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 78 毫秒
1.
We tested the hypothesis that extracellular lactate regulates the function of pericyte-containing retinal microvessels. Although abluminally positioned pericytes appear to adjust capillary perfusion by contracting and relaxing, knowledge of the molecular signals that regulate the contractility of these mural cells is limited. Here, we focused on lactate because this metabolic product is in the retinal extracellular space under both physiological and pathophysiological conditions. In microvessels freshly isolated from the adult rat retina, we used perforated-patch pipettes to monitor ionic currents, fura-2 to measure calcium levels, and time-lapse photography to visualize changes in mural cell contractility and lumen diameter. During lactate exposure, pericyte calcium rose; these cells contracted, and lumens constricted. This contractile response appears to involve a cascade of events resulting in the inhibition of Na+/Ca2+ exchangers (NCXs), the decreased of which function causes pericyte calcium to increase and contraction to be triggered. On the basis of our observation that gap junction uncouplers minimized the lactate-induced rise in pericyte calcium, we propose that the NCXs inhibited by lactate are predominately located in the endothelium. Indicative of the importance of endothelial/pericyte gap junctions, uncouplers of these junctions switched the pericyte response to lactate from contraction to relaxation. In addition, we observed that hypoxia, which closes microvascular gap junctions, also switched lactate's effect from vasocontraction to vasorelaxation. Thus the response of pericyte-containing retinal microvessels to extracellular lactate is metabolically modulated. The ability of lactate to serve as a vasoconstrictor when energy supplies are ample and a vasodilator under hypoxic conditions may be an efficient mechanism to link capillary function with local metabolic need.  相似文献   

2.
Pan ZH  Hu HJ  Perring P  Andrade R 《Neuron》2001,32(1):89-98
Transmitter release in neurons is thought to be mediated exclusively by high-voltage-activated (HVA) Ca(2+) channels. However, we now report that, in retinal bipolar cells, low-voltage-activated (LVA) Ca(2+) channels also mediate neurotransmitter release. Bipolar cells are specialized neurons that release neurotransmitter in response to graded depolarizations. Here we show that these cells express T-type Ca(2+) channel subunits and functional LVA Ca(2+) currents sensitive to mibefradil. Activation of these currents results in Ca(2+) influx into presynaptic terminals and exocytosis, which we detected as a capacitance increase in isolated terminals and the appearance of reciprocal currents in retinal slices. The involvement of T-type Ca(2+) channels in bipolar cell transmitter release may contribute to retinal information processing.  相似文献   

3.
Published work has shown that endothelin-1-induced contractility of bovine retinal pericytes is reduced after culture in high concentrations of glucose. The purpose of the present study was to establish the profile of endothelin-1-induced calcium transients in pericytes and to identify changes occurring after culture in high concentrations of glucose. Glucose had no effect on basal levels of cytosolic calcium or on endothelin-1-induced calcium release from intracellular stores. However, influx of calcium from the extracellular medium after endothelin-1 stimulation was reduced in pericytes that had been cultured in 25 mM D-glucose. L-type Ca(2+) currents were identified by patch clamping. The L-type Ca(2+) channel agonist, (-)-Bay K8644, caused less influx of calcium from the extracellular medium in pericytes that had been cultured in 25 mM D-glucose than in those cultured with 5 mM D-glucose. However, 3-O-methylglucose, a nonmetabolizable analogue of glucose which can cause glycation, had similar effects to those of high concentrations of glucose. The results suggest that reduced function of the L-type Ca(2+) channel that occurs in bovine retinal pericytes after culture in high concentrations of D-glucose is probably due to glycation of a channel protein.  相似文献   

4.
ANG II constricts descending vasa recta (DVR) through Ca(2+) signaling in pericytes. We examined the role of PKC DVR pericytes isolated from the rat renal outer medulla. The PKC blocker staurosporine (10 microM) eliminated ANG II (10 nM)-induced vasoconstriction, inhibited pericyte cytoplasmic Ca(2+) concentration ([Ca(2+)](cyt)) elevation, and blocked Mn(2+) influx into the cytoplasm. Activation of PKC by either 1,2-dioctanoyl-sn-glycerol (10 microM) or phorbol 12,13-dibutyrate (PDBu; 1 microM) induced both vasoconstriction and pericyte [Ca(2+)](cyt) elevation. Diltiazem (10 microM) blocked the ability of PDBu to increase pericyte [Ca(2+)](cyt) and enhance Mn(2+) influx. Both ANG II- and PDBu-induced PKC stimulated DVR generation of reactive oxygen species (ROS), measured by oxidation of dihydroethidium (DHE). The effect of ANG II was only significant when ANG II AT(2) receptors were blocked with PD-123319 (10 nM). PDBu augmentation of DHE oxidation was blocked by either TEMPOL (1 mM) or diphenylene iodonium (10 microM). We conclude that ANG II and PKC activation increases DVR pericyte [Ca(2+)](cyt), divalent ion conductance into the cytoplasm, and ROS generation.  相似文献   

