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 15 毫秒
1.
High saturated fat (HF-S) diets increase intramyocellular lipid, an effect ameliorated by omega-3 fatty acids in vitro and in vivo, though little is known about sex- and muscle fiber type-specific effects. We compared effects of standard chow, HF-S, and 7.5% HF-S replaced with fish oil (HF-FO) diets on the metabolic profile and lipid metabolism gene and protein content in red (soleus) and white (extensor digitorum longus) muscles of male and female C57BL/6 mice (n = 9-12/group). Weight gain was similar in HF-S- and HF-FO-fed groups. HF-S feeding increased mesenteric fat mass and lipid marker, Oil Red O, in red and mixed muscle; HF-FO increased interscapular brown fat mass. Compared to chow, HF-S and HF-FO increased expression of genes regulating triacylglycerol synthesis and fatty acid transport, HF-S suppressed genes and proteins regulating fatty acid oxidation, whereas HF-FO increased oxidative genes, proteins and enzymes and lipolytic gene content, whilst suppressing lipogenic genes. In comparison to HF-S, HF-FO further increased fat transporters, markers of fatty acid oxidation and mitochondrial content, and reduced lipogenic genes. No diet-by-sex interactions were observed. Neither diet influenced fiber type composition. However, some interactions between muscle type and diet were observed. HF-S induced changes in triacylglycerol synthesis and lipogenic genes in red, but not white, muscle, and mitochondrial biogenesis and oxidative genes were suppressed by HF-S and increased by HF-FO in red muscle only. In conclusion, HF-S feeding promotes lipid storage in red muscle, an effect abrogated by the fish oil, which increases mediators of lipolysis, oxidation and thermogenesis while inhibiting lipogenic genes. Greater storage and synthesis, and lower oxidative genes in red, but not white, muscle likely contribute to lipid accretion encountered in red muscle. Despite several gender-dimorphic genes, both sexes exhibited a similar HF-S-induced metabolic and gene expression profile; likewise fish oil was similarly protective in both sexes.  相似文献   

