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 46 毫秒
1.
The specific binding of [3H] (?)alprenolol to sites in frog erythrocyte membranes provides a tool for directly assessing ligand binding to adenylatecyclase coupled β-adrenergic receptors. Hill Plots of such binding data yield slopes (nH=“Hill Coefficients”) less than 1.0, suggesting that negatively cooperative interactions among the β-adrenergic receptors may occur. The existence of such negative cooperativity was confirmed by a direct kinetic method. The dissociation of receptor bound [3H] (?)alprenolol was studied under two conditions: 1) with dilution of the ligand-receptor complex sufficient to prevent rebinding of the dissociated tracer and 2) with this same dilution in the presence of excess unlabeled (?)alprenolol. If the sites are independent, the dissociation rates must be the same in both cases. However, the presence of (?)alprenolol increases the rate of [3H] (?)alprenolol dissociation, indicating that negatively cooperative interactions among the β-adrenergic receptor binding sites do occur.  相似文献   

2.
The N-terminal domain of HIV-1 glycoprotein 41?000 (FP; residues 1-23; AVGIGALFLGFLGAAGSTMGARSCONH2) participates in fusion processes underlying virus-cell infection. Here, we use physical techniques to study the secondary conformation of synthetic FP in aqueous, structure-promoting, lipid and biomembrane environments. Circular dichroism and conventional, 12C-Fourier transform infrared (FTIR) spectroscopy indicated the following α-helical levels for FP in 1-palmitoyl-2-oleoylphosphatidylglycerol (POPG) liposomes∼hexafluoroisopropanol (HFIP)>trifluoroethanol (TFE)>phosphate-buffered saline (PBS). 12C-FTIR spectra also showed disordered FP structures in these environments, along with substantial β-structures for FP in TFE or PBS. In further experiments designed to map secondary conformations to specific residues, isotope-enhanced FTIR spectroscopy was performed using a suite of FP peptides labeled with 13C-carbonyl at multiple sites. Combining these 13C-enhanced FTIR results with molecular simulations indicated the following model for FP in HFIP: α-helix (residues 3-16) and random and β-structures (residues 1-2 and residues 17-23). Additional 13C-FTIR analysis indicated a similar conformation for FP in POPG at low peptide loading, except that the α-helix extends over residues 1-16. At low peptide loading in either human erythrocyte ghosts or lipid extracts from ghosts, 13C-FTIR spectroscopy showed α-helical conformations for the central core of FP (residues 5-15); on the other hand, at high peptide loading in ghosts or lipid extracts, the central core of FP assumed an antiparallel β-structure. FP at low loading in ghosts probably inserts deeply as an α-helix into the hydrophobic membrane bilayer, while at higher loading FP primarily associates with ghosts as an aqueous-accessible, β-sheet. In future studies, 13C-FTIR spectroscopy may yield residue-specific conformations for other membrane-bound proteins or peptides, which have been difficult to analyze with more standard methodologies.  相似文献   

3.
Abstract

Vinblastine (VLB) and its derivatives have been used for clinical first-line drugs to treat various cancers. Due to the resistance and serious side effects from using VLB and its derivatives, there is a need to discover and develop novel VLB derivatives with high activity against cancer cells. In order to better discover and develop new VLB derivatives, we need to study the structural basis of VLB's anti-cancer cytotoxicity and the mechanism of its interaction with α,β-tubulins. Based on the crystal structure of α,β-microtubule complex protein, the molecular dynamics method including the sampling PMF method was used to study the variation of dissociation free energy (ΔG) of α,β-tubulins under different system conditions, and then from which to study the mechanism of the interaction between VLB and α,β-tubulins. The obtained results show that the dissociation of pure α,β-tubulins requires 197.8?kJ·mol?1 for ΔG. When the VLB molecule exists between the interface of α,β-tubulins, the dissociation ΔG of α,β-tubulins reaches 220.5?kJ·mol?1, which is greater than that of pure α,β-tubulin. The VLB molecule is formed by connecting a vindoline moiety (VM) molecule with a catharanthine moiety (CM) molecule through a carbon-carbon bond, which is a larger molecule. When the CM molecule exists in the middle of α,β-tubulin interface, the dissociation ΔG of α,β-tubulins is 46.2?kJ·mol?1, during which the CM moves with β-tubulin. When the VM molecule exists between the middle of α,β-tubulin interface, the dissociation ΔG of α,β-tubulins is 86.7?kJ·mol?1, during which it moves with α-tubulin. Therefore, the VLB molecule is like a double-sides tape to stick α-tubulin and β-tubulin together. The VLB molecule intervenes the dynamic equilibrium between dissociation and aggregation of α-tubulin and β-tubulin by a double-sides sticking mechanism to exert high activity with toxicity against cancer cell. Besides, our results demonstrate that VLB has its structural basis for anticancer cytotoxicity due to its two compositions composed of a CM molecule and a VM molecule although they have little toxicity against cancer cell alone.  相似文献   