5.
Loss of pericytes from the capillary wall is a hallmark of diabetic retinopathy, however, the pathogenic significance of this phenomenon is unclear. In previous mouse gene knockout models leading to pericyte deficiency, prenatal lethality has so far precluded analysis of postnatal consequences in the retina. We now report that endothelium-restricted ablation of platelet-derived growth factor-B generates viable mice with extensive inter- and intra-individual variation in the density of pericytes throughout the CNS. We found a strong inverse correlation between pericyte density and the formation of a range of retinal microvascular abnormalities strongly reminiscent of those seen in diabetic humans. Proliferative retinopathy invariably developed when pericyte density was <50% of normal. Our data suggest that a reduction of the pericyte density is sufficient to cause retinopathy in mice, implying that pericyte loss may also be a causal pathogenic event in human diabetic retinopathy.  相似文献   

6.
Primary culture of microvascular endothelial cells from bovine retina   总被引:11,自引:0,他引:11  
Summary To provide an in vitro system for studying retinal capillary function we have developed methods for isolation and culture of microvascular endothelial cells from retina. Retinal microvessels were isolated by homogenization of the retina and collection of the microvessels onto nylon mesh. Treatment of the isolated microvessels with collagenase and dispase followed by Percoll gradient centrifugation yielded endothelial cells that were largely free of pericytes. A homogeneous population of endothelial cells that were capable of at least six population doublings was obtained by plating onto a fibronectin coated substrate in plasma derived serum. The endothelial origin of these cells was confirmed by the presence of Factor VIII antigen, angiotensin converting enzyme activity, numerous tight junctions, and a cell surface that did not bind platelets. A second cell type, which did not exhibit these cell markers and which is presumably the intramural pericyte, was obtained when the isolated microvessels were plated on tissue culture grade plastic in fetal bovine serum. Supported by Research Grants 5R01-EY03772 and 5R01-ES02380 from the U.S. Public Health Service (G. W. G.) and Established Investigator Award 31-107 from the American Heart Association (A. L. B.).  相似文献   

7.
Glutamate is well established as an excitatory neurotransmitter in the vertebrate retina. Its role as a modulator of retinal function, however, is poorly understood. We used immunocytochemistry and calcium imaging techniques to investigate whether metabotropic glutamate receptors are expressed in the chicken retina and by identified GABAergic amacrine cells in culture. Antibody labeling for both metabotropic glutamate receptors 1 and 5 in the retina was consistent with their expression by amacrine cells as well as by other retinal cell types. In double-labeling experiments, most metabotropic glutamate receptor 1-positive cell bodies in the inner nuclear layer also label with anti-GABA antibodies. GABAergic amacrine cells in culture were also labeled by metabotropic glutamate receptor 1 and 5 antibodies. Metabotropic glutamate receptor agonists elicited Ca(2+) elevations in cultured amacrine cells, indicating that these receptors were functionally expressed. Cytosolic Ca(2+) elevations were enhanced by metabotropic glutamate receptor 1-selective antagonists, suggesting that metabotropic glutamate receptor 1 activity might normally inhibit the Ca(2+) signaling activity of metabotropic glutamate receptor 5. These results demonstrate expression of group I metabotropic glutamate receptors in the avian retina and suggest that glutamate released from bipolar cells onto amacrine cells might act to modulate the function of these cells.  相似文献   

