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.
The gene coding for d-psicose 3-epimerase (DPEase) from Clostridium sp. BNL1100 was cloned and expressed in Escherichia coli. The recombinant enzyme was purified by Ni-affinity chromatography. It was a metal-dependent enzyme and required Co2+ as optimum cofactor. It displayed catalytic activity maximally at pH 8.0 and 65 °C (as measured over 5 min). The optimum substrate was d-psicose, and the K m, turnover number (k cat), and catalytic efficiency (k cat/K m) for d-psicose were 227 mM, 32,185 min?1, and 141 min?1 mM?1, respectively. At pH 8.0 and 55 °C, 120 g d-psicose l?1 was produced from 500 g d-fructose l?1 after 5 h.  相似文献   

2.
Hamster liver glutathione peroxidase was purified to homogeneity in three chromatographic steps and with 30% yield. The purified enzyme had a specific activity of approximately 500 μmol cumene hydroperoxide reduced/min/mg of protein at 37 °C, pH 7.6, and 0.25 mm GSH. The enzyme was shown to be a tetramer of indistinguishable subunits, the molecular weight of which was approximately 23,000 as estimated by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. A single isoelectric point of 5.0 was attributed to the active enzyme. Amino acid analysis determined that selenocysteine, identified as its carboxymethyl derivative, was the only form of selenium. One residue of cysteine was found to be present in each glutathione peroxidase subunit. The presence of tryptophan was colorimetrically determined. Pseudo-first-order kinetics of inactivation of the enzyme by iodoacetate was observed at neutral pH with GSH as the only reducing agent. An optimal pH of 8.0 at 37 °C and an activation energy of 3 kcal/mol at pH 7.6 were found. A ter-uni-ping-pong mechanism was shown by the use of an integrated-rate equation. At pH 7.6, the apparent second-order rate constants for reaction of glutathione peroxidase with hydroperoxides were as follows: k1 (t-butyl hydroperoxide), 7.06 × 105 mm min?1; k1 (cumene hydroperoxide), 1.04 × 106 mm?1 min?1; k1 (p-menthane hydroperoxide), 1.2 × 106 mm?1 min?1; k1 (diisopropylbenzene hydroperoxide), 1.7 × 106 mm?1 min?1; k1 (linoleic acid hydroperoxide), 2.36 × 106 mm?1 min?1; k1 (ethyl hydroperoxide), 2.5 × 106 mm?1 min?1; and k1 (hydrogen peroxide), 2.98 × 106 mm?1 min?1. It is concluded that for bulky hydroperoxides, the more hydrophobic the substrate, the faster its reduction by glutathione peroxidase.  相似文献   

3.
An alkaline serine-proteinase from Bacillus sp. PN51 isolated from bat feces collected in Phang Nga, Thailand, was purified and characterized. The molecular mass was estimated to be 35.0 kDa. The N-terminal 25 amino acid sequence was about 70% identical with that of Natrialba magadii halolysin-like extracellular serine protease. The enzyme showed the highest proteinase activity at 60 °C at pH 10.0. The activity was strongly inhibited by PMSF and chymostatin. The proteinase activity was not affected by the presence of 2% urea, 2% H2O2, 12% SDS, 15% triton X-100, or 15% tween 80. The proteinase preferred Met, Leu, Phe, and Tyr residues at the P1 position, in descending order. The k cat, K m and k cat/K m values for Z-Val-Lys-Met-MCA were 16.8±0.14 min?1, 5.1±0.28 μM, and 3.3±0.28 μM?1 min?1 respectively. This is the first report of an alkaline serine-proteinase with extremely high stability against detergents such as SDS.  相似文献   

4.
(1) The kinetic parameters of rat pancreatic adenylate cyclase were evaluated, using GTP, p[NH]ppG or GTPγS as nucleotide activator, cholecystokinin as peptide hormone, and GDPβS and dibutyryl cyclic GMP as inhibitors of guanosine triphosphate and CCK-8, respectively. The time courses of activation and the degree of activation at steady state (EA/ETOT) were compatible with a simple two-state model of activation-deactivation based on a pseudo-monomolecular activation process (rate constant k+2, and a deactivation process (rate constant koff) that included, depending on the activating nucleotide, the hydrolysis of GTP (rate constant k2) and/or the dissociation of the intact nucleotide (rate constant k?1), so that EA/ETOT = k+1/(k+1 + k2 + k?a). (2) The hormone CCK-8 increased the value of k+1 with GTP dose-dependently, from 0.2 to 10.9 min?1. The value of k?1 increased 0.01 to 0.3 min?1 but the value of k2 was unaltered at 7 min?1, so that EA/ETOT increased 15-fold, from 4% to 61%. (3) A cholera toxin pretreatment at 30 μg/ml allowed also a large increase in EA/ETOT with GTP (up to 51%) but the underlying mechanism was different. It consisted of a 14-fold decrease in the koff value of the GTP-activated enzyme (from 7 min? to 0.5 min?1) that corresponded to a reduction in GTPase activity. When testing the system with p[NH]ppG, two added effects of the cholera toxin pretreatment were observed: a 4-fold increase in the value of k+1 (from 0.2 to 0.8 min?1) and the occurrence of a significant 0.3 min?1 value for k?1.  相似文献   