2.
During anaerobic growth of Escherichia coli, pyruvate formate-lyase (PFL) and lactate dehydrogenase (LDH) channel pyruvate toward a mixture of fermentation products. We have introduced a third branch at the pyruvate node in a mutant of E. coli with a mutation in pyruvate dehydrogenase (PDH*) that renders the enzyme less sensitive to inhibition by NADH. The key starting enzymes of the three branches at the pyruvate node in such a mutant, PDH*, PFL, and LDH, have different metabolic potentials and kinetic properties. In such a mutant (strain QZ2), pyruvate flux through LDH was about 30%, with the remainder of the flux occurring through PFL, indicating that LDH is a preferred route of pyruvate conversion over PDH*. In a pfl mutant (strain YK167) with both PDH* and LDH activities, flux through PDH* was about 33% of the total, confirming the ability of LDH to outcompete the PDH pathway for pyruvate in vivo. Only in the absence of LDH (strain QZ3) was pyruvate carbon equally distributed between the PDH* and PFL pathways. A pfl mutant with LDH and PDH* activities, as well as a pfl ldh double mutant with PDH* activity, had a surprisingly low cell yield per mole of ATP (YATP) (about 7.0 g of cells per mol of ATP) compared to 10.9 g of cells per mol of ATP for the wild type. The lower YATP suggests the operation of a futile energy cycle in the absence of PFL in this strain. An understanding of the controls at the pyruvate node during anaerobic growth is expected to provide unique insights into rational metabolic engineering of E. coli and related bacteria for the production of various biobased products at high rates and yields.In Escherichia coli as well as in other aerobic organisms, sugars such as glucose are metabolized in two separate steps: glycolysis, which converts glucose to pyruvate, and tricarboxylic acid (TCA) cycle enzymes, which oxidize acetyl coenzyme A (acetyl-CoA) to CO2 (5, 9). The pyruvate dehydrogenase complex (PDH) connects the glycolytic reactions to TCA cycle enzymes by catalyzing the production of acetyl-CoA from pyruvate. Because of its unique central role in metabolism, PDH is regulated at both the genetic and the biochemical level (7, 12, 27, 33, 34). The NADH generated during the complete oxidation of sugar is reoxidized to NAD+ by O2 through the respiratory electron transport pathway with accompanying energy production (11). Optimum coupling of these enzyme reactions helps to maintain the internal ratios of [NADH] to [NAD+] (redox balance) and of [ATP] to [ADP] plus [AMP] in order to support growth at the highest rate.The absence of O2 or another external electron acceptor during the growth of E. coli (anaerobic conditions) forces the bacterium to minimize the contribution of the TCA cycle enzymes to biosynthesis from catabolism (4, 14). Under these conditions, pyruvate or acetyl-CoA derived from pyruvate serves as the electron acceptor (reduced to lactate or ethanol, respectively) to maintain the redox balance. The enzymes responsible for redox balance in anaerobic E. coli are pyruvate formate-lyase (PFL), lactate dehydrogenase (LDH), and alcohol/aldehyde dehydrogenase (adhE; ADH-E). The main products of the fermentation of E. coli are a mixture of organic acids, such as acetate, lactate, and formate, in addition to ethanol (2, 4). Succinate, derived from phosphoenolpyruvate (PEP), is a minor product of fermentation and normally accounts for less than 5% of the total products produced from glucose by the culture.Anaerobic growth of E. coli, compared to aerobic growth, is also limited by energy, leading to an increase in glycolytic flux (19). The conversion of pyruvate to acetate and ethanol yields an additional ATP per glucose, suggesting that this would be the preferred route for pyruvate oxidation during anaerobic growth. This is accomplished by the PFL-dependent production of acetyl-CoA and further conversion to acetate (Fig. (Fig.1).1). This preference for PFL has been demonstrated `with several bacteria under carbon limitation conditions imposed either in a chemostat or in the presence of a poor carbon source (10, 20, 23). This additional ATP also elevates the ATP yield per glucose to 3, with an increase in the growth rate, and has been shown to be essential for the anaerobic growth of E. coli in xylose-mineral salts medium (13). The absence of this third ATP in a pfl mutant has been reported to increase glycolytic flux to lactate to compensate for this decrease in ATP yield per glucose (39). However, the flow of pyruvate carbon to acetate is tempered by the need to maintain redox balance, and this is achieved by the conversion of a second acetyl-CoA to ethanol by ADH-E. Under conditions of energy excess due to a declining growth rate, lactate production is expected to support redox balance maintenance without the additional ATP from the PFL-ADH-E pathway (Fig. (Fig.1).1). The production of this mixture of products in an appropriate ratio helps to maintain the redox balance under anaerobic conditions while also maximizing the ATP yield per glucose to support high growth rates and cell yields.Open in a separate windowFIG. 1.Anaerobic metabolic pathways of E. coli carrying the lpd101 mutation (PDH*).No PDH-based fermentation reaction to ethanol that can also help maintain cellular redox balance in an anaerobic cell has evolved in E. coli or other closely related bacteria. PDH activity is inhibited by NADH, normally found at higher levels in anaerobically growing cultures than in aerobic cultures (12, 18, 34, 35). Based on genome sequences available in GenBank, the genes encoding the components of PDH are not found in strictly anaerobic bacteria.We have recently described a mutation (lpd101) in the dihydrolipoamide dehydrogenase (LPD) of the PDH that allowed the enzyme to function in anaerobic cells (designated PDH* here) (17, 18). With this altered PDH*, an anaerobic cell can have three different pathways for pyruvate metabolism (Fig. (Fig.1).1). The three main enzymes that utilize pyruvate as a substrate, PDH*, PFL, and LDH, have different apparent Km values for pyruvate (0.4, 2.0, and 7.2 mM, respectively) (1, 18, 37, 41). PDH requires NAD+ for activity (apparent Km, 0.07 mM), while LDH is dependent on NADH (apparent Km, 0.2 mM) as the second substrate (18, 37).The PDH* serves as the first enzyme in a pathway that oxidatively decarboxylates pyruvate to acetyl-CoA and NADH, followed by reduction of the acetyl-CoA by alcohol dehydrogenase to ethanol in a two-step process using 2 NADHs (Fig. (Fig.1).1). The NADH produced during the conversion of glucose to acetyl-CoA dictates that the acetyl-CoA generated by PDH be used for redox balance (ethanol) and not for ATP generation (acetate), unless some of the NADH is used for biosynthesis by the growing cell (17). PDH* and LDH serve essentially the same physiological role in the cell, oxidizing NADH to support continued operation of glycolysis, although it is not readily apparent with PDH*. Although PDH* contributes to an increase in NADH pool, the redox balance is still maintained by coupling PDH* to NADH-dependent reduction of acetyl-CoA to ethanol by ADH-E (Fig. (Fig.1).1). This potential competition between LDH and PDH has been eliminated in the wild type through inhibition of the activity of PDH by NADH (12, 18, 32). However, the in vivo role of PDH* in a mutant that has all three pathways has not been investigated, since the flow of pyruvate through any of the three reactions during growth and postgrowth fermentation of sugars to products is expected to be dependent on the redox state, the ATP requirement, and other physiological conditions of the anaerobic cell. Using a combination of metabolic flux analysis and mutations in one or more of the genes encoding these enzymes, we have evaluated the flow of pyruvate carbon among the three potential pathways. The results are presented in this communication.  相似文献   

