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 250 毫秒
1.
Abstract— The effects of accumulated endogenous GABA on the activity of L-glutamic acid decarboxylase (GAD) were studied in mouse brain. When the content of GABA in the brain was increased after administration in vivo of aminooxyacetic acid (AOAA), there was a reduction of GAD activity which could not be reversed by the addition of pyridoxal-5′-phosphate (PLP). Since inhibition of GAD activity by AOAA could be readily reversed by PLP, the reduction of GAD activity measured in the presence of added PLP indicated a decrease in the level of GAD apoenzyme. Similarly, increase of GABA content by hydrazine was also accompanied by a reduction in the level of GAD. Thiosemicarbazide and hydroxylamine did not affect the content of GABA appreciably, and in both cases levels of GAD remained unchanged when measured in the presence of added PLP. The correlation of the reduction in the levels of GAD with the increases in content of GABA suggests that GABA may regulate its own synthesizing enzyme by feedback repression.  相似文献   

2.
Abstract: Ethanolamine O-sulphate (EOS) dissolved in the drinking water (5mg-ml−1) was administered ad libitum to rats for 26 days. At the end of this period, glutamate decarboxylase (GAD) and GABA-transaminase (GABA-T) activities, 4-aminobutyrate (GABA) concentration, and the levels of six other amino acids were measured in various brain regions. Significant inhibition of GABA-T accompanied by significant increases in GABA content were observed throughout the brain, although the magnitudes of these effects varied according to region. GAD activity was significantly reduced in most brain regions, although this effect was apparently not related to cofactor availability or the direct actions of EOS or increased GABA concentration. Glutamine levels were significantly reduced to approximately 72% of control values in all brain regions. Aspartate levels were significantly reduced to approximately 84% of control values in all regions except the striatum and cerebellum. Minor changes in other amino acid levels were also detected. These neurochemical changes which accompanied the primary effect of EOS on GABA-T are discussed in terms of indirect secondary metabolic changes rather than nonspecific enzyme inhibition by EOS.  相似文献   

3.
We have previously shown that short-lasting reduction of cerebral blood flow by bilateral clamping of carotid arteries (BCCA) results in long-lasting increase in regional GABA concentration and decrease in seizure susceptibility in rats. In the present experiments, the effect of BCCA on GABA turnover and the enzymes involved in GABA synthesis and degradation were studied in rats. Regional GABA turnover was measured by means of GABA accumulation induced by the GABA-transaminase (GABA-T) inhibitor aminooxyacetic acid (AOAA). Fourteen days after BCCA, GABA turnover was significantly increased in hippocampus, substantia nigra and cortex, but not different from sham-operated controls in several other brain regions, including striatum, hypothalamus and cerebellum. The activity of glutamate decarboxylase (GAD) measured ex vivo did not show any changes in investigated structures, while the activity of GABA-T was slightly increased in hippocampus. The increased GABA turnover in some brain regions may explain our previous findings of increased GABA content in these brain regions and decreased sensitivity of BCCA treated animals to the GABAA-receptor antagonist bicuculline.  相似文献   

4.
We recently reported that the mammalian brain has two forms of the GABA synthetic enzyme glutamate decarboxylase (GAD, E.C. 4.1.1.15), which are the products of two genes. The two forms, which we call GAD65 and GAD67, differ from each other in sequence, molecular size, subcellular distribution, and interactions with the cofactor pyridoxal phosphate (PLP), with GAD65 activity more dependent than that of GAD67 on the continued presence of exogenous PLP. The existence of two GAD genes suggests that individual GABA neurons may be subject to differential regulation of GABA production. We have examined the expression of these two forms of GAD during postnatal development of the rat striatum to determine whether different classes of GABA neurons selectively express different amounts of the two GAD mRNAs. Here we present evidence for a dramatic developmental difference in the expression of the two mRNAs during postnatal development of the rat striatum. Using in situ hybridization to the two GAD mRNAs, we observed a selective increase in GAD65 mRNA during the second postnatal week, at the time when striatal matrix neurons innervate the substantia nigra (SN). PLP-dependent enzyme activity in the midbrain increases in parallel with increased expression of GAD65 mRNA in the striatum. We hypothesize that the innervation of the SN by striatal neurons triggers an increase in GAD65. The changing ratios of GAD65 and GAD67 in the striatum may contribute to the well-documented changes in seizure susceptibility that occur in early life.  相似文献   