4.
The presence of muscarinic receptors in sheep and rat pineals was detected by binding of [3H]quinuclidinyl benzilate ([3H]QNB), a potent and specific muscarinic antagonist. [3H]QNB binding to sheep pineal membrane resuspensions was saturable and reversible, with a rate constant for association at 37°C of 6×108M?1min?1 and a rate constant for dissociation of 1×10?2min?1. Kinetic and saturation experiments yielded an equilibrium dissociation constant of 13–18 pM and a concentration of binding sites equivalent to 1.1 pmol/g of original wet weight. This is only about 5% of the level of β-adrenergic receptors. Competition by a variety of cholinergic drugs confirmed the muscarinic nature of the binding sites. Experiments in rats failed to detect a significant decrease in pineal [3H]QNB binding following bilateral superior cervical ganglionectomy, suggesting that the binding sites are not localized exclusively on sympathetic terminals.  相似文献   

5.
Analyses of brain phospholipid fatty acid profiles reveal a selective deficiency and enrichment in eicosapentaenoic acid (EPA) and docosahexaenoic acid (DHA), respectively. In order to account for this difference in brain fatty acid levels, we hypothesized that EPA is more rapidly β-oxidized upon its entry into the brain. Wild-type C57BL/6 mice were perfused with either 14C-EPA or 14C-DHA via in situ cerebral perfusion for 40 s, followed by a bicarbonate buffer to wash out the residual radiolabeled polyunsaturated fatty acid (PUFA) in the capillaries. 14C-PUFA-perfused brains were extracted for chemical analyses of neutral lipid and phospholipid fatty acids. Based on the radioactivity in aqueous, total lipid, neutral lipid and phospholipid fractions, volume of distribution (VD, μl/g) was calculated. The VD between 14C-EPA- and 14C-DHA-perfused samples was not statistically different for total lipid, neutral lipids or total phospholipids. However, the VD of 14C-EPA in the aqueous fraction was 2.5 times higher than that of 14C-DHA (p=0.025), suggesting a more extensive β-oxidation than DHA. Furthermore, radiolabeled palmitoleic acid, a fatty acid that can be synthesized de novo, was detected in brain phospholipids from 14C-EPA but not from 14C-DHA-perfused mice suggesting that β-oxidation products of EPA were recycled into endogenous fatty acid biosynthetic pathways. These findings suggest that low levels of EPA in brain phospholipids compared to DHA may be the result of its rapid β-oxidation upon uptake by the brain.  相似文献   

6.
The interaction with model membranes of a peptide, EqtII1–32, corresponding to the N‐terminal region of the pore‐forming toxin equinatoxin II (EqtII) has been studied using solid‐state NMR and molecular dynamics (MD) simulations. The distances between specifically labeled nuclei in [19F‐para]Phe16‐[1‐13C]Leu19 and [19F‐para]Phe16‐[15N]Leu23 analogs of EqtII1–32 measured by REDOR in lyophilized peptide were in agreement with published crystal and solution structures. However, in both DMPC and mixed DMPC:SM membrane environments, significant changes in the distances between the labeled amino acid pairs were observed, suggesting changes in helical content around the experimentally studied region, 16–23, in the presence of bilayers. 19F‐31P REDOR experiments indicated that the aromatic ring of Phe16 is in contact with lipid headgroups in both membrane environments. For the DMPC:SM mixed bilayers, a closer interaction between Phe16 side chains and lipid headgroups was observed, but an increase in distances was observed for both labeled amino acid pairs compared with those measured for EqtII1–32 in pure DMPC bilayers. The observed differences between DMPC and DMPC:SM bilayers may be due to the greater affinity of EqtII for the latter. MD simulations of EqtII1–32 in water, on a pure DMPC bilayer, and on a mixed DMPC:SM bilayer indicate significant peptide secondary structural differences in the different environments, with the DMPC‐bound peptide adopting helical formations at residues 16–24, whereas the DMPC:SM‐bound peptide exhibits a longer helical stretch, which may contribute to its enhanced activity against PC:SM compared with pure PC bilayers. Proteins 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