3.
The dorsal root ganglion (DRG) is a highly vulnerable site in diabetic neuropathy. Under diabetic conditions, the DRG is subjected to tissue ischemia or lower ambient oxygen tension that leads to aberrant metabolic functions. Metabolic dysfunctions have been documented to play a crucial role in the pathogenesis of diverse pain hypersensitivities. However, the contribution of diabetes-induced metabolic dysfunctions in the DRG to the pathogenesis of painful diabetic neuropathy remains ill-explored. In this study, we report that pyruvate dehydrogenase kinases (PDK2 and PDK4), key regulatory enzymes in glucose metabolism, mediate glycolytic metabolic shift in the DRG leading to painful diabetic neuropathy. Streptozotocin-induced diabetes substantially enhanced the expression and activity of the PDKs in the DRG, and the genetic ablation of Pdk2 and Pdk4 attenuated the hyperglycemia-induced pain hypersensitivity. Mechanistically, Pdk2/4 deficiency inhibited the diabetes-induced lactate surge, expression of pain-related ion channels, activation of satellite glial cells, and infiltration of macrophages in the DRG, in addition to reducing central sensitization and neuroinflammation hallmarks in the spinal cord, which probably accounts for the attenuated pain hypersensitivity. Pdk2/4-deficient mice were partly resistant to the diabetes-induced loss of peripheral nerve structure and function. Furthermore, in the experiments using DRG neuron cultures, lactic acid treatment enhanced the expression of the ion channels and compromised cell viability. Finally, the pharmacological inhibition of DRG PDKs or lactic acid production substantially attenuated diabetes-induced pain hypersensitivity. Taken together, PDK2/4 induction and the subsequent lactate surge induce the metabolic shift in the diabetic DRG, thereby contributing to the pathogenesis of painful diabetic neuropathy.  相似文献   

4.
The mitochondrial pyruvate dehydrogenase complex (PDC) is down-regulated by phosphorylation catalyzed by pyruvate dehydrogenase kinase (PDK) isoforms 1–4. Overexpression of PDK isoforms and therefore reduced PDC activity prevails in cancer and diabetes. In the present study, we investigated the role of the invariant C-terminal DW-motif in inhibition of human PDK2 by dichloroacetate (DCA). Substitutions were made in the DW-motif (Asp-382 and Trp-383) and its interacting residues (Tyr-145 and Arg-149) in the other subunit of PDK2 homodimer. Single and double mutants show 20–60% residual activities that are not stimulated by the PDC core. The R149A and Y145F/R149A mutants show drastic increases in apparent IC50 values for DCA, whereas binding affinities for DCA are comparable with wild-type PDK2. Both R149A and Y145F variants exhibit increased similar affinities for ADP and ATP, mimicking the effects of DCA. The R149A and the DW-motif mutations (D382A/W383A) forestall binding of the lipoyl domain of PDC to these mutants, analogous to wild-type PDK2 in the presence of DCA and ADP. In contrast, the binding of a dihydrolipoamide mimetic AZD7545 is largely unaffected in these PDK2 variants. Our results illuminate the pivotal role of the DW-motif in mediating communications between the DCA-, the nucleotide-, and the lipoyl domain-binding sites. This signaling network locks PDK2 in the inactive closed conformation, which is in equilibrium with the active open conformation without DCA and ADP. These results implicate the DW-motif anchoring site as a drug target for the inhibition of aberrant PDK activity in cancer and diabetes.  相似文献   