5.
Abstract: A previous study of the metabolism of 6-[18F]-fluoro-l -3,4-dihydroxyphenylalanine (FDOPA) in rats pretreated with carbidopa contained information amenable to kinetic analysis. Using these data, tracer transfer coefficients and metabolic rate constants were estimated. After intravenous injection, FDOPA in circulation was O-methylated (kD0 = 0.055 min?1), and the metabolite (O-methyl-FDOPA) escaped from plasma with a rate constant (kM?1) of 0.01 min?1. The initial clearance of FDOPA to striatum (KD1) was 0.07 ml g?1 min?1, and the equilibrium distribution volume (VDe) was 0.67 ml g?1. The initial clearance of O-methyl-FDOPA to striatum (KM1) was 0.08 ml g?1 min?1, and the equilibrium distribution volume (VMe) was 0.75 ml g?1. The rate constant of FDOPA decarboxylation (kD3) was 0.17 min?1 in striatum. The elimination of 6-[18F]fluorodopamine (FDA) from striatum suggested an apparent rate constant for monoamine oxidase activity (k7) of 0.055 min?1. 6-[18F]Fluorohomovanillic acid (FHVA) was formed from 6-[18F]fluoro-l -3,4-dihydroxyphenylacetic acid with a rate constant (k11) of 0.083 min?1, and FHVA was eliminated from striatum (k9) with a rate constant of 0.12 min?1. The steady-state concentration ratios of FDA and its metabolites were shown to be functions of these rate constants.  相似文献   

6.
The gene coding for ribose-5-phosphate isomerase (Rpi) from Thermotoga lettingae TMO was cloned and expressed in E. coli. The recombinant enzyme was purified by Ni-affinity chromatography. It converted d-psicose to d-allose maximally at 75 °C and pH 8.0 with a 32 % conversion yield. The k m, turnover number (k cat), and catalytic efficiency (k cat k m ?1 ) for substrate d-psicose were 64 mM, 6.98 min?1 and 0.11 mM?1 min?1 respectively.  相似文献   

7.
A study of the sulfhydryl groups of rat brain hexokinase   总被引:1,自引:0,他引:1  
Rat brain hexokinase (ATP: d-hexose-6-phosphotransferase, EC 2.7.1.1) is rapidly inactivated by reaction with 5,5′-dithiobis-(2-nitrobenzoate). The inactivation follows monophasic first-order kinetics in either the absence of ligands (k = 0.641 min?1 at 25 °C) or in the presence of saturating levels of ATP (free or complexed with Mg2+) or P1; the inactivation rate is slightly increased (k ? 0.7 min ?1) in the presence of ATP or P1. In contrast, glucose and glucose-6-P markedly decrease the inactivation rate; inactivation in the presence of these ligands is biphasic, with two first-order rates (k ? 0.5 min?1 and 0.01 min?1) being distinguishable.The enzyme contains 14 sulfhydryl groups which react with 5,5′-dithiobis-(2-nitrobenzoate); reaction of these groups in the native enzyme is complete after 2 hr at 25 °C, or in approx 5 min with the urea or guanidine-denatured enzyme. In the native enzyme, three classes of sulfhydryl groups are distinguishable and are designated as F-, I-, or S-type based on their fast (k ? 0.7 min?1), intermediate (k ? 0.5-0.7 min?1), or slow (k ? 0.02 min?1 rates of reaction with 5,5′-dithiobis-(2-nitrobenzoate). The correlation of inactivation rates with the rates for reaction of the I-type sulfhydryls indicates that the I-type sulfhydryls include residues necessary for catalytic activity. The F-type residues are clearly not required for activity.The effects of ATP, P1, glucose, and glucose-6-P on the reactivity of the sulfhydryls have been determined. As in the absence of ligands, S-, I-, and F-type sulfhydryls could be distinguished. In the presence of saturating concentrations of these ligands, the F, I, and S classes of sulfhydryls contained respectively: with ATP, 1, 4, and 7 residues; with P1, 1, 3, and 7 residues; with glucose, 1, 2, and 5 residues; with glucose-6-P, 1, 2, and 1 residues. Comparison with rate constants for inactivation in the presence of these ligands again indicated that I-type sulfhydryls were particularly important in maintenance of enzyme activity. The present results indicate considerable similarity between the reactivity of the sulfhydryl residues in rat brain hexokinase and the sulfhydryls of the bovine brain enzyme [V. D. Redkar and U. W. Kenkare (1972), J. Biol. Chem., 247, 7576–7584].  相似文献   