7.
The nucleotide sequence from the 5′ terminus inward of one third of mouse α- and βmaj-globin messenger RNAs has been established. In addition, using 5′ 32P end-labeled mRNAs as substrates and S1 and T1 nucleases as probes for single-stranded regions, the secondary structures of mouse and rabbit α- and β-globin mRNAs have been analyzed. Our results indicate that the AUG initiator codon in both mouse and rabbit β-globin mRNA is quite susceptible to cleavage with S1 and T1 nucleases, suggesting that it resides in a single-stranded exposed region. In contrast, the initiator AUG in the α-globin mRNA of both species is inaccessible to cleavage, indicating that it is either buried by tertiary structure or is base-paired. Since the rate of initiation of protein synthesis with β-globin mRNA in rabbit reticulocyte is 30–40% faster than for α-globin mRNA, these results imply a possible correlation between the differential rates of initiation with these two mRNAs and the accessibility of the respective AUG initiator codons.  相似文献   

8.
(?) [3H]Dihydroalprenolol, a potent competitive β-adrenergic antagonist can be used to directly study β-adrenergic receptors by ligand binding techniques in an intact cell system, the frog erythrocyte. At 37°, binding reached equilibrium within 1 minute. Upon addition of excess unlabeled propranolol, complete dissociation of receptor bound ligand occurred within 1 minute. The characteristics of (?) [3Hdihydroalprenolol binding to β-adrenergic receptors in intact cells were quite similar to those previously demonstrated with isolated membrane fractions. The equilibrium dissociation constant for (?) [3H]dihydroalprenolol was 1.5 nM. Order of potency of agonists and antagonists in competing for the binding sites was appropriate for the β-adrenergic receptor as was the stereospecificity of binding ((?) isomers more potent than (+) isomers). Saturation studies with these intact cells indicated 1700 binding sites/cell in excellent agreement with the number previously estimated from membrane studies. Preincubation of cells with 10?5M isoproterenol produced a 36% fall in number of β-adrenergic receptors. It is concluded that (?) [3H]dihydroalprenolol can be used to directly study the properties and regulation of β-adrenergic receptors in intact cell as well as broken cell preparations.  相似文献   

9.
Shih PC  Li GC  Yang KJ  Chen W  Tzou DL 《Steroids》2011,76(6):558-563
Solid-state 1H/13C cross-polarization/magic angle spinning (CP/MAS) NMR spectroscopy has been applied to two steroid compounds: dehydroepiandrosterone (DHEA) and spironolactone (SPI), to analyze their conformations at the atomic level. In the absence of lipid, the high-resolution 13C CP/MAS NMR signals of DHEA and SPI in a powder form reveal multiple patterns, with splittings of 30-160 Hz, indicating the existence of multiple conformations. In the mimic lipid environment formed by mixing 1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC) and 1,2-diheptanoyl-sn-glycero-3-phosphocholine (DHPC) in a molar ratio 3:1, the resulting DHEA and SPI spectra revealed mostly singlet patterns, suggesting that these steroids undergo a conformational change leading to a specific conformation in the lipid environment. Evidence from chemical shift isotropy and anisotropy analysis indicates that DHEA might adopt conformations subtly different from that seen in solution and in the powder form. In conclusion, we demonstrate by solid-state NMR that the structures of DHEA and SPI may adopt slightly different conformations in different chemical environments.  相似文献   