5.
The Escherichia coli pyruvate dehydrogenase multienzyme complex contains multiple copies of three enzymatic components, E1p, E2p, and E3, that sequentially carry out distinct steps in the overall reaction converting pyruvate to acetyl-CoA. Efficient functioning requires the enzymatic components to assemble into a large complex, the integrity of which is maintained by tethering of the displaced, peripheral E1p and E3 components to the E2p core through non-covalent binding. We here report the crystal structure of a subcomplex between E1p and an E2p didomain containing a hybrid lipoyl domain along with the peripheral subunit-binding domain responsible for tethering to the core. In the structure, a region at the N terminus of each subunit in the E1p homodimer previously unseen due to crystallographic disorder was observed, revealing a new folding motif involved in E1p-E2p didomain interactions, and an additional, unexpected, flexibility was discovered in the E1p-E2p didomain subcomplex, both of which probably have consequences in the overall multienzyme complex assembly. This represents the first structure of an E1p-E2p didomain subcomplex involving a homodimeric E1p, and the results may be applicable to a large range of complexes with homodimeric E1 components. Results of HD exchange mass spectrometric experiments using the intact, wild type 3-lipoyl E2p and E1p are consistent with the crystallographic data obtained from the E1p-E2p didomain subcomplex as well as with other biochemical and NMR data reported from our groups, confirming that our findings are applicable to the entire E1p-E2p assembly.  相似文献   

6.
Abstract: Activity of the pyruvate dehydrogenase complex (PDHC) was measured in seven brain regions of themale rat at various times during the postnatal period usingan arylamine acetyltransferase coupled assay. Three daysafter birth, PDHC activity was found to be < 15% ofadult values in all brain regions with the exception of hypothalamus and medulla-pons (30% of adult values ineach case). Activity of the enzyme complex in these latterregions attained adult levels by 21 days postnatally, some 5-15 days ahead of that found in cerebral cortex, striatum, hippocampus, and cerebellum. Such differences in PDHC maturation reflect the greater degree of earlymaturity of the phylogenetically older brain structures. Cerebellar PDHC developed more slowly than in otherbrain regions to attain only 40% of adult levels by thetime of weaning. The pattern of maturation of cerebellarPDHC is paralleled by increased incorporation of glucoseinto cerebral amino acids and by the pattern of develop-ment of parallel fiber synaptogenesis. These findings sug-gest that PDHC may play a key role in the regional de-velopment of metabolic compartmentation and the asso-ciated maturation of cerebral function in the rat.  相似文献   

7.
On the basis of the homodimeric X-ray structure of dihydrolipoamide dehydrogenase from Azotobacter vinelandii we demonstrate by protein modeling techniques that two dimeric units of this enzyme can associate to a tetrameric structure with intense contacts between the building blocks. Complementary structures of the respective other unit in the tetramer contribute to the active sites. The coenzyme FAD becomes shielded from the environment, thus its binding is stabilized. By energy minimization techniques binding energies and RMS-values were computed and the contact areas between the building blocks were determined to quantify the interaction. In the cell tetramerization of dihydrolipoamide dehydrogenase will be realized upon its incorporation as an enzyme component into the pyruvate dehydrogenase multienzyme complex and will have consequences for the structure and subunit stoichiometry of the complex. Especially, the multiplicity of the three enzyme components, i.e. pyruvate dehydrogenase, dihydrolipoamide acetyltransferase and dihydrolipoamide dehydrogenase in the enzyme complex must be 24:24:24 instead of 24:24:12 assumed so far.Electronic Supplementary Material available.  相似文献   

8.
Baker P  Carere J  Seah SY 《Biochemistry》2012,51(22):4558-4567
BphJ, a nonphosphorylating acylating aldehyde dehydrogenase, catalyzes the conversion of aldehydes to form acyl-coenzyme A in the presence of NAD(+) and coenzyme A (CoA). The enzyme is structurally related to the nonacylating aldehyde dehydrogenases, aspartate-β-semialdehyde dehydrogenase and phosphorylating glyceraldehyde-3-phosphate dehydrogenase. Cys-131 was identified as the catalytic thiol in BphJ, and pH profiles together with site-specific mutagenesis data demonstrated that the catalytic thiol is not activated by an aspartate residue, as previously proposed. In contrast to the wild-type enzyme that had similar specificities for two- or three-carbon aldehydes, an I195A variant was observed to have a 20-fold higher catalytic efficiency for butyraldehyde and pentaldehyde compared to the catalytic efficiency of the wild type toward its natural substrate, acetaldehyde. BphJ forms a heterotetrameric complex with the class II aldolase BphI that channels aldehydes produced in the aldol cleavage reaction to the dehydrogenase via a molecular tunnel. Replacement of Ile-171 and Ile-195 with bulkier amino acid residues resulted in no more than a 35% reduction in acetaldehyde channeling efficiency, showing that these residues are not critical in gating the exit of the channel. Likewise, the replacement of Asn-170 in BphJ with alanine and aspartate did not substantially alter aldehyde channeling efficiencies. Levels of activation of BphI by BphJ N170A, N170D, and I171A were reduced by ≥3-fold in the presence of NADH and ≥4.5-fold when BphJ was undergoing turnover, indicating that allosteric activation of the aldolase has been compromised in these variants. The results demonstrate that the dehydrogenase coordinates the catalytic activity of BphI through allostery rather than through aldehyde channeling.  相似文献   