8.
Abstract: The aim of this study was to purify microvessels from bovine retina and also to cultivate bovine retinal endothelial cells (BRECs) or intramural pericytes, to determine their fatty acid composition. Microvessels were obtained after Dounce homogenization of the retina followed by centrifugation on albumin cushion and finally microvessels in the pellet were trapped on a 100-µm nylon filter. Contamination of microvessel preparations by neuronal tissue, assessed after both microscopic examination and western blotting with a monoclonal antibody raised against rhodopsin, was minor. In the entire bovine retina, docosahexaenoic acid (DHA) represented 23.3% of the total fatty acids and there was about three times less arachidonic acid (AA) (8.2%) than DHA. In contrast, DHA and AA levels were almost equivalent in the retinal microvessels with ∼10% of total fatty acids. When compared with intact microvessels, the DHA proportion of confluent monolayers of both BRECs or pericytes in primary cultures dropped to ∼2% of the total fatty acids, whereas AA was unchanged. Culture medium supplementation with unesterified DHA (10 µ M ) restored the DHA proportion of BRECs close to the microvascular value at the expense of linoleic acid without affecting AA very much. In contrast, DHA supplementation in pericytes increased the DHA proportion of these cells at the expense of AA. In conclusion, DHA of intact microvessels represented 10% of the total fatty acids, which was close to the AA proportion. Mild DHA supplementation of BRECs or pericytes in primary cultures restored their DHA proportion to the original microvessel value. This high percentage of polyunsaturated fatty acids in retinal microvessels should allow us to test the hypothesis that oxidation products derived from these fatty acids may be involved in the pathogenic process leading to diabetic retinopathy.  相似文献   

9.
Membrane depolarization is the signal that triggers release of neurotransmitter from nerve terminals. As a result of depolarization, voltage-dependent Ca(2+) channels open, level of intracellular Ca(2+) increases. and release of neurotransmitter commences. Previous study had shown that in rat brain synaptosomes, muscarinic acetylcholine (ACh) receptors (mAChRs) interact with soluble NSF attachment protein receptor proteins of the exocytic machinery in a voltage-dependent manner. It was suggested that this interaction might control the rapid, synchronous release of acetylcholine. The present study investigates the mechanism for such a voltage-dependent interaction. Here we show that depolarization shifts mAChRs, specifically the m2 receptor subtype, to a low affinity state toward its agonists. At resting potential, mAChRs are in a high affinity state (K(d) of approximately 20 nM) and they shift to a low affinity state (K(d) of tens of microM) upon membrane depolarization. In addition, interaction between m2 receptor subtype and the exocytic machinery increases with receptor occupancy. Both phenomena are independent of Ca(2+) influx. We propose that these results may explain control of ACh release from nerve terminals. At resting potential the exocytic machinery is clamped due to its interaction with the occupied mAChR and depolarization relieves this interaction. This, together with Ca(2+) influx, enables release of ACh to commence.  相似文献   

10.
Microvascular pericytes are believed to be involved in various functions such as regulation of capillary blood flow and endothelial proliferation. Since pericytes represent a morphologically heterogeneous cell population ranging from circular smooth musclelike to elongated fibroblast-like morphology it is possible that regulation of blood flow (via contractility) and control of endothelial proliferation (as well as other metabolic functions) may be accomplished by different subsets of pericytes. In the present study we provide evidence for heterogeneity of pericytes at the molecular level by using two novel technical approaches. These are (a) immunostaining of whole mounts of the microvascular beds of the rat mesentery and bovine retina and (b) immunoblotting studies of microdissected retinal microvessels. We show that pericytes of true capillaries (midcapillaries) apparently lack the smooth muscle isoform of alpha-actin whereas transitional pericytes of pre- and postcapillary microvascular segments do express this isoform. Thus, regulation of capillary blood flow may be accomplished by the smooth muscle-related pre- and postcapillary pericytes whereas the nonmuscle pericytes of true capillaries may play a role in other functions.  相似文献   

11.
We tested the hypothesis that constriction of descending vasa recta (DVR) is mediated by voltage-gated calcium entry. K(+) channel blockade with BaCl(2) (1 mM) or TEACl (30 mM) depolarized DVR smooth muscle/pericytes and constricted in vitro-perfused vessels. Pericyte depolarization by 100 mM extracellular KCl constricted DVR and increased pericyte intracellular Ca(2+) ([Ca(2+)](i)). The K(ATP) channel opener pinacidil (10(-7)-10(-4) M) hyperpolarized resting pericytes, repolarized pericytes previously depolarized by ANG II (10(-8) M), and vasodilated DVR. The DVR vasodilator bradykinin (10(-7) M) also reversed ANG II depolarization. The L-type Ca(2+) channel blocker diltiazem vasodilated ANG II (10(-8) M)- or KCl (100 mM)-preconstricted DVR, and the L-type agonist BayK 8644 constricted DVR. The plateau phase of the pericyte [Ca(2+)](i) response to ANG II was inhibited by diltiazem. These data support the conclusion that DVR vasoreactivity is controlled through variation of membrane potential and voltage-gated Ca(2+) entry into the pericyte cytoplasm.  相似文献   