10.
The binding of 99mTc to negatively-charged and neutral unilamellar lipid vesicles was investigated in the absence and presence of the ligand diethylenetriaminepentaacetic acid (DTPA) covalently attached to the headgroup of phosphatidylethanolamine at the surface of the membrane. Even in the absence of DTPA on the membrane surface, 99mTc reduced by Sn bound to the membrane surface but rapidly dissociated from the vesicles in the presence of plasma in vitro. When DTPA was present on the membrane surface, dissociation of 99mTc from the vesicle surface in plasma was much reduced. The dissociation of 99mTc from the surface of negatively-charged vesicles was less than for neutral vesicles in the absence of ligand but was markedly greater than for vesicles containing the ligand DTPA, suggesting that the binding of 99mTc to vesicles with surface-attached DTPA could not be explained solely on the basis of the negative charge provided by the DTPA. In vitro experiments using 14C-labeled lipids as well as in vivo imaging studies indicated that dissociation of 99mTc from the surface of the vesicle did not arise predominantly because of lipid exchange with plasma components or due to cleavage of Tc-DTPA from the vesicle surface. For vesicles with surface-attached DTPA, 99mTc dissociation from the vesicle surface in plasma was further reduced by addition of the antioxidant ascorbate.  相似文献   

11.
Two-dimensional 1H/31P dipolar heteronuclear correlation (HETCOR) magic-angle spinning nuclear magnetic resonance (NMR) is used to investigate the correlation of the lipid headgroup with various intra- and intermolecular proton environments. Cross-polarization NMR techniques involving 31P have not been previously pursued to a great extent in lipid bilayers due to the long 1H-31P distances and high degree of headgroup mobility that averages the dipolar coupling in the liquid crystalline phase. The results presented herein show that this approach is very promising and yields information not readily available with other experimental methods. Of particular interest is the detection of a unique lipid backbone-water intermolecular interaction in egg sphingomyelin (SM) that is not observed in lipids with glycerol backbones like phosphatidylcholines. This backbone-water interaction in SM is probed when a mixing period allowing magnetization exchange between different 1H environments via the nuclear Overhauser effect (NOE) is included in the NMR pulse sequence. The molecular information provided by these 1H/31P dipolar HETCOR experiments with NOE mixing differ from those previously obtained by conventional NOE spectroscopy and heteronuclear NOE spectroscopy NMR experiments. In addition, two-dimensional 1H/13C INEPT HETCOR experiments with NOE mixing support the 1H/31P dipolar HETCOR results and confirm the presence of a H2O environment that has nonvanishing dipolar interactions with the SM backbone.  相似文献   

12.
Abstract

The association (k+1) and dissociation (k-1) rate constants, and the equilibrium thermodynamic binding parameters (ΔG°, ΔH° and ΔS°) of the β-adrenergic ligand [125Iodo]cyanopindolol (ICYP) were studied in a crude lung membrane preparation of rats of different ages. There was no difference in k+1-values for the different age groups, while the k-1-values were in all cases difficult to measure: almost no dissociation of ICYP from its binding site occurs. The thermodynamic properties were not affected by age. It is concluded that, in these experimental conditions, age has no effect on the kinetic parameters of the binding of ICYP to the β-adrenoceptors in rat lung.  相似文献   

13.
The tellurium oxyanion TeO32− has been used in the treatment of infectious diseases caused by mycobacteria. However, many pathogenic bacteria show tellurite resistance. Several tellurite resistance genes have been identified, and these genes mediate responses to diverse extracellular stimuli, but the mechanisms underlying their functions are unknown. To shed light on the function of KP-TerD, a 20.5 -kDa tellurite resistance protein from a plasmid of Klebsiella pneumoniae, we have determined its three-dimensional structure in solution using NMR spectroscopy. KP-TerD contains a β-sandwich formed by two five-stranded β-sheets and six short helices. The structure exhibits two negative clusters in loop regions on the top of the sandwich, suggesting that KP-TerD may bind metal ions. Indeed, thermal denaturation experiments monitored by circular dichroism and NMR studies reveal that KP-TerD binds Ca2+. Inductively coupled plasma-optical emission spectroscopy shows that the binding ratio of KP-TerD to Ca2+ is 1:2. EDTA (ethylenediaminetetraacetic acid) titrations of Ca2+-saturated KP-TerD monitored by one-dimensional NMR yield estimated dissociation constants of 18  and 200 nM for the two Ca2+-binding sites of KP-TerD. NMR structures incorporating two Ca2+ ions define a novel bipartite Ca2+-binding motif that is predicted to be highly conserved in TerD proteins. Moreover, these Ca2+-binding sites are also predicted to be present in two additional tellurite resistance proteins, TerE and TerZ. These results suggest that some form of Ca2+ signaling plays a crucial role in tellurite resistance and in other responses of bacteria to multiple external stimuli that depend on the Ter genes.  相似文献   