9.
Insulin resistance, the key defect in type 2 diabetes (T2D), is associated with a low capacity to adapt fuel oxidation to fuel availability, i.e., metabolic inflexibility. This, in turn, contributes to a further damage of insulin signaling. Effectiveness of T2D treatment depends in large part on the improvement of insulin sensitivity and metabolic adaptability of the muscle, the main site of whole-body glucose utilization. We have shown previously in mice fed an obesogenic high-fat diet that a combined use of n-3 long-chain polyunsaturated fatty acids (n-3 LC-PUFA) and thiazolidinediones (TZDs), anti-diabetic drugs, preserved metabolic health and synergistically improved muscle insulin sensitivity. We investigated here whether n-3 LC-PUFA could elicit additive beneficial effects on metabolic flexibility when combined with a TZD drug rosiglitazone. Adult male C57BL/6N mice were fed an obesogenic corn oil-based high-fat diet (cHF) for 8 weeks, or randomly assigned to various interventions: cHF with n-3 LC-PUFA concentrate replacing 15% of dietary lipids (cHF+F), cHF with 10 mg rosiglitazone/kg diet (cHF+ROSI), cHF+F+ROSI, or chow-fed. Indirect calorimetry demonstrated superior preservation of metabolic flexibility to carbohydrates in response to the combined intervention. Metabolomic and gene expression analyses in the muscle suggested distinct and complementary effects of the interventions, with n-3 LC-PUFA supporting complete oxidation of fatty acids in mitochondria and the combination with n-3 LC-PUFA and rosiglitazone augmenting insulin sensitivity by the modulation of branched-chain amino acid metabolism. These beneficial metabolic effects were associated with the activation of the switch between glycolytic and oxidative muscle fibers, especially in the cHF+F+ROSI mice. Our results further support the idea that the combined use of n-3 LC-PUFA and TZDs could improve the efficacy of the therapy of obese and diabetic patients.  相似文献   

10.
Pyruvate is located at a metabolic junction of assimilatory and dissimilatory pathways and represents a switch point between respiratory and fermentative metabolism. In Escherichia coli, the pyruvate dehydrogenase complex (PDHC) and pyruvate formate-lyase are considered the primary routes of pyruvate conversion to acetyl-CoA for aerobic respiration and anaerobic fermentation, respectively. During glucose fermentation, the in vivo activity of PDHC has been reported as either very low or undetectable, and the role of this enzyme remains unknown. In this study, a comprehensive characterization of wild-type E. coli MG1655 and a PDHC-deficient derivative (Pdh) led to the identification of the role of PDHC in the anaerobic fermentation of glucose. The metabolism of these strains was investigated by using a mixture of 13C-labeled and -unlabeled glucose followed by the analysis of the labeling pattern in protein-bound amino acids via two-dimensional 13C,1H NMR spectroscopy. Metabolite balancing, biosynthetic 13C labeling of proteinogenic amino acids, and isotopomer balancing all indicated a large increase in the flux of the oxidative branch of the pentose phosphate pathway (ox-PPP) in response to the PDHC deficiency. Because both ox-PPP and PDHC generate CO2 and the calculated CO2 evolution rate was significantly reduced in Pdh, it was hypothesized that the role of PDHC is to provide CO2 for cell growth. The similarly negative impact of either PDHC or ox-PPP deficiencies, and an even more pronounced impairment of cell growth in a strain lacking both ox-PPP and PDHC, provided further support for this hypothesis. The three strains exhibited similar phenotypes in the presence of an external source of CO2, thus confirming the role of PDHC. Activation of formate hydrogen-lyase (which converts formate to CO2 and H2) rendered the PDHC deficiency silent, but its negative impact reappeared in a strain lacking both PDHC and formate hydrogen-lyase. A stoichiometric analysis of CO2 generation via PDHC and ox-PPP revealed that the PDHC route is more carbon- and energy-efficient, in agreement with its beneficial role in cell growth.  相似文献   