12.
Inducible nitric oxide synthase (iNOS) is an essential mediator in diabetic vascular lesions and known to be regulated by activation of Ca2+/calmodulin-dependent protein kinase II (CaMKII). The aim of this study was to investigate whether CaMKII affects iNOS-mediated pericyte death in the retina of diabetic mice with early stage disease. Total- and phospho-CaMKII, iNOS, and active caspase-3 protein levels were assessed by Western blotting, and CaMKII activity was measured by kinase assay. iNOS-related pericyte death was assessed by double immunofluorescent staining for iNOS and α-smooth muscle actin, followed by the TUNEL assay. Autocamtide-2-related inhibitory peptide (AIP), a specific inhibitor of CaMKII, was injected into the right vitreous 2 days before sacrifice of mice, to examine the effect of CaMKII inactivation in diabetic retinas. The levels of total- and phospho-CaMKII, iNOS, and active caspase-3 protein, and CaMKII activity were significantly increased in the diabetic retinas compared with those of control retinas. Furthermore, TUNEL-positive signals colocalized with iNOS-immunoreactive pericytes in the same retinas. However, inactivation of CaMKII by AIP treatment inhibited all these changes, which was accompanied by less pericyte loss. Our results demonstrate that CaMKII contributes to iNOS-related death of pericytes in the diabetic retina and that inactivation of this enzyme may be a potential treatment for retinal vascular lesion.  相似文献   

13.
The purpose of this study was to establish and characterize a retinal pericyte cell line from retinal capillaries of transgenic rats harboring the temperature-sensitive simian virus 40 large T-antigen gene (tsA58 Tg rat), and to apply this to the co-culture with a retinal capillary endothelial cell line. The conditionally immortalized rat retinal pericyte cell lines (TR-rPCTs), which express a temperature-sensitive large T-antigen, were obtained from two tsA58 Tg rats. These cell lines had a multicellular nodule morphology and reacted positively with von Kossa staining, a marker of calcification. TR-rPCTs cells expressed mRNA of pericyte markers such as rat intercellular adhesion molecule-1, platelet-derived growth factor-receptor beta, angiopoietin-1, and osteopontin. Western blot analysis indicated that alpha-smooth muscle actin (alpha-SMA) was expressed in TR-rPCT3 and 4 cells. In contrast, alpha-SMA was induced by transforming growth factor-beta1 and its enhancement was reduced by basic fibroblast growth factor in TR-rPCT1 and 2 cells. When TR-rPCT1 cells were cultured with a rat retinal endothelial cell line (TR-iBRB2) in a contact co-culture system, the number of TR-iBRB2 cells were significantly reduced in comparison with that of a single culture of TR-iBRB2 cells, suggesting that physical contact between pericytes and retinal endothelial cells is important for the growth of retinal endothelial cells. In conclusion, conditionally immortalized retinal pericyte cell lines were established from tsA58 Tg rats. These cell lines exhibited the properties of retinal pericytes and can be applied in co-culture systems with a retinal capillary endothelial cell line.  相似文献   

14.
We showed that muscarinic acetylcholine (ACh)-stimulation increased the cellular content of cADPR in the pancreatic acinar cells from normal mice but not in those from CD38 knockout mice. By monitoring ACh-evoked increases in the cytosolic Ca(2+) concentration ([Ca(2+)](i)) using fura-2 microfluorimetry, we distinguished and characterized the Ca(2+) release mechanisms responsive to cADPR. The Ca(2+) response from the cells of the knockout mice (KO cells) lacked two components of the muscarinic Ca(2+) release present in wild mice. The first component inducible by the low concentration of ACh contributed to regenerative Ca(2+) spikes. This component was abolished by ryanodine treatment in the normal cells and was severely impaired in KO cells, indicating that the low ACh-induced regenerative spike responses were caused by cADPR-dependent Ca(2+) release from a pool regulated by a class of ryanodine receptors. The second component inducible by the high concentration of ACh was involved in the phasic Ca(2+) response, and it was not abolished by ryanodine treatment. Overall, we conclude that muscarinic Ca(2+) signaling in pancreatic acinar cells involves a CD38-dependent pathway responsible for two cADPR-dependent Ca(2+) release mechanisms in which the one sensitive to ryanodine plays a crucial role for the generation of repetitive Ca(2+) spikes.  相似文献   