14.
Noradrenaline (0.1–5 μM, in the presence of 5 μM propranolol to block β-receptors), ATP (100 μM) and angiotensin II (0.1 μM), which are thought to increase cytosolic Ca2+ concentration by mobilizing Ca2+ from internal stores, increased the lipid fluidity as measured by diphenylhexatriene fluorescence polarization in plasma membranes isolated from rat liver. The effect of noradrenaline was dose-dependent and blocked by the α-antagonists phenoxybenzamine (50 μM) and phentolamine (1 μM). The response to a maximal dose of noradrenaline (5 μM) and that to ATP (100 μM) were not cumulative, suggesting that both agents use a common mechanism to alter the membrane lipid fluidity. In contrast, the addition of noradrenaline (5 μM) along with the foreign amphiphile Na+-oleate (1–30 μM) resulted in an increase in membrane lipid fluidity which was equivalent to the sum of individual responses to the two agents. In the absence of Mg2+, reducing free Ca2+ concentration by adding EGTA increased membrane lipid fluidity and abolished the effect of noradrenaline, suggesting that Ca2+ is involved in the mechanism by which the hormone exerts its effect on plasma membranes. Noradrenaline (5 μM) and angiotensin II (0.1 μM) also promoted a small release of 45Ca2+ (16 pmol/mg membrane proteins) from prelabelled plasma membranes. The effect of noradrenaline was suppressed by the α-antagonist phentolamine (5 μM). It is proposed that noradrenaline, via α-adrenergic receptors and other Ca2+-mobilizing hormones, increases membrane lipid fluidity by displacing a small pool of Ca2+ bound to phospholipids, removing thus the mechanical constraints brought about by this ion.  相似文献   

15.
It has been suggested that cholesterol may modulate amyloid-β (Aβ) formation, a causative factor of Alzheimer’s disease (AD), by regulating distribution of the three key proteins in the pathogenesis of AD (β-amyloid precursor protein (APP), β-secretase (BACE1) and/or presenilin 1 (PS1)) within lipid rafts. In this work we tested whether cholesterol accumulation upon NPC1 dysfunction, which causes Niemann Pick type C disease (NPC), causes increased partitioning of APP into lipid rafts leading to increased CTF/Aβ formation in these cholesterol-rich membrane microdomains. To test this we used CHO NPC1−/− cells (NPC cells) and parental CHOwt cells. By sucrose density gradient centrifugation we observed a shift in fl-APP/CTF compartmentalization into lipid raft fractions upon cholesterol accumulation in NPC vs. wt cells. Furthermore, γ-secretase inhibitor treatment significantly increased fl-APP/CTF distribution in raft fractions in NPC vs. wt cells, suggesting that upon cholesterol accumulation in NPC1-null cells increased formation of APP-CTF and its increased processing towards Aβ occurs in lipid rafts. Our results support that cholesterol overload, such as in NPC disease, leads to increased partitioning of APP/CTF into lipid rafts resulting in increased amyloidogenic processing of APP in these cholesterol-rich membranes. This work adds to the mechanism of the cholesterol-effect on APP processing and the pathogenesis of Alzheimer’s disease and supports the role of lipid rafts in these processes.  相似文献   