5.
The pyridoxal 5‐phosphate (PLP) cofactor is a significant organic molecule in medicinal chemistry. It is often found covalently bound to lysine residues in proteins to form PLP dependent enzymes. An example of this family of PLP dependent enzymes is γ‐aminobutyric acid aminotransferase (GABA‐AT) which is responsible for the degradation of the neurotransmitter GABA. Its inhibition or inactivation can be used to prevent the reduction of GABA concentration in brain which is the source of several neurological disorders. As a test case for PLP dependent enzymes, we have performed molecular dynamics simulations of GABA‐AT to reveal the roles of the protein residues and its cofactor. Three different states have been considered: the apoenzyme, the holoenzyme, and the inactive state obtained after the suicide inhibition by vigabatrin. Different protonation states have also been considered for PLP and two key active site residues: Asp298 and His190. Together, 24 independent molecular dynamics trajectories have been simulated for a cumulative total of 2.88 µs. Our results indicate that, unlike in aqueous solution, the PLP pyridine moiety is protonated in GABA‐AT. This is a consequence of a pKa shift triggered by a strong charge–charge interaction with an ionic “diad” formed by Asp298 and His190 that would help the activation of the first half‐reaction of the catalytic mechanism in GABA‐AT: the conversion of PLP to free pyridoxamine phosphate (PMP). In addition, our MD simulations exhibit additional strong hydrogen bond networks between the protein and PLP: the phosphate group is held in place by the donation of at least three hydrogen bonds while the carbonyl oxygen of the pyridine ring interacts with Gln301; Phe181 forms a π–π stacking interaction with the pyridine ring and works as a gate keeper with the assistance of Val300. All these interactions are hypothesized to help maintain free PMP in place inside the protein active site to facilitate the second half‐reaction in GABA‐AT: the regeneration of PLP‐bound GABA‐AT (i.e., the holoenzyme). Proteins 2016; 84:875–891. © 2016 Wiley Periodicals, Inc.  相似文献   

6.
The time course of the effects of aminooxyacetic acid, γ-vinyl GABA, γ-acetylenic GABA, gabaculine, ethanolamine-O-sulphate (EOS) and valproic acid (VPA) on brain GABA content and the activities of glutamic acid decarboxylase (GAD) and GABA aminotransferase (GABA-T), the enzymes involved in biosynthesis and degradation of GABA, was re-determined and compared with the action on the electroconvulsive threshold in mice. All drugs caused significant increases in the seizure threshold, and the temporal pattern of this effect correlated rather well with the induced elevation of brain GABA. However, no clear relationship was found between the extent of GABA increase and the relative increase of seizure threshold. Except for VPA, the time course of the increment in brain GABA followed closely the inhibition of GABA-T. The activity of GAD was gradually decreased by γ-acetylenic GABA and a slow decline of GAD activity was also observed after γ-vinyl GABA. EOS and gabaculine suggesting a feedback repression of GAD synthesis by highly elevated GABA concentrations. Concomitant with significant reduction of GAD activity, a decrease in seizure threshold occurred though brain GABA levels remained markedly elevated. On the other hand, following administration of VPA the effect of GABA levels was paralleled by an increase in GAD activity indicating that the GABA-elevating action of this drug can be attributed at least in part to an activation of GABA synthesis. The data suggest that reduction of GAD activity may be an inevitable consequence of increasing brain GABA concentrations over a certain extent and this effect seems to limit the anticonvulsant efficacy of GABA-T inhibitors.  相似文献   

7.
The regulation of glutamate decarboxylase (GAD; EC 4.1.1.15) was studied by using cultures of cerebral cortical neurons from rat brain grown in serum-free medium. About 50% of the neurons in the cultures were gamma-aminobutyric acid (GABA)ergic as determined by two double-staining procedures. Immunoblotting experiments with four anti-GAD sera that recognize the two forms to varying degrees, demonstrated that the cultures contained the two forms of GAD that are present in rat brain (apparent molecular masses = 63 and 66 kDa). GAD activity was reduced by 60-70% when intracellular GABA levels were increased by incubating the cultures with the GABA-transaminase inhibitor gamma-vinyl-GABA for greater than 5-10 h or with 1 mM GABA itself. Neither baclofen nor muscimol (100 microM) affected GAD activity. Immunoblotting experiments showed that only the larger of the two forms of GAD (66 kDa) was decreased by elevated GABA levels. These results, together with previous results indicating that the smaller form of GAD is more strongly regulated by pyridoxal 5'-phosphate (the cofactor for GAD), suggest that the two forms of GAD are regulated by different mechanisms.  相似文献   