11.
Warburg effect of aerobic glycolysis in hepatic M1 macrophages is a major cause for metabolic dysfunction and inflammatory stress in non-alcoholic fatty liver disease (NAFLD). Plant-derived triterpene celastrol markedly inhibited macrophage M1 polarization and adipocyte hypertrophy in obesity. The present study was designed to identify the celastrol-bound proteins which reprogrammed metabolic and inflammatory pathways in M1 macrophages. Pyruvate kinase M2 (PKM2) was determined to be a major celastrol-bound protein. Peptide mapping revealed that celastrol bound to the residue Cys31 while covalent conjugation altered the spatial conformation and inhibited the enzyme activity of PKM2. Mechanistic studies showed that celastrol reduced the expression of glycolytic enzymes (e.g., GLUT1, HK2, LDHA, PKM2) and related signaling proteins (e.g., Akt, HIF-1α, mTOR), shifted aerobic glycolysis to mitochondrial oxidative phosphorylation and skewed macrophage polarization from inflammatory M1 type to anti-inflammatory M2 type. Animal experiments indicated that celastrol promoted weight loss, reduced serum cholesterol level, lipid accumulation and hepatic fibrosis in the mouse model of NAFLD. Collectively, the present study demonstrated that celastrol might alleviate lipid accumulation, inflammation and fibrosis in the liver via covalent modification of PKM2.  相似文献   

12.
Stochastic winter weather events are predicted to increase in occurrence and amplitude at northern latitudes and organisms are expected to cope through phenotypic flexibility. Small avian species wintering in these environments show acclimatization where basal metabolic rate (BMR) and maximal thermogenic capacity (MSUM) are typically elevated. However, little is known on intra-seasonal variation in metabolic performance and on how population trends truly reflect individual flexibility. Here we report intra-seasonal variation in metabolic parameters measured at the population and individual levels in black-capped chickadees ( Poecile atricapillus ). Results confirmed that population patterns indeed reflect flexibility at the individual level. They showed the expected increase in BMR (6%) and MSUM (34%) in winter relative to summer but also, and most importantly, that these parameters changed differently through time. BMR began its seasonal increase in November, while MSUM had already achieved more than 20% of its inter-seasonal increase by October, and declined to its starting level by March, while MSUM remained high. Although both parameters co-vary on a yearly scale, this mismatch in the timing of variation in winter BMR and MSUM likely reflects different constraints acting on different physiological components and therefore suggests a lack of functional link between these parameters.  相似文献   

13.
Accumulation of high level of lactate can negatively impact cell growth during fed-batch culture process. In this study, we attempted to knockout the lactate dehydrogenase A (LDHA) gene in CHO cells in order to attenuate the lactate level. To prevent the potential deleterious effect of pyruvate accumulation, consequent to LDHA knockout, on cell culture, we chose a pyruvate dehydrogenase kinase 1, 2, and 3 (PDHK1, 2, and 3) knockdown cell line in which to knock out LDHA alleles. Around 3,000 clones were screened to obtain 152 mutants. Only heterozygous mutants were identified. An attempt to knockout the remaining wild-type allele from one such heterozygote yielded only two mutants after screening 567 clones. One had an extra valine. Another evidenced a duplication event, possessing at lease one wild-type and two different frameshifted alleles. Both mutants still retained LDH activity. Together, our data strongly suggest that a complete knockout of LDHA is lethal in CHO cells, despite simultaneous down-regulation of PDHK1, 2, and 3.  相似文献   

14.
Objective: To investigate the contribution of meal and snack replacements for long‐term weight maintenance and risk factor reduction in obese patients. Research Methods and Procedures: Prospective, randomized, two‐arm, parallel intervention for 12 weeks followed by a prospective single‐arm 4‐year trial in a University Hospital clinic. One hundred patients, >18 years old and with a body mass index > 25 and ≤ 40 kg/m2, were prescribed a 1200 to 1500 kcal/d control diet (Group A) or an isoenergetic diet, including two meal and snack replacements (vitamin‐ and mineral‐fortified shakes, soups, and bars) and one meal high in fruits and vegetables (Group B). Following a 3 months of weight loss, all patients were prescribed the same energy‐restricted diet (1200 to 1500 kcal) with one meal and one snack replacement for an additional 4 years. Results: All 100 patients were evaluated at 12 weeks. Mean percentage weight loss was 1.5 ± 0.4% and 7.8 ± 0.5% (mean ± SEM) for Groups A and B, respectively. At 12 weeks systolic blood pressure, plasma triacylglycerol, glucose, and insulin concentrations were significantly reduced in Group B, whereas no changes occurred in Group A. After 4 years, 75% of the patients were evaluated. Total mean weight loss was 3.2 ± 0.8% for Group A and 8.4 ± 0.8% (mean ± SEM) for Group B. Both groups showed significant improvement in blood glucose and insulin (p < 0.001), but only Group B showed significant improvement in triacylglycerol and systolic blood pressure compared to baseline values (p < 0.001). Discussion: Providing a structured meal plan via vitamin‐ and mineral‐fortified liquid meal replacements is a safe and effective dietary strategy for obese patients. Long‐term maintenance of weight loss with meal replacements can improve certain biomarkers of disease risk.  相似文献   