16.
R Blomstrand  L Kager  O Lantto 《Life sciences》1973,13(8):1131-1141
In order to elucidate the mechanisms involved in the acute ethanol-induced liver triglyceride accumulation, the oxidation, esterification and β-keto acid formation have been studied in rat and human liver slices after incubation with albumin bound, long chain fatty acids (palmitic. oleic and linoleic acids).The addition of alcohol to rat and human liver slices depressed the formation of 14CO2 from palmitic acid-1-14C, oleic acid-1-14C and linoleic acid-1-14C. The esterification to triglycerides and phospholipids was increased and the formation of β-keto acids was decreased by alcohol.Addition of 4-methylpyrazole, an inhibitor of liver alcohol dehydrogenase, almost prevented the alcohol effect on the lipid metabolism of the liver slices. The oxidation of alcohol is thus obligatory for the decreased β-oxidation of fatty acids, the increased esterification and for the decreased formation of β-keto acids. The results suggest that ethanol oxidation inhibits β-oxidation of fatty acids and that this primary effect leads to accumulation of liver triglycerides by increased esterification of plasma free fatty acids.  相似文献   

17.
The peptide-lipid interaction of a β-hairpin antimicrobial peptide tachyplesin-1 (TP-1) and its linear derivatives are investigated to gain insight into the mechanism of antimicrobial activity. 31P and 2H NMR spectra of uniaxially aligned lipid bilayers of varying compositions and peptide concentrations are measured to determine the peptide-induced orientational disorder and the selectivity of membrane disruption by tachyplesin. The disulfide-linked TP-1 does not cause any disorder to the neutral POPC and POPC/cholesterol membranes but induces both micellization and random orientation distribution to the anionic POPE/POPG membranes above a peptide concentration of 2%. In comparison, the anionic POPC/POPG bilayer is completely unaffected by TP-1 binding, suggesting that TP-1 induces negative curvature strain to the membrane as a mechanism of its action. Removal of the disulfide bonds by substitution of Cys residues with Tyr and Ala abolishes the micellization of POPE/POPG bilayers but retains the orientation randomization of both POPC/POPG and POPE/POPG bilayers. Thus, linear tachyplesin derivatives have membrane disruptive abilities but use different mechanisms from the wild-type peptide. The different lipid-peptide interactions between TP-1 and other β-hairpin antimicrobial peptides are discussed in terms of their molecular structure.  相似文献   

18.
Calf brain membranes have previously been shown to enzymatically transfer N-acetyl[14C]glucosamine from UDP-N-acetyl[14C]glucosamine into N-acetyl[14C]glucosami-nylpyrophosphoryldolichol, N,N′-diacetyl[14C]chitobiosylpyrophosphoryldolichol and a minor labeled product with the chemical and chromatographic properties of a [14C]trisaccharide lipid (Waechter, C. J., and Harford, J. B. (1977) Arch. Biochem. Biophys.181, 185–198). This paper demonstrates that incubating calf brain membranes containing endogenous, prelabeled N-acetyl[14C]glucosaminyl lipids with unlabeled GDP-mannose enhances the formation of the [14C]trisaccharide lipid. The intact [14C]trisaccharide lipid behaves like a dolichol-bound trisaccharide, in which the glycosyl group is linked via a pyrophosphate bridge, when chromatographed on SG-81 paper or DEAE-cellulose. Mild acid treatment releases a water-soluble product that comigrates with authentic β-Man-(1→4)-β-GlcNAc(1→4)-GlcNAc. The free [14C]trisaccharide is converted to N,N′-diacetyl[14C]chitobiose by incubation with a highly purified β-mannosidase. These findings indicate that the trisaccharide lipid formed by calf brain membranes is β-mannosyl-N,N′-diacetylchito-biosylpyrophosphoryldolichol. The two glycosyltransferases responsible for the enzymatic conversion of the N-acetylglucosaminyl lipid to the trisaccharide lipid have been studied using exogenous, purified [14C]glycolipid substrates. Calf brain membranes enzymatically transfer N-acetylglucosamine from UDP-N-acetylglucosamine to exogenous N-acetyl[14C] glucosaminylpyrophosphoryldolichol to form [14C]disaccharide lipid. The biosynthesis of [14C]disaccharide lipid is stimulated by unlabeled UDP-N-acetylglucosamine under conditions that inhibit N-acetylglucosaminylpyrophosphoryldolichol synthesis. Unlike the formation of N-acetylglucosaminylpyrophosphoryldolichol the enzymatic addition of the second N-acetylglucosamine residue is not inhibited by tunicamycin. Exogenous purified [14C] disaccharide lipid is enzymatically mannosylated by calf brain membranes to form the [14C] trisaccharide lipid. The formation of the [14C]trisaccharide lipid from exogenous [14C] disaccharide lipid is stimulated by unlabeled GDP-mannose and Mg2+, and inhibited by EDTA. Exogenous dolichyl monophosphate is also inhibitory. These results strongly suggest that the calf brain mannosyltransferase involved in the synthesis of the trisaccharide lipid requires a divalent cation and utilizes GDP-mannose, not mannosylphosphoryldolichol, as the direct mannosyl donor.  相似文献   