8.
gamma-Aminobutyric acid (GABA) and its synthetic enzyme, glutamate decarboxylase (GAD), are not limited to the nervous system but are also found in nonneural tissues. The mammalian brain contains at least two forms of GAD (GAD67 and GAD65), which differ from each other in size, sequence, immunoreactivity, and their interaction with the cofactor pyridoxal 5'-phosphate (PLP). We used cDNAs and antibodies specific to GAD65 and GAD67 to study the molecular identity of GADs in peripheral tissues. We detected GAD and GAD mRNAs in rat oviduct and testis. In oviduct, the size of GAD, its response to PLP, its immunoreactivity, and its hybridization to specific RNA and DNA probes all indicate the specific expression of the GAD65 gene. In contrast, rat testis expresses the GAD67 gene. The GAD in these two reproductive tissues is not in neurons but in nonneural cells. The localization of brain GAD and GAD mRNAs in the mucosal epithelial cells of the oviduct and in spermatocytes and spermatids of the testis shows that GAD is not limited to neurons and that GABA may have functions other than neurotransmission.  相似文献   

9.
The interactions between glutamate decarboxylase (GAD) and its cofactor pyridoxal phosphate (PLP) play a key role in the regulation of GAD activity. The enzyme has two isoforms, GAD65 and GAD67. A comparison of binding constants, rate constants, and kinetic profiles for the formation of holoenzyme (holoGAD65 and holoGAD67) revealed that the two isoforms interact distinctively with the cofactor. GAD67 exhibits a higher binding constant for PLP binding, making it more difficult to dissociate PLP from holoGAD67 than holoGAD65. Meanwhile, PLP binding occurs at a much slower rate for GAD67 than GAD65, as evidenced by lower rate constants and a slower initial rate of the holoenzyme formation. Job's plots revealed a stoichiometry of 1:1 for PLP binding to GAD65 before and after the saturation level of PLP, while 1:2 for PLP binding to GAD67 prior to the saturation of PLP and 1:1 at the saturation level of PLP. These results suggested that the two binding sites of GAD65 exhibit similar affinities for PLP. In contrast, one binding site of GAD67 exhibits a significantly higher affinity for PLP than the other binding site. Based on these findings, it was proposed that a slower PLP binding to GAD67 than GAD65 and a less ease to dissociate PLP from holoGAD67 than holoGAD65 are important underlying factors. This attributes to GAD67 being more highly saturated by PLP and GAD65 being less saturated by PLP. A larger conformation change constant for GAD67 than GAD65 supported a significant conformational change induced by the initial PLP binding to GAD67, which affects the other binding site affinity of GAD67. The present studies provided valuable insights into distinctive properties between the two isoforms of GAD.  相似文献   

10.
The recent identification of two genes encoding distinct forms of the GABA synthetic enzyme, glutamate decarboxylase (GAD), raises the possibility that varying expression of the two genes may contribute to the regulation of GABA production in individual neurons. We investigated the postnatal development the two forms of GAD in the rat cerebellum. The mRNA for GAD67, the form which is less dependent on the presence of the cofactor, pyridoxal phosphate (PLP), is present at birth in presumptive Purkinje cells and increases during postnatal development. GAD67 mRNA predominates in the cerebellum. The mRNA for GAD65, which displays marked PLP-dependence for enzyme activity, cannot be detected in cerebellar cortex by in situ hybridization until P7 in Purkinje cells, and later in other GABA neurons. In deep cerebellar nuclei, which mature prenatally, both forms of GAD mRNA can be detected at birth. The amounts of immunoreactice GAD and GAD enzyme activity parallel changes in mRNA levels. We suggest that the delayed appearance of GAD65 is coincident with synapse formation between GABA neurons and their targets during the second postnatal week. GAD67 mRNA may be present prior to synaptogenesis to produce GABA for trophic and metabolic functions.Special issue dedicated to Dr. Eugene Roberts.  相似文献   

