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 265 毫秒
1.
S X Lin  J P Shi  X D Cheng  Y L Wang 《Biochemistry》1988,27(17):6343-6348
A Blue Sephadex G-150 affinity column adsorbs the arginyl-tRNA synthetase of Escherichia coli K12 and purifies it with high efficiency. The relatively low enzyme content was conveniently purified by DEAE-cellulose chromatography, affinity chromatography, and fast protein liquid chromatography to a preparation with high activity capable of catalyzing the esterification of about 23,000 nmol of arginine to the cognate tRNA per milligram of enzyme within 1 min, at 37 degrees C, pH 7.4. The turnover number is about 27 s-1. The purification was about 1200-fold, and the overall yield was more than 30%. The enzyme has a single polypeptide chain of about Mr 70,000 and binds arginine and tRNA with 1:1 stoichiometry. For the aminoacylation reaction, the Km values at pH 7.4, 37 degrees C, for various substrates were determined: 12 microM, 0.9 mM, and 2.5 microM for arginine, ATP, and tRNA, respectively. The Km value for cognate tRNA is higher than those of most of the aminoacyl-tRNA synthetase systems so far reported. The ATP-PPi exchange reaction proceeds only in the presence of arginine-specific tRNA. The Km values of the exchange at pH 7.2, 37 degrees C, are 0.11 mM, 2.9 mM, and 0.5 mM for arginine, ATP, and PPi, respectively, with a turnover number of 40 s-1. The pH dependence shows that the reaction is favored toward slightly acidic conditions where the aminoacylation is relatively depressed.  相似文献   

2.
O Goerlich  E Holler 《Biochemistry》1984,23(2):182-190
The synthesis of diadenosine 5',5"'-P1-,P4-tetraphosphate (Ap4A) catalyzed by phenylalanyl-tRNA synthetase in the presence of Zn2+ involves the same partial reactions (synthesis of phenylalanyladenylate and transfer of the adenylate moiety to ATP) as occur in the absence of this metal ion. However, transfer is strongly stimulated while adenylate synthesis is depressed. Also inhibited are pyrophosphorolysis of phenylalanyladenylate and transfer of phenylalanine from the adenylate to cognate tRNA, because overall tRNA phenylalanylation is depressed [Mayaux, J.-F., & Blanquet, S. (1981) Biochemistry 20, 4647-4654], whereas binding of tRNA to the synthetase is not. At moderate concentrations of Zn2+, and in the presence of 5 microM phenylalanine and 0.5 mM ATP, transfer of AMP is rate limiting, while at higher concentrations of Zn2+ synthesis of adenylate is rate determining. The Zn2+ concentration optimum for stimulation depends on the concentration of phenylalanine and ATP. The effects of Zn2+ are mediated through two classes of binding site(s) on the synthetase, the half-saturations of which are 1-4 and 20-30 microM Zn2+, respectively. Binding of Zn2+ to the second class of site(s) causes inhibition of the synthetase, whereas binding to the first class is responsible for activation and inhibition, which may be caused by a conformational change. Evidence for the latter is the observed decrease in protein intrinsic fluorescence intensity and the decrease in fluorescence intensity of 6-(p-toluidinyl)naphthalene-2-sulfonate, which is used as a reporter group. The kinetics of the binding reaction show a saturation dependence on Zn2+, also suggesting that a conformational change occurs.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
Catalytic behavior of Pseudomonas cepacia lipase in w/o microemulsions   总被引:3,自引:0,他引:3  
The activity of purified Pseudomonas cepacia lipase has been investigated in esterification reactions of various aliphatic alcohols with natural fatty acids. The reactions were carried out in microemulsions formed in isooctane by bis-(2-ethylhexyl)sulfosuccinate sodium salt (AOT). Kinetic studies showed that the reaction follows a ping-pong bi-bi mechanism with inhibition by both substrates. The apparent kinetic parameters of the reaction were found to be K(m octanol) = 310 mM, K(m lauric acid) = 78 mM, and V(max) = 250 mumol min(-1) mg(-1). The same system was used for the synthesis of mono- and diglycerides from glycerol and lauric acid, which was successful at very low w(o) values. The catalytic behavior of P. cepacia lipase was also studied in esterification reactions performed in a nonionic microemulsion system formulated by tetraethyleneglycoldodecylether (C(12)E(4)). The optimum activity was found at about w(o) = 8. The apparent values of V(max app) and K(m app) for octanol were calculated and found to be 100 mumol min(-1) mg(-1) and 76 mM, respectively. (c) 1995 John Wiley & Sons, Inc.  相似文献   