15.
The relation between the activation (phosphorylation) state of pyruvate dehydrogenase complex (PDHC; EC 1.2.4.1, EC 2.3.1.12, and EC 1.6.4.3) and the rate of pyruvate oxidation has been examined in isolated, metabolically active, and tightly coupled mitochondria from rat cerebral cortex. With pyruvate and malate as the substrates, the activation state of PDHC decreased on addition of ADP, while the rates of oxygen uptake and 14CO2 formation from [1-14C]pyruvate increased. The lack of correlation between the activation state of PDHC and rate of pyruvate oxidation was seen in media containing 5, 30, or 100 mM KCl. Both the activation state of PDHC and pyruvate oxidation increased, however, when KCl was increased from 5 to 100 mM. Although the PDHC is inactivated by an ATP-dependent kinase (EC 2.7.1.99), direct measurement of ATP and ADP failed to show a consistent relationship between the activation state of PDHC and either ATP levels or ATP/ADP ratios. Comparison of the activation state of PDHC in uncoupled or oligomycin-treated mitochondria also failed to correlate PDHC activation state to adenine nucleotides. In brain mitochondria, unlike those from other tissues, the activation state of PDHC does not seem to be related clearly to the rate of pyruvate oxidation, or to the mitochondrial adenylate energy charge.  相似文献   

16.
The Escherichia coli pyruvate dehydrogenase complex (PDHc) catalyzing conversion of pyruvate to acetyl-CoA comprises three components: E1p, E2p, and E3. The E2p is the five-domain core component, consisting of three tandem lipoyl domains (LDs), a peripheral subunit binding domain (PSBD), and a catalytic domain (E2pCD). Herein are reported the following. 1) The x-ray structure of E2pCD revealed both intra- and intertrimer interactions, similar to those reported for other E2pCDs. 2) Reconstitution of recombinant LD and E2pCD with E1p and E3p into PDHc could maintain at least 6.4% activity (NADH production), confirming the functional competence of the E2pCD and active center coupling among E1p, LD, E2pCD, and E3 even in the absence of PSBD and of a covalent link between domains within E2p. 3) Direct acetyl transfer between LD and coenzyme A catalyzed by E2pCD was observed with a rate constant of 199 s−1, comparable with the rate of NADH production in the PDHc reaction. Hence, neither reductive acetylation of E2p nor acetyl transfer within E2p is rate-limiting. 4) An unprecedented finding is that although no interaction could be detected between E1p and E2pCD by itself, a domain-induced interaction was identified on E1p active centers upon assembly with E2p and C-terminally truncated E2p proteins by hydrogen/deuterium exchange mass spectrometry. The inclusion of each additional domain of E2p strengthened the interaction with E1p, and the interaction was strongest with intact E2p. E2p domain-induced changes at the E1p active site were also manifested by the appearance of a circular dichroism band characteristic of the canonical 4′-aminopyrimidine tautomer of bound thiamin diphosphate (AP).  相似文献   