8.
Brush border membranes of the rabbit renal tubule have an ATPase which was stimulated 60% by 50 mm HCO3?. The Ka for HCO3? was 36 mm. Kinetic studies of the “HCO3?-ATPase” indicate that HCO3? had no effect on the Km for ATP and ATP did not alter the Ka for HCO3?. Several anions, notably SO32?, also accelerated the rate of dephosphorylation of ATP. The V for “SO32?-ATPase” was fivefold greater than that for “HCO3?-ATPase.” The Ka for SO32? was 0.78 mm. Other anions including Cl? and phosphates, did not enhance ATPase activity. Thus, of the anions present in the glomerular filtrate in appreciable concentrations only HCO3? stimulated the luminal membrane enzyme. The anion-stimulated ATPase activity increased sharply from pH 6.1 to 7.1 and moderately with higher pH. The renal ATPase was not inhibited by SCN? nor methyl sulfonyl chloride and was relatively insensitive to oligomycin and quercetin. Carbonyl cyanide p-trifluoromethoxy phenylhydrazone increased the basal rate of the membranal ATPase, suggesting that the ATPase activity is limited by transmembrane H+ flux. Carbonic anhydrase significantly increased the HCO3?-stimulated ATPase activity. This increment was blocked by Diamox. These findings provide evidence consistent with the hypothesis that the brush border membrane ATPase is involved in the extrusion of H+ from tubular cell to lumen and support suggested interrelationships between HCO3?-stimulated ATPase, H+ secretion, and bicarbonate transport in the kidney.  相似文献   

9.
The role of exposed tyrosine side-chains in enzyme-catalysed reactions has been studied for porcine-pancreatic alpha-amylase, sweet-potato beta-amylase, and Aspergillus niger glucamylase using N-acetylimidazole as the specific protein reagent. The changes in activity binding affinity (Δk?1/k+1), and kinetic parameters (Km,k2) due to acetylation of the phenolic hydroxyl groups have been determined. Acetylation of each enzyme occurred by an “apparent” first-order reaction with a rate constant of 0.72–1.4 x 10?1min?1. Acetylation increased the apparent Km (soluble starch as the substrate) for each enzyme (appreciably for alpha-amylase and glucamylase), whereas k2 remained unchanged. Similarly, for each enzyme, the binding affinity for immobilised cyclohexa-amylose decreased appreciably, whereas the catalytic activity was reduced only to a small degree (and remained unchanged for beta-amylase). It is concluded that the tyrosine groups located in the active centre of each enzyme have a substrate-binding function.  相似文献   

10.
The oxygen-transporting protein, hemocyanin (Hc), of the garden snail Helix aspersa maxima (HaH) was isolated and kinetically characterized. Kinetic parameters of the reaction of catalytic oxidation of catechol to quinone, catalyzed by native HaH were determined: the V max value amounted to 22 nmol min?1 mg?1, k cat to 1.1 min?1. Data were compared to those reported for other molluscan Hcs and phenoloxidases (POs). The o-diphenoloxidase activity of the native HaH is about five times higher than the activity determined for the Hcs of the terrestrial snail Helix pomatia and of the marine snail Rapana thomasiana (k cat values of 0.22 and 0.25 min?1, respectively). The K m values obtained for molluscan Hcs from different species are comparable to those for true POs, but the low catalytic efficiency of Hcs is probably related to inaccessibility of the active sites to potential substrates. Upon treatment of HaH with subtilisin DY, the enzyme activity against substrate catechol was considerably increased. The relatively high proteolytically induced o-diPO activity of HaH allowed using it for preparation of a biosensor for detection of catechol.  相似文献   

