phosphorylation from ATP. In E1P, the Ca2+ ions are occluded with no access to either side of the membrane, and Ca2+ is released to the luminal side after the conformational transition to E2P, likely in exchange for protons being countertransported. The structural organization and domain movements leading to Ca2+ translocation have recently been elucidated by crystallization of SERCA in various conformational states thought to represent intermediates in the pump cycle (47). SERCA is made up of 10 membrane-spanning mostly helical segments, M1–M10 (numbered from the N terminus), of which M4–M6 and M8 contribute liganding groups for Ca2+ binding, and a cytoplasmic headpiece separated into three distinct domains, named A (“actuator”), P (“phosphorylation”), and N (“nucleotide binding”). The A-domain appears to undergo considerable movement during the functional cycle. In the E1/E1P states, the highly conserved TGE183S loop of the A-domain is at great distance from the catalytic center containing nucleotide-binding residues and the phosphorylated Asp351 of the P-domain, but during the Ca2E1P → E2P transition, the A-domain rotates ∼90° around an axis perpendicular to the membrane, thereby moving the TGE183S loop into close contact with the catalytic site such that Glu183 can catalyze dephosphorylation of E2P (8, 9). During the dephosphorylation, Glu183 likely coordinates the water molecule attacking the aspartyl phosphoryl bond and withdraws a hydrogen. Hence, the movement of the A-domain during the Ca2E1P → E2P transition is the event that changes the catalytic specificity from kinase activity to phosphatase activity. During the dephosphorylation of E2P → E2, there is only a slight change of the position of the A-domain, and a large back-rotation is needed to reach the E1 form from E2; thus, the A-domain rotation defines the difference between the E1/E1P class of conformations and the E2/E2P class. Because the A-domain is physically connected to transmembrane helices M1–M3 through the linker segments A-M1, A-M2, and A-M3, the A-domain movement occurring during the Ca2E1P → E2P transition may be a key event in the opening of the Ca2+ sites toward the lumen, thus explaining the coupling of ATP hydrolysis to Ca2+ translocation. An important unanswered question is, however, how the movement of the A-domain is brought about. Which are the driving forces that destabilize Ca2E1P and/or stabilize E2P such that the energy-transducing Ca2E1P → E2P transition takes place? To answer this, it seems important to elucidate the exact roles of the linkers. Intriguing results have been obtained by Suzuki and co-workers, who demonstrated the importance of the A-M1 linker in connection with luminal release of Ca2+ from E2P (10). In this study, we have addressed the role of the A-M3 linker. An alignment of two crystal structures thought to resemble the Ca2E1P and E2·Pi forms (5), respectively, is shown in Fig. 1. The A-domain rotation is associated with formation of a helix (“A3 helix”) in the N-terminal part of the A-M3 linker, and this helix seems to interact with a helix bundle consisting of the P5–P7 helices of the P-domain, a feature exhibited by all published crystal structures of the E2 type (cf. supplemental Fig. S1 and Ref. 11). Moreover, when structures of similar crystallographic resolution are compared (as in Fig. 1), the non-helical part of the A-M3 linker in E2-type structures has a higher relative temperature factor (“B-factor”) than the corresponding segment in Ca2E1P (Fig. 1C, thick part colored orange-red for high temperature factor), thus suggesting a higher degree of freedom of movement relative to Ca2E1P. Hence, the A-M3 linker appears more strained in Ca2E1P compared with E2 forms, and the greater flexibility of the linker in E2 forms may promote the formation of the A3 helix.Open in a separate windowSCHEME 1.Ca2+-ATPase reaction cycle.Open in a separate windowFIGURE 1.A-M3 linker configuration in E1- and E2-type crystal structures. Crystal structures with Protein Data Bank codes 2zbd (Ca2E1P analog) and 1wpg (E2·Pi analog) are shown aligned. A, overview of structure 2zbd in bluish colors with green A-M3 linker and structure 1wpg in reddish colors with wheat A-M3 linker. B, magnification of the A-M3 linker (corresponding to the red box in A) with arrows indicating site 1, between Glu243 and Gln244, and site 2, between Gly233 and Lys234, in both conformations. The green A-M3 linker to the right is structure 2zbd. The wheat A-M3 linker to the left is structure 1wpg. Note the kinked A3 helix forming part of the latter structure. C, same A-M3 linker structures as in B but with the magnitude of the temperature factor (B-factor) indicated in colors (red > orange > yellow > green > blue) and by tube diameter. Because the two crystal structures selected here as E1- and E2-type representatives have similar crystallographic resolution (2.40 and 2.30 Å, respectively), the differences in temperature factor in specific regions provide direct information about chain flexibility.Here, we have determined the functional consequences of changing the length (and thereby likely the strain) of the A-M3 linker. Polyglycine and glycine-proline loops of varying lengths were inserted at two different sites in the linker (Fig. 1), and deletions were also studied. Rather unexpectedly, we were able to insert as many as 41 residues in one of the sites without loss of expression or ability to react with Ca2+ and ATP, forming Ca2E1P, but the Ca2E1P → E2P transition was greatly affected.  相似文献   

20.