17.
The grass-specific herbicide haloxyfop, ((±)-2-[4-((3-chloro-5-(trifluoromethyl)-2-pyridinyl)oxy)-phenoxy] propionic acid) has been shown to inhibit lipid synthesis and respiration, to cause the accumulation of amino acids, and not to affect cellular sugar or ATP levels. Thus studies were carried out with enzyme activities from corn (Zea mays L.) (haloxyfop sensitive) and soybean (Glycine max [L.] Merr.) (haloxyfop tolerant) to locate the possible inhibition sites among the glycolytic and tricarboxylic acid (TCA) cycle enzymes. Following along the oxidative metabolism pathway of sugars, the pyruvate dehydrogenase complex (PDC) was the first enzyme among the glycolytic enzymes that demonstrated noticeable inhibition by 1 millimolar haloxyfop. Kinetic studies with corn and soybean PDC from both purified etioplasts and mitochondria gave Ki values of from 1 to 10 millimolar. Haloxyfop also inhibited the activity of the TCA cycle enzyme, the α-ketoglutarate dehydrogenase complex (α-KGDC) which carries out the same reaction as PDC except for the substitution of α-ketoglutarate for pyruvate as one of the substrates. The Ki values were somewhat lower in this case (near 1 millimolar). The relatively high Ki values for both enzyme complexes would indicate that these may not be the herbicidal sites of inhibition, but it is possible that the herbicide could be concentrated in compartments and/or the substrate concentrations may be well below optimal. Likewise little difference was seen in the haloxyfop inhibition of the enzyme activities from the sensitive species, corn, and from the tolerant species, soybean, so the selectivity of the herbicide is not evident from these results. The inhibition of the PDC and α-KGDC as the mode of action of haloxyfop is, however, consistent with the observed physiological effects of the herbicide, and these are the only enzymic activities so far found to be sensitive to haloxyfop.  相似文献   

18.
Abstract: The effect of 3-hydroxybutyrate on pyruvate decarboxylation by neonatal rat brain mitochondria and synaptosomes was investigated. The rate of [1 -14C]pyruvate decarboxylation (1 mm final concentration) by brain synaptosomes derived from 8-day-old rats was inhibited by 10% in the presence of 2 mm -d ,l -3-hydroxybutyrate and by more than 20% in the presence of 20 mm -d ,l -3-hydroxybutyrate. The presence of 2 mm -l ,d -3-hydroxybutyrate did not affect the rate of [1-14T]pyruvate decarboxylation (1 mm final concentration) by brain mitochondria; however, at a concentration of 20 mm -d ,l -3-hydroxybutyrate, a marked inhibition was seen in preparations from both 8-day-old (35% inhibition) and 21-day-old (24% inhibition) but not in those from adult rats. Although the presence of 100 mm -K+ in the incubation medium stimulated the rate of pyruvate decarboxylation by approximately 50% compared with the rate in the presence of 1 mm -K+, the presence of 20 mm -d ,l -3-hydroxybutyrate still caused a marked inhibition in both media (1 and 100 mm -K+). The presence of 20 mm -d ,l -3-hydroxybutyrate during the incubation caused an approximately 20% decrease in the level of the active form of the pyruvate dehydrogenase complex in brain mitochondria from 8-day-old rats. The concentrations of ATP, ADP, NAD+, NADH, acetyl CoA, and CoA were measured in brain mitochondria from 8-day-old rats incubated in the presence of 1 mm -pyruvate alone or 1 mm -pyruvate plus 20 mm -d ,l -3-hydroxybutyrate. Neither the ATP/ADP nor the NADH/NAD+ ratio showed significant changes. The acetyl CoA/CoA ratio was significantly increased by more than twofold in the presence of 3-hydroxybutyrate. The possible mechanisms and physiological significance of 3-hydroxybutyrate inhibition of pyruvate decarboxylation in neonatal rat brain mitochondria are discussed.  相似文献   

19.
20.
The influence of oxygen and temperature on the inactivation of pyruvate, Pi dikinase and NADP-malate dehydrogenase was studied in Zea mays. O2 was required for inactivation of both pyruvate, Pi dikinase and NADP-malate dehydrogenase in the dark in vivo. The rate of inactivation under 2% O2 was only slightly lower than that at 21% O2. The in vitro inactivation of pyruvate, Pi dikinase, while dependent on adenine nucleotides (ADP + ATP), did not require O2.

The postillumination inactivation of pyruvate, Pi dikinase in leaves was strongly dependent on temperature. As temperature was decreased in the dark, there was a lag period of increasing length (e.g. at 17°C there was a lag of about 25 minutes) before inactivation proceeded. Following the lag period, the rate of inactivation decreased with decreasing temperature. The half-time for dark inactivation was about 7 minutes at 32°C and 45 minutes at 17°C. The inactivation of pyruvate, Pi dikinase in vitro following extraction from illuminated leaves was also strongly dependent on temperature, but occurred without a lag period. In contrast, NADP-malate dehydrogenase was rapidly inactivated in leaves (half-time of approximately 3 minutes) during the postillumination period without a lag, and there was little effect of temperature between 10 and 32°C. The results are discussed in relation to known differences in the mechanism of activation/inactivation of the two enzymes.

  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