11.
The intracerebroventricular injection of pyridoxal phosphate (PLP, 0.125-1.25 μmol/rat) causes epileptic seizures (4 min → 1 min) that are preventable or reversible by GABA (1 μmol/rat), by muscimol (O.025 μmol/rat), or by diazepam (1.75 μmol/rat). At the peak of PLP-induced convulsions, the activities of GAD and GABA-T in 14 regions of rat brain remained unaltered, whereas the concentrations of PLP remained elevated. The PLP-induced convulsion was blocked by DABA (10 μmol/rat) but was not altered by β-alanine (50 μmol/rat). The previous in vitro studies have shown that PLP increases the uptake of [3H]GABA into synaptosomes and inhibits the binding of [3H]GABA to synaptic membranes. These data suggest that PLP-induced convulsion is due to reduced availability of GABA to its recognition sites, rather than to alteration in the activity of GABA metabolizing enzymes, or unavailability of PLP as a coenzyme for GAD and GABA-T. Since the duration of PLP-induced epileptic seizures is short and can be prevented by GABA agonists, PLP may be used as a tool to study the nature of GABA-mediated neuroinhibition and the properties of GABA receptor sites.  相似文献   

12.
The presence of gamma-aminobutyric acid (GABA) as well as glutamic acid decarboxylase (GAD) and GABA-transaminase (GABA-T) enzymes was demonstrated in the cockroach (Periplaneta americana) brain. Isonicotinic acid hydrazide (INH) in vivo (2.19 mumol/g) inhibited brain GAD activity, the inhibition lasted for about 2 hours and the normal activity levels reappeared at 4 h after INH administration. Brain GABA levels increased initially but then declined and were restored to normal levels at 4 h after INH administration. GABA-T activity was strongly inhibited by INH and a total 100% inhibition was observed at 2-3 h following INH treatment. The GABA-T activity, however, began to recover after 3 h but only 37% of the total enzyme activity was released from inhibition. Mercaptopropionic acid (MPA) in vivo (32 micrograms/g) inhibited brain GAD activity and depleted GABA level also. Results indicate that INH response of the cockroach brain GABA system is similar to that reported for the chick brain but differs from that of the mammalian brain.  相似文献   

13.
The effect of diazepam and nitrazepam (0.5, 1, 2.5, and 5 mg/kg) on GABA-transaminase activity in the rat cerebellar cortex and hippocampus was investigated by quantitative histochemistry. Both the drugs suppressed the enzyme activity in the hippocampus; the maximal effect was produced by diazepam. It is suggested that the observed alterations are caused by the enzyme inhibition and also by a decrease of GABA turnover rate.  相似文献   

14.
Summary The formation of GABA from L-glutamate was investigated in homogenates of rat brain, liver, and kidney, using highly purified [14C]-L-glutamic acid as substrate and a thin-layer chromatographic separation of products. In agreement with other workers, liberation of [14C]-CO2 was found to be stoichiometric with GABA formation in brain homogenates, but not in liver or kidney extracts. Subcellular fractionation and dialysis experiments suggested that most of the GABA synthesis in these peripheral tissues, unlike brain, does not occur via a direct decarboxylation of glutamate and requires one or more cofactors other than pyridoxal phosphate. NAD stimulated GABA formation in dialyzed extracts, and inhibition of GABA-transaminase, bothin vitro andin vivo, caused marked inhibition of GABA formation from glutamate in peripheral extracts. Although a very low GAD activity in liver and kidney cannot be excluded, these experiments suggest a major pathway from glutamate to GABA in these homogenates which includes (1) conversion of glutamate to -ketoglutarate by glutamate dehydrogenase or transaminases, (2) conversion of -ketoglutarate to succinic semialdehyde, and (3) formation of GABA from succinic semialdehyde and glutamate by GABA-transaminase.  相似文献   