The inositol 1,4,5-trisphosphate (IP3) receptor (IP3R) is an intracellular IP3-gated calcium (Ca2+) release channel and plays important roles in regulation of numerous Ca2+-dependent cellular responses. Many intracellular modulators and IP3R-binding proteins regulate the IP3R channel function. Here we identified G-protein-coupled receptor kinase-interacting proteins (GIT), GIT1 and GIT2, as novel IP3R-binding proteins. We found that both GIT1 and GIT2 directly bind to all three subtypes of IP3R. The interaction was favored by the cytosolic Ca2+ concentration and it functionally inhibited IP3R activity. Knockdown of GIT induced and accelerated caspase-dependent apoptosis in both unstimulated and staurosporine-treated cells, which was attenuated by wild-type GIT1 overexpression or pharmacological inhibitors of IP3R, but not by a mutant form of GIT1 that abrogates the interaction. Thus, we conclude that GIT inhibits apoptosis by modulating the IP3R-mediated Ca2+ signal through a direct interaction with IP3R in a cytosolic Ca2+-dependent manner.The inositol 1,4,5-trisphosphate (IP3)3 receptor (IP3R) consisting of three subtypes, IP3R1, IP3R2, and IP3R3, is a tetrameric intracellular IP3-gated calcium (Ca2+) release channel localized at the endoplasmic reticulum (ER) with its NH2 terminus and COOH-terminal tail (CTT) exposed to the cytoplasm (1, 2; see Fig. 1A). IP3Rs are composed of five functional domains. The long NH2-terminal cytoplasmic region contains three domains, a coupling/suppressor domain, an IP3-binding core domain, and an internal coupling domain. The COOH-terminal region has a six-membrane spanning channel domain and a short cytoplasmic CTT “gatekeeper domain” that is critical for IP3R channel opening (2, 3). Ca2+ release activity of the IP3R channel is regulated by many intracellular modulators (ATP, calmodulin, and Ca2+), protein kinases, and IP3R-binding proteins (2, 4), and the tight regulation of IP3R channel activity by these factors generates various spatial and temporal intracellular Ca2+ patterns such as Ca2+ spikes and Ca2+ oscillations, leading to numerous cellular responses (1, 2, 5, 6).Open in a separate windowFIGURE 1.GIT1 and GIT2 bind to all three subtypes of IP3R. A, schematic of ER residential IP3R. The CTT of IP3R1 is used as bait in a yeast two-hybrid screen. B, schematic representation of GIT1, GIT2, and two GIT1 fragments identified from the yeast two-hybrid screen. Functional domains are indicated. ARF-GAP, ARF-specific GTPase-activating protein domain; ANK-REP, ankyrin repeats; CC, coiled-coil domains; SHD, the Spa2-homology domain; EF, EF-hand; IQ, IQ-like motifs; aa, amino acid. C, GIT1 binds to IP3R1 in vitro. GST and GST-IP3R1/CTT were incubated with mouse brain lysate for a pull-down assay. The input and pulled-down samples were probed with α-GIT1. D and E, GIT1 binds to IP3R1 in vivo. Mouse brain lysates were processed to control IgG and α-IP3R1 (D) or α-GIT1 (E) for IP. The input and IP samples were probed with α-GIT1 and α-IP3R1. F and G, both GIT1 and GIT2 bind to all three IP3R subtypes. HeLa cells coexpressing GFP-fused IP3R1, IP3R2, or IP3R3 and mRFP-fused GIT1 (F) or GIT2 (G) were processed for IP using α-RFP. The input and IP samples were blotted with α-GFP (top) and α-RFP (bottom).One of the physiological roles of IP3R-mediated Ca2+ signaling is a pro-apoptotic regulator during apoptosis. Ca2+ released from ER can stimulate several key enzymes activated during apoptosis such as endonucleases (7) and calpain (8). In addition, the close proximity of ER to mitochondria may facilitate the mitochondrial overload of Ca2+ released from the IP3Rs with certain apoptotic stimuli, triggering the opening of the mitochondrial permeability transition pore and the release of apoptotic signaling molecules, such as cytochrome c and apoptosis-inducing factor, which leads to the activation of caspases (5, 6). Moreover, several key components of apoptotic cascades, such as cytochrome c (9) and anti-apoptosis proteins Bcl-2 (10, 11) and Bcl-XL (12), have been reported to interact with the internal coupling domain and/or the CTT of IP3R and enhance the Ca2+-release activity of IP3Rs during apoptosis. In this study, we identified the ubiquitously expressed G-protein-coupled receptor kinase-interacting proteins (GIT) (13), GIT1 and GIT2, as novel IP3R-binding proteins that bind to the CTT of IP3R and inhibit apoptosis by regulation of IP3R-mediated Ca2+ signal.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号