11.
Cystathionine β-synthase (CBS) catalyzes the pyridoxal 5′-phosphate (PLP)-dependent condensation of l-serine and l-homocysteine to form l-cystathionine in the first step of the reverse transsulfuration pathway. Residue S289 of yeast CBS, predicted to form a hydrogen bond with the pyridine nitrogen of the PLP cofactor, was mutated to alanine and aspartate. The kcat/Kml-Ser of the S289A mutant is reduced by a factor of ~ 800 and the β-replacement activity of the S289D mutant is undetectable. Fluorescence energy transfer between tryptophan residue(s) of the enzyme and the PLP cofactor, observed in the wild-type enzyme and diminished in the S289A mutant, is absent in S289D. These results demonstrate that residue S289 is essential in maintaining the properties and orientation of the pyridine ring of the PLP cofactor. The reduction in activity of ytCBS-S289A suggests that ytCBS catalyzes the α,β-elimination of l-Ser via an E1cB mechanism.  相似文献   

12.
A recombinant alcohol dehydrogenase (ADH) from Kangiella koreensis was purified as a 40 kDa dimer with a specific activity of 21.3 nmol min?1 mg?1, a K m of 1.8 μM, and a k cat of 1.7 min?1 for all-trans-retinal using NADH as cofactor. The enzyme showed activity for all-trans-retinol using NAD + as a cofactor. The reaction conditions for all-trans-retinol production were optimal at pH 6.5 and 60 °C, 2 g enzyme l?1, and 2,200 mg all-trans-retinal l?1 in the presence of 5 % (v/v) methanol, 1 % (w/v) hydroquinone, and 10 mM NADH. Under optimized conditions, the ADH produced 600 mg all-trans-retinol l?1 after 3 h, with a conversion yield of 27.3 % (w/w) and a productivity of 200 mg l?1 h?1. This is the first report of the characterization of a bacterial ADH for all-trans-retinal and the biotechnological production of all-trans-retinol using ADH.  相似文献   

13.
The dissociation of insulin from human insulin antibodies has been investigated using a technique that is rapid and does not require addition of excess unlabelled insulin. A slow (k1 = 2·1?3 min?1 and a fast (k2 = 4·10?2 min?1) dissociating antibody component were identified in all studies. These have been shown to correspond, respectively, to the high and low affinity antibody components of equilibrium binding studies. The range of k1 and k2 values and their response to temperature change is small. Insulin resistance and stability of diabetes are not related to properties of antibody dissociation. Dissociation is faster in the presence of high (6–850 nM) insulin concentration due to increased binding to the fast dissociating component without change in the dissociation rate constants. When incubation time is increased beyond achivement of maximal binding there is a time-dependent rise in binding to the slow dissociating component, with a concomitant fall in k1. The traditional concept that equilibrium is established at maximum binding requires further examination.  相似文献   

14.
We studied the influence of inorganic nitrogen sources (NO3 ? or NH4 +) and potassium deficiency on expression and activity of plasma membrane (PM) H+-ATPase in sorghum roots. After 15 d of cultivation at 0.2 mM K+, the plants were transferred to solutions lacking K+ for 2 d. Then, K+ depletion assays were performed in the presence or absence of vanadate. Further, PMs from K+-starved roots were extracted and used for the kinetic characterization of ATP hydrolytic activity and the immunodetection of PM H+-ATPase. Two major genes coding PM H+-ATPase (SBA1 and SBA2) were analyzed by real-time PCR. PM H+-ATPase exhibited a higher Vmax and Km in NH4 +-fed roots compared with NO3 ? -fed roots. The optimum pH of the enzyme was slightly lower in NO3 ? -fed roots than in NH4 +-fed roots. The vanadate sensitivity was similar. The expressions of SBA1 and SBA2 increased in roots grown under NH4 +. Concomitantly, an increased content of the enzyme in PM was observed. The initial rate of K+ uptake did not differ between plants grown with NO3 ? or NH4 +, but it was significantly reduced by vanadate in NH4 +-grown plants.  相似文献   