4.
In nucleic acid polymerization reaction, pyrophosphorolysis is the reversal of nucleotide addition, in which the terminal nucleotide is excised in the presence of inorganic pyrophosphate (PPi). The CCA enzymes are unusual RNA polymerases, which catalyze CCA addition to positions 74-76 at the tRNA 3′ end without using a nucleic acid template. To better understand the reaction mechanism of CCA addition, we tested pyrophosphorolysis of CCA enzymes, which are divided into two structurally distinct classes. Here, we show that only class II CCA enzymes catalyze pyrophosphorolysis and that the reaction can initiate from all three CCA positions and proceed processively until the removal of nucleotide C74. Pyrophosphorolysis of class II enzymes establishes a fundamental difference from class I enzymes, and it is achieved only with the tRNA structure and with specific divalent metal ions. Importantly, pyrophosphorolysis enables class II enzymes to efficiently remove an incorrect A75 nucleotide from the 3′ end, at a rate much faster than the rate of A75 incorporation, suggesting the ability to perform a previously unexpected quality control mechanism for CCA synthesis. Measurement of kinetic parameters of the class II Escherichia coli CCA enzyme reveals that the enzyme catalyzes pyrophosphorolysis slowly relative to the forward nucleotide addition and that it exhibits weak binding affinity to PPi relative to NTP, suggesting a mechanism in which PPi is rapidly released after each nucleotide addition as a driving force to promote the forward synthesis of CCA.  相似文献   

5.
The transport pathways for dibasic amino acids were investigated in brush border membrane vesicles (BBMV) from the anterior-middle (AM) and posterior (P) regions of Bombyx mori midgut. In the absence of K(+), a low-affinity saturable transport of arginine in both AM- and P-BBMV (K(m) 1.01 mM, V(max) 4.07 nmol/7s/mg protein and K(m) 1.38 mM, V(max) 2.26 nmol/7s/mg protein, respectively) was detected. Arginine influx was dependent on the membrane electrical potential (Deltapsi) and increased raising the alkalinity of the external medium from pH 7.2 to 10.6. Competition experiments indicated the following order of substrate affinity: arginine, homoarginine, N(G)-monomethylarginine, N(G)-nitroarginine>lysine>ornithine>cysteine>methionine. Leucine, valine and BCH (2-amino-2-norbornanecarboxylic acid) did not inhibit arginine influx. In the presence of external K(+), the influx of arginine as a function of arginine concentration fitted to a complex saturation kinetics compatible with both a low-affinity and a high-affinity component. The latter (K(m) 0.035 mM, V(max) 2.54 nmol/7s/mg protein) was fully characterized. The influx rate had an optimum at pH 8.8, was strongly affected by Deltapsi and was homogeneous along the midgut. The substrate affinity rank was: homoarginine>arginine, N(G)-monomethylarginine>cysteine, lysine>N(G)-nitroarginine>ornithine>methionine. Leucine and amino acids with a hydrophobic side chain were not accepted. This system is also operative in the absence of potassium, with the same order of specificity but a very low activity. Lysine influx is mediated by two more transport systems, the leucine uniport and the K(+)/leucine symport specific for amino acids with a hydrophobic side chain that recognizes lysine at extravesicular pH values (pH(out)) exceeding 9. Both the uniport and the symport differ from the cationic transport systems so far identified in mammals because they are unaffected by N-ethylmaleimide, have no significant affinity for neutral amino acids in the presence of the cation and show a striking difference in their optimum pH.  相似文献   