19.
Ganglioside GM1 beta-galactosidase: studies in human liver and brain   总被引:10,自引:0,他引:10  
A microcolumn assay for ganglioside GM1 β-galactosidase (EC 3.2.1.23) has been developed using GM1 tritiated exclusively in the terminal galactose residue. The reaction is stimulated up to 100-fold by anionic and cationic detergents; this stimulation is inhibited by neutral detergents. 4-Methylumbelliferyl β-d-galactopyranoside is hydrolyzed about seven times more rapidly than GM1 in human brain (gray matter) and liver. Agarose gel filtration separated two forms of GM1 β-galactosidase in both brain and liver. The major form (ganglioside GM1 β-galactosidase A) had a molecular weight of 60–70 × 103 and the minor form (ganglioside GM1 β-galactosidase B) 600–800 × 103. The liver and brain GM1 β-galactosidases and 4-methylumbelliferyl β-galactosidase A cochromatographed on fractionation. The two forms of the enzyme in liver isolated by gel filtration corresponded to the two major forms found on starch gel electrophoresis and were converted to electrophoretically slower-moving forms after treatment with neuraminidase (EC 3.2.1.8, Cl. perfringens) suggesting that both are sialylated glycoproteins. The activity of GM1 β-galactosidase in the brain and liver tissue of patients with GM1 gangliosidosis Types I and II was less than 2% of control values. The mutation in each GM1 gangliosidosis appears to result in a severe reduction of activity of two ganglioside GM1 β-galactosidases.  相似文献   

20.
《Cellular signalling》2014,26(2):295-305
Alcohol-induced liver injury is the most common liver disease in which fatty acid metabolism is altered. It is thought that altered NAD+/NADH redox potential by alcohol in the liver causes fatty liver by inhibiting fatty acid oxidation and the activity of tricarboxylic acid cycle reactions. β-Lapachone (βL), a naturally occurring quinone, has been shown to stimulate fatty acid oxidation in an obese mouse model by activating adenosine monophosphate-activated protein kinase (AMPK). In this report, we clearly show that βL reduced alcohol-induced hepatic steatosis and induced fatty acid oxidizing capacity in ethanol-fed rats. βL treatment markedly decreased hepatic lipids while serum levels of lipids and lipoproteins were increased in rats fed ethanol-containing liquid diets with βL administration. Furthermore, inhibition of lipolysis, enhancement of lipid mobilization to mitochondria and upregulation of mitochondrial β-oxidation activity in the soleus muscle were observed in ethanol/βL-treated animals compared to the ethanol-fed rats. In addition, the activity of alcohol dehydrogenase, but not aldehyde dehydrogenase, was significantly increased in rats fed βL diets. βL-mediated modulation of NAD+/NADH ratio led to the activation of AMPK signaling in these animals. Conclusion: Our results suggest that improvement of fatty liver by βL administration is mediated by the upregulation of apoB100 synthesis and lipid mobilization from the liver as well as the direct involvement of βL on NAD+/NADH ratio changes, resulting in the activation of AMPK signaling and PPARα-mediated β-oxidation. Therefore, βL-mediated alteration of NAD+/NADH redox potential may be of potential therapeutic benefit in the clinical setting.  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