15.
M Lupu-Meiri  H Shapira  Y Oron 《FEBS letters》1990,262(2):165-169
We tested the contribution of extracellular calcium (Ca2+) to membrane electrical responses to acetylcholine (ACh) in native Xenopus oocytes. Removal of Cao caused a decrease in both the rapid (D1) and the slow (D2) chloride currents that comprise the common depolarizing response to ACh in native oocyte. The effect of Ca2+o removal on the muscarinic response was mimicked by the addition of 1 mM Mn2+, an effective antagonist of calcium influx, though not by antagonists of voltage-sensitive calcium channels. When oocytes were challenged with ACh in Ca2(+)-free medium, subsequent addition of 1.8 mM CaCl2 resulted in a rapid, often transient, depolarizing current. Similarly to the Ca2+o-dependent component of membrane electrical responses, the Ca2(+)-evoked current was reversibly abolished by Mn2+, though not by antigonists of voltage-sensitive calcium channels. Depletion of cellular calcium potentiated the Ca2(+)-evoked current, implying negative feedback of calcium channels by calcium. Injection of 10-100 fmol of inositol 1,4,5-trisphosphate (IP3) resulted in a two-component depolarizing current. IP3 injection promoted the appearance of Ca2+o-evoked current that was significantly potentiated by previous calcium depletion. We suggest that activation of cell-membrane muscarinic receptors causes opening of apparently voltage-insensitive and verapamil or diltiazem-resistant calcium channels. These channels may be activated by IP3 or its metabolites, which increase following the activation of cell membrane receptors coupled to a phospholipase C. The channels may be identical to receptor-operated channels described in other model systems.  相似文献   

16.
Neurons, sensory cells and endocrine cells secrete neurotransmitters and hormones to communicate with other cells and to coordinate organ and system function. Validation that a substance is used as an extracellular signaling molecule by a given cell requires a direct demonstration of its secretion. In this protocol we describe the use of biosensor cells to detect neurotransmitter release from endocrine cells in real-time. Chinese hamster ovary cells expressing the muscarinic acetylcholine (ACh) receptor M3 were used as ACh biosensors to record ACh release from human pancreatic islets. We show how ACh biosensors loaded with the Ca(2+) indicator Fura-2 and pressed against isolated human pancreatic islets allow the detection of ACh release. The biosensor approach is simple; the Ca(2+) signal generated in the biosensor cell reflects the presence (release) of a neurotransmitter. The technique is versatile because biosensor cells expressing a variety of receptors can be used in many applications. The protocol takes ~3 h.  相似文献   

17.
Recent findings suggest that astrocytes respond to neuronally released neurotransmitters with Ca2+ elevations. These Ca2+ elevations may trigger astrocytes to release glutamate, affecting neuronal activity. Neuronal activity is also affected by modulatory neurotransmitters that stimulate G protein-coupled receptors. These neurotransmitters, including acetylcholine and histamine, might affect neuronal activity by triggering Ca2+-dependent release of neurotransmitters from astrocytes. However, there is no physiological evidence for histaminergic or cholinergic receptors on astrocytes in situ. We asked whether astrocytes have these receptors by imaging Ca2+-sensitive dyes sequestered by astrocytes in hippocampal slices. Our results show that immunocytochemically identified astrocytes respond to carbachol and histamine with increases in intracellular free Ca2+ concentration. The H1 histamine receptor antagonist chlorpheniramine inhibited responses to histamine. Similarly, atropine and the M1-selective muscarinic receptor antagonist pirenzepine inhibited carbachol-elicited responses. Astrocyte responses to histamine and carbachol were compared with responses elicited by alpha1-adrenergic and metabotropic glutamate receptor agonists. Individual astrocytes responded to different subsets of receptor agonists. Ca2+ oscillations were the prevalent response pattern only with metabotropic glutamate receptor stimulation. Finally, functional alpha1-adrenergic receptors and muscarinic receptors were not detected before postnatal day 8. Our data show that astrocytes have acetylcholine and histamine receptors coupled to Ca2+. Given that Ca2+ elevations in astrocytes trigger neurotransmitter release, it is possible that these astrocyte receptors modulate neuronal activity.  相似文献   