15.
The subcellular distribution of pyridoxal phosphate (PLP) was studied in mouse brain, as well as the effect of pyridoxal phosphate-γ-glutamyl hydrazone (PLPGH—a convulsant drug which decreases both PLP levels and glutamate decarboxylase activity [GAD] in whole brain) upon both the PLP concentration and the GAD activity in subcellular fractions. An electron microscopic evaluation of the subcellular particles of control and PLPGH-treated animals was also carried out. The main findings were the following: (1) PLP was localized mainly in the supernatant and crude mitochondrial fractions; two-thirds of the amount present in the latter were located in the subfraction containing pure mitochondria, and the remainder was in the synaptosomal fraction. After osmotic disruption of synaptosomes, PLP was found in both the intrasynaptosomal mitochondria and the synaptoplasm. (2) Treatment of mice with PLPGH decreased levels of PLP in several brain fractions, this effect being much more notable in the soluble fractions than in the particulate fractions. After osmotic disruption of the synaptosomes, a specific decrease of PLP in the synaptoplasm was observed. (3) Treatment with PLPGH produced also an inhibition of GAD activity in most of the fractions studied, when this enzyme was assayed in the absence of PLP. In general, the inhibition was greater in those fractions in which levels of PLP were also affected. In synaptosomes, this correlation between the decreased levels of PLP and decreased activity of GAD occurred only in the synaptoplasm. (4) The activation of GAD by PLP added to incubation mixtures was much greater in those fractions from PLPGH-treated animals which displayed extensive inhibition of GAD, in comparison to the corresponding fractions from control animals. (5) No ultrastructural changes were detected in the subcellular fractions from treated animals. Our results show that the decreases of both the levels of PLP and the activity of GAD (as previously found in whole brain) actually occur in the synaptosomes, a finding that supports the hypothesis that the role of PLP in the mechanisms controlling excitability can be explained, at least in part, by its regulatory action on GAD activity, which in turn determines the rate of GABA synthesis at the nerve endings.  相似文献   

16.
Regulation of γ-Aminobutyric Acid Synthesis in the Brain   总被引:3,自引:3,他引:0  
Abstract: γ-Aminobutyric acid (GABA) is synthesized in brain in at least two compartments, commonly called the transmitter and metabolic compartments, and because reglatory processes must serve the physiologic function of each compartment, the regulation of GABA synthesis presents a complex problem. Brain contains at least two molecular forms of glutamate decarboxylase (GAD), the principal synthetic enzyme for GABA. Two forms, termed GAD65 and GAD67, are the products of two genes and differ in sequence, molecular weight, interaction with the cofactor, pyridoxal 5′-phosphate (pyridoxal-P), and level of expression among brain regions. GAD65 appears to be localized in nerve terminals to a greater degree than GAD67, which appears to be more uniformly distributed throughout the cell. The interaction of GAD with pyridoxal-P is a major factor in the short-term regulation of GAD activity. At least 50% of GAD is present in brain as apoenzyme (GAD without bound cofactor; apoGAD), which serves as a reservoir of inactive GAD that can be drawn on when additional GABA synthesis is needed. A substantial majority of apoGAD in brain is accounted for by GAD65, but GAD67 also contributes to the pool of apoGAD. The apparent localization of GAD65 in nerve terminals and the large reserve of apo-GAD65 suggest that GAD65 is specialized to respond to short-term changes in demand for transmitter GABA. The levels of apoGAD and the holoenzyme of GAD (holoGAD) are controlled by a cycle of reactions that is regulated by physiologically relevant concentrations of ATP and other polyanions and by inorganic phosphate, and it appears possible that GAD activity is linked to neuronal activity through energy metabolism. GAD is not saturated by glutamate in synaptosomes or cortical slices, but there is no evidence that GABA synthesis in vivo is regulated physiologically by the availability of glutamate. GABA competitively inhibits GAD and converts holo- to apoGAD, but it is not clear if intracellular GABA levels are high enough to regulate GAD. There is no evidence of short-term regulation by second messengers. The syntheses of GAD65 and GAD67 proteins are regulated separately. GAD67 regulation is complex; it not only is present as apoGAD67, but the expression of GAD67 protein is regulated by two mechanisms: (a) by control of mRNA levels and (b) at the level of translation or protein stability. The latter mechanism appears to be mediated by intracellular GABA levels.  相似文献   