15.
The effects of K+, Na+ and ATP on the gastric (H+ + K+)-ATPase were investigated at various pH. The enzyme was phosphorylated by ATP with a pseudo-first-order rate constant of 3650 min?1 at pH 7.4. This rate constant increased to a maximal value of about 7900 min?1 when pH was decreased to 6.0. Alkalinization decreased the rate constant. At pH 8.0 it was 1290 min?1. Additions of 5 mM K+ or Na+, did not change the rate constant at acidic pH, while at neutral or alkaline pH a decrease was observed. Dephosphorylation of phosphoenzyme in lyophilized vesicles was dependent on K+, but not on Na+. Alkaline pH increased the rate of dephosphorylation. K+ stimulated the ATPase and p-nitrophenylphosphatase activities. At high concentrations K+ was inhibitory. Below pH 7.0 Na+ had little or no effect on the ATPase and p-nitrophenylphosphatase, while at alkaline pH, Na+ inhibited both activities. The effect of extravesicular pH on transport of H+ was investigated. At pH 6.5 the apparent Km for ATP was 2.7 μM and increased little when K+ was added extravesicularly. At pH 7.5, millimolar concentrations of K+ increased the apparent Km for ATP. Extravesicular K+ and Na+ inhibited the transport of H+. The inhibition was strongest at alkaline pH and only slight at neutral or acidic pH, suggesting a competition between the alkali metal ions and hydrogen ions at a common binding site on the cytoplasmic side of the membrane. Two H+-producing reactions as possible candidates as physiological regulators of (H+ + K+)-ATPase were investigated. Firstly, the hydrolysis of ATP per se, and secondly, the hydration of CO2 and the subsequent formation of H+ and HCO3?. The amount of hydrogen ions formed in the ATPase reaction was highest at alkaline pH. The H+/ATP ratio was about 1 at pH 8.0. When CO2 was added to the reaction medium there was no change in the rate of hydrogen ion transport at pH 7.0, but at pH 8.0 the rate increased 4-times upon the addition of 0.4 mM CO2. The results indicate a possible co-operation in the production of acid between the H+ + K+-ATPase and a carbonic anhydrase associated with the vesicular membrane.  相似文献   

16.
The hemicellulose xylan constitutes a major portion of plant biomass, a renewable feedstock available for conversion to biofuels and other bioproducts. β-xylosidase operates in the deconstruction of the polysaccharide to fermentable sugars. Glycoside hydrolase family 43 is recognized as a source of highly active β-xylosidases, some of which could have practical applications. The biochemical details of four GH43 β-xylosidases (those from Alkaliphilus metalliredigens QYMF, Bacillus pumilus, Bacillus subtilis subsp. subtilis str. 168, and Lactobacillus brevis ATCC 367) are examined here. Sedimentation equilibrium experiments indicate that the quaternary states of three of the enzymes are mixtures of monomers and homodimers (B. pumilus) or mixtures of homodimers and homotetramers (B. subtilis and L. brevis). k cat and k cat/K m values of the four enzymes are higher for xylobiose than for xylotriose, suggesting that the enzyme active sites comprise two subsites, as has been demonstrated by the X-ray structures of other GH43 β-xylosidases. The K i values for d-glucose (83.3–357 mM) and d-xylose (15.6–70.0 mM) of the four enzymes are moderately high. The four enzymes display good temperature (K t 0.5?~?45 °C) and pH stabilities (>4.6 to <10.3). At pH 6.0 and 25 °C, the enzyme from L. brevis ATCC 367 displays the highest reported k cat and k cat/K m on natural substrates xylobiose (407 s?1, 138 s?1?mM?1), xylotriose (235 s?1, 80.8 s?1?mM?1), and xylotetraose (146 s?1, 32.6 s?1?mM?1).  相似文献   

17.
D-Lactate dehydrogenase (D-LDH) from Pediococcus pentosaceus ATCC 25745 was found to produce D-3-phenyllactic acid from phenylpyruvate. The optimum pH and temperature for enzyme activity were pH 5.5 and 45 °C. The Michaelis-Menten constant (K m), turnover number (k cat), and catalytic efficiency (k cat?K m) values for the substrate phenylpyruvate were estimated to be 1.73 mmol/L, 173 s?1, and 100 (mmol/L)?1 s?1 respectively.  相似文献   