18.
Human mesenchymal stem cells (HMSC) have the potential to differentiate into many cell types. The physiological properties of HMSCs including their Ca(2+) signaling pathways, however, are not well understood. We investigated Ca(2+) influx and release functions in HMSCs. In Ca(2+) imaging experiments, spontaneous Ca(2+) oscillations were observed in 36 of 50 HMSCs. The Ca(2+) oscillations were completely blocked by the application of 10 micro M cyclopiazonic acid (CPA) or 1 micro M thapsigargin (TG). A brief application of 1 micro M acetylcholine (ACh) induced a transient increase of [Ca(2+)](i) but the application of caffeine (10 mM) did not induce any Ca(2+) transient. When the stores were depleted with Ca(2+)-ATPase blockers (CPA or TG) or muscarinic agonists (ACh), store-operated Ca(2+) (SOC) entry was observed. Using the patch-clamp technique, store-operated Ca(2+) currents (I(SOC)) could be recorded in cells treated with ACh or CPA, but voltage-operated Ca(2+) currents (VOCCs) were not elicited in most of the cells (17/20), but in 15% of cells examined, small dihydropyridine (DHP)-sensitive Ca(2+) currents were recorded. Using RT-PCR, mRNAs were detected for inositol 1,4,5-trisphosphate receptor (InsP(3)R) type I, II, and III and DHP receptors alpha1A and alpha1H were detected, but mRNA was not detected for ryanodine receptor (RyR) or N-type Ca(2+) channels. These results suggest that in undifferentiated HMSCs, Ca(2+) release is mediated by InsP(3)Rs and Ca(2+) entry through plasma membrane is mainly mediated by the SOCs channels with a little contribution of VOCCs.  相似文献   

19.
Ca(2+) influx through L-type channels is critical for numerous physiological functions. Relatively little is known about modulation of neuronal L-type Ca(2+) channels. We studied modulation of neuronal Ca(V)1.2c channels heterologously expressed in HEK293 cells with each of the known muscarinic acetylcholine receptor subtypes. Galphaq/11-coupled M1, M3, and M5 receptors each produced robust inhibition of Ca(V)1.2c, whereas Galphai/o-coupled M2 and M4 receptors were ineffective. Channel inhibition through M1 receptors was studied in detail and was found to be kinetically slow, voltage-independent, and pertussis toxin-insensitive. Slow inhibition of Ca(V)1.2c was blocked by coexpressing RGS2 or RGS3T or by intracellular dialysis with antibodies directed against Galphaq/11. In contrast, inhibition was not reduced by coexpressing betaARK1ct or Galphat. These results indicate that slow inhibition required signaling by Galphaq/11, but not Gbetagamma, subunits. Slow inhibition did not require Ca(2+) transients or Ca(2+) influx through Ca(V)1.2c channels. Additionally, slow inhibition was insensitive to pharmacological inhibitors of phospholipases, protein kinases, and protein phosphatases. Intracellular BAPTA prevented slow inhibition via a mechanism other than Ca(2+) chelation. The cardiac splice-variant of Ca(V)1.2 (Ca(V)1.2a) and a splice-variant of the neuronal/neuroendocrine Ca(V)1.3 channel also appeared to undergo slow muscarinic inhibition. Thus, slow muscarinic inhibition may be a general characteristic of L-type channels having widespread physiological significance.  相似文献   

20.
1. Agonist activation of rat retina muscarinic receptors results in suppression of cyclic AMP (cAMP) generation and enhanced phosphoinositide hydrolysis. 2. Pharmacological manipulations that elevate cAMP or stable analogues of cAMP attenuate the acetylcholine (ACh)-induced enhancement of phosphoinositide hydrolysis. We postulate that cross-talk between adenylate cyclase and phospholipase C signal transducing systems probably exists in rat retina, as has been described for other systems. 3. Intraocular administration of pertussis toxin attenuated the response of both adenylate cyclase and phospholipase C to muscarinic stimulation, suggesting that some retinal muscarinic receptors are apparently coupled to their effector systems via pertussis toxin sensitive G proteins.  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