17.
P J Schechter  Y Tranier  J Grove 《Life sciences》1979,24(13):1173-1182
Two irreversible enzyme-activated GABA-transaminase inhibitors, gabaculine (5-amino cyclohex-1, 3-dienyl carboxylic acid) and an isomer, isogabaculine (3-amino cyclohex-1, 5-dienyl carboxylic acid), were investigated in mice for their effects on brain GABA metabolism and on seizures induced by a variety of stimuli. Biochemical and pharmacological activities of the two inhibitors were very similar. Both produce dose- and time-related, sustained inhibition of GABA-T activity and, to a lesser extent, of GAD activity and long-lasting increases in brain GABA concentrations. Both protect mice against audiogenic seizures and significantly decrease the frequency of seizures induced by isoniazid, thiosemicarbazide and pentylenetetrazol. They do not affect the frequency of seizures induced by strychnine, bicuculline or picrotoxin and do not alter the threshold to electroconvulsive shock. Although the effects of gabaculine and isogabaculine on brain GABA metabolism resemble those of other GABA-T inhibitors, important differences in pharmacological activities exist.  相似文献   

18.
BACKGROUND AND AIMS: Glutamic acid decarboxylase (GAD, EC 4.1.1.15) catalyses the conversion of glutamate to gamma-aminobutyric acid (GABA). The 65 kDa isoform, GAD65 is a potent autoantigen in type 1 diabetes, whereas GAD67 is not. A hybrid cDNA was created by fusing a human cDNA for amino acids 1-101 of GAD67 to a human cDNA for amino acids 96-585 of GAD65; the recombinant (r) protein was expressed in yeast and was shown to have equivalent immunoreactivity to mammalian brain GAD with diabetes sera. We here report on enzymatic and molecular properties of rGAD67/65. METHODS: Studies were performed on enzymatic activity of rGAD67/65 by production of 3H-GABA from 3H-glutamate, enzyme kinetics, binding to the enzyme cofactor pyridoxal phosphate (PLP), stability according to differences in pH, temperature and duration of storage, and antigenic reactivity with various GAD-specific antisera. RESULTS: The properties of rGAD67/65 were compared with published data for mammalian brain GAD (brackets). These included a specific enzyme activity of 22.7 (16.7) nKat, optimal pH for enzymatic activity 7.4 (6.8), K(m) of 1.3 (1.3) mM, efficient non-covalent binding to the cofactor PLP, and high autoantigenic potency. The stability of rGAD67/65 was optimal over 3 months at -80 degrees C, or in lyophilized form at -20 degrees C. CONCLUSIONS: Hybrid rGAD67/65 has enzymatic and other properties similar to those of the mixed isoforms of GAD in preparations from mammalian brain as described elsewhere, in addition to its previously described similar immunoreactivity.  相似文献   

19.
Abstract— γ-Vinyl GABA (4-amino-hex-5-enoic acid, RMI 71754) is a catalytic inhibitor of GABA-T in vitro. When given by a peripheral route to mice, it crosses the blood-brain barrier and induces a long-lasting, dose-dependent, irreversible inhibition of brain GABA transaminase (GABA-T). Glutamate decarboxylase (GAD) is only slightly affected even at the highest doses used. γ -Vinyl GABA has little or no effect on brain succinate semialdehyde dehydrogenase, aspartate transaminase and alanine transaminase activities. GABA-T inhibition is accompanied by a sustained dose-dependent increase of brain GABA concentration. From the rate of accumulation of GABA it was estimated that GABA turnover in brain was at least 6.5 μmol/g/h. Based on recovery of enzyme activity the half-life of GABA-T was found to be 3.4 days, that of GAD was estimated to be about 2.4 days. γ -Vinyl GABA should be valuable for manipulations of brain GABA metabolism.  相似文献   

20.
Abstract— Investigations of the respective effects of dietary vitamin B6 deficiency and 4-deoxypyridoxine (a vitamin B6 antagonist) on GABA metabolism in rat brain have been carried out. No convulsions were observed in rats subjected to either treatment. GABA levels were lowered by both treatments, the greatest diminutions being found with the dietary deficiency. Glutamic acid decarboxylase activity was reduced under both conditions, but the loss of activity in the B6 deficiency experiments could be attributed to cofactor depletion, whereas in the deoxypyridoxine experiments the loss of activity appears to be due to lower levels of available apoenzyme. The activity of GABA-transaminase was not affected by deoxypyridoxine treatment and only moderately reduced in the B.5 deficient animals.  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