18.
Pyridoxal 5′-phosphate (PLP) and CO2 were competitive for the same site on d-ribulose bisphosphate carboxylase (EC 4.1.1.39), presumably an ?-amino group of a lysyl residue. An apparent noncompetitive inhibition occurred with respect to ribulose bisphosphate (RuP2). The time course of inactivation was first order with respect to both PLP and RuP2 carboxylase concentration. The extinction coefficient was found to be 5800 ± 800 m?1 cm?1 for the maximal absorbance at 432 nm of the enzyme/PLP Schiff base. The titration of the enzyme with PLP gave a biphasic double reciprocal plot. The number of amino groups reacting with PLP was calculated to be 9.5 and 19 per molecule of enzyme, at low and high concentrations of PLP. Half of the amino groups available for reaction with PLP at either concentration could be protected by RuP2. When the RuP2 carboxylase/PLP complex was reduced with NaBH4 in the absence of substrates, only 20% of the enzyme activity was recovered, but, in the presence of RuP2 or bicarbonate, the recovery of enzyme activity was 80 or 25%, respectively. It is concluded that there are eight primary groups that react with PLP, one at each of the eight catalytic sites of the RuP2 carboxylase molecule. It is postulated that this amine is essential for formation and/or activation of an enzyme/CO2 complex. In the absence of CO2, this amine may react with the carbonyl function of RuP2. This provides an explanation for the inactivation of RuP2 carboxylase upon preincubation with RuP2 in the absence of CO2 and even the inactivation of the activated enzyme at RuP2 concentrations above 0.5 mm. The second set of eight primary amino groups which react with PLP may involve the CO2 regulatory site.  相似文献   

19.
1. The oligomeric dicyclohexylcarbodiimide (DCCD)-binding protein of mitochondrial ATPase was studied using (a) the relationship between [14C]DCCD binding and inhibition of ATPase activities and (b) the analysis of the kinetics of inhibition. 2. The [14C]DCCD binding to bovine heart mitochondria is linearly proportional to the inhibition of ATP hydrolysis up to a 50% decrease of the original activity resulting in 0.6 mol DCCD bound covalently to the specific inhibitory site (Hous?t?k, J., Svoboda, P., Kopecký, J., Kuz?ela, S?. and Drahota, Z. (1981) Biochim. Biophys. Acta 634, 331–339) per mol of the fully inhibited enzyme. 3. Kinetics of the inhibition of both the ATPase activity (heart and liver mitochondria) and ADP-stimulated respiration (liver) reveal that 1 mol DCCD per mol ATPase eliminates both the synthetic and the hydrolytic activities. It is inferred that the activity-binding correlation underestimates the number of DCCD-reactive sites. 4. The second-order rate constant of the DCCD-ATPase interaction (k) is inversely related to the concentration of membranes, indicating that DCCD reaches the inhibitory site by concentrating in the hydrophobic (phospholipid) environment. 5. At a given concentration of liver mitochondria, comparable k values are obtained both for the inhibition of ATP hydrolysis (k=5.35·102M?1·min?1) and ADP-stimulated respiration (k=5.67·102M?1·min?1). 6. It is concluded that both the synthetic and the hydrolytic functions of ATPase are inhibited via a common single DCCD-reactive site. This site is represented by one of the several polypeptide chains forming the oligomer of the DCCD-binding protein. The inhibitor-ATPase interaction does not exhibit cooperativity, indicating that the preferential reactivity towards DCCD is an inherent property of the inhibitory site.  相似文献   

20.

Objective

To demonstrate the feasibility of simultaneous acquisition of 18F-FDG-PET, diffusion-weighted imaging (DWI) and T1-weighted dynamic contrast-enhanced MRI (T1w-DCE) in an integrated simultaneous PET/MRI in patients with head and neck squamous cell cancer (HNSCC) and to investigate possible correlations between these parameters.

Methods

17 patients that had given informed consent (15 male, 2 female) with biopsy-proven HNSCC underwent simultaneous 18F-FDG-PET/MRI including DWI and T1w-DCE. SUVmax, SUVmean, ADCmean, ADCmin and K trans, k ep and v e were measured for each tumour and correlated using Spearman’s ρ.

Results

Significant correlations were observed between SUVmean and K trans (ρ = 0.43; p ≤ 0.05); SUVmean and k ep (ρ = 0.44; p ≤ 0.05); K trans and k ep (ρ = 0.53; p ≤ 0.05); and between k ep and v e (ρ = -0.74; p ≤ 0.01). There was a trend towards statistical significance when correlating SUVmax and ADCmin (ρ = -0.35; p = 0.08); SUVmax and K trans (ρ = 0.37; p = 0.07); SUVmax and k ep (ρ = 0.39; p = 0.06); and ADCmean and v e (ρ = 0.4; p = 0.06).

Conclusion

Simultaneous 18F-FDG-PET/MRI including DWI and T1w-DCE in patients with HNSCC is feasible and allows depiction of complex interactions between glucose metabolism, microcirculatory parameters and cellular density.  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