6.
The kinetic properties of honeybee arginine phosphokinase (APK), which catalyzes the reaction: Arginine phosphate + ADP + H+ ? arginine + ATP, have been studied.In the direction of ATP synthesis, the pH optimum was around pH 7.2 and the activation energy over the range 18–44 °C was about 10,500 cal/mole. The optimum ratio of Mg2+:ADP was about 4:1.In the direction of arginine phosphate (AP) synthesis, the enzyme had a pH optimum around pH 8.3. The energy of activation for the reaction over the range 22–39 °C was about 7500 cal/mole. The optimum ratio of Mg2+:ATP was about 1:1.The initial velocities of the reactions in the direction of ATP and AP synthesis were measured at varying concentrations of one substrate while the concentration of the other substrate was held constant at several levels. The double reciprocal plots of the data obtained yielded a series of intersecting lines, indicating that the enzyme has a sequential mechanism. Radioisotope exchange experiment showed that arginine phosphokinase did not catalyze ATP ? ADP exchange in the absence of arginine. Product inhibition studies showed that arginine was competitive with AP and noncompetitive with ADP; whereas ATP was competitive with ADP and noncompetitive with arginine. The results from initial velocity, radioisotope exchange, and product inhibition studies suggested that the enzyme has a rapid equilibrium, random mechanism.  相似文献   

7.
The pH dependence and kinetics parameters of renin-angiotensinogen reactions were determined using wild-type and S84G mutant human renins and wild-type and H13Y mutant sheep angiotensinogens. It is explained in this report that (i) renin catalyzes acidic and basic reactions of which the optimum pHs are 5.5 and 7.5-8.2 respectively, both of which produce angiotensin I; (ii) Ser84 specific to human renin accelerates the acidic reaction by 75-110% through elevation of V(max), and shifts the optimum pH of the basic reaction from 7.5 to 8.0-8.2; and (iii) His13 specific to sheep angiotensinogen accelerates the acidic and basic reactions by 25-42% through reduction of K(m). It is concluded from these results that the coexistence of Ser84 in renin and His13 in angiotensinogen brings a pH profile of two separate peaks at pHs 5.5 and 8.2 to the reaction of human renin and sheep angiotensinogen.  相似文献   

8.
To study the role of 5-methylcytidine in the aminoacylation of mammalian tRNA, bulk tRNA specifically deficient in 5-methylcytidine was isolated from the livers of mice treated with 5-azacytidine (18 mg/kg) for 4 days. For comparison, more extensively altered tRNA was isolated from the livers of mice treated with DL-ethionine (100 mg/kg) plus adenine (48 mg/kg) for 3 days. The amino acid acceptor capacity of these tRNAs was determined by measuring the incorporation of one of eight different 14C-labeled amino acids or a mixture of 14C-labeled amino acids in homologous assays using a crude synthetase preparation isolated from untreated mice. The 5-methylcytidine-deficient tRNA incorporated each amino acid to the same extent as fully methylated tRNA. The tRNA from DL-ethionine-treated livers showed an overall decreased amino-acylation capacity for all amino acids tested. The 5-methylcytidine-deficient tRNA from DL-ethionine-treated mice were further characterized as substrates in homologous rate assays designed to determine the Km and V of the aminoacylation reaction using four individual 14C-labeled amino acids and a mixture of 14C-labeled amino acids. The Km and V of the reactions for all amino acids tested using 5-methylcytidine-deficient tRNA as substrate were essentially the same as for fully methylated tRNA. However, the Km and V were increased when liver tRNA from mice treated with DL-ethionine plus adenine was used as substrate in the rate reaction with [14C]lysine as label. Our results suggest that although extensively altered tRNA is a poorer substrate than control tRNA in both extent and rate of aminoacylation, 5-methylcytidine in mammalian tRNA is not involved in the recognition of the tRNA by the synthetase as measured by aminoacylation activity.  相似文献   

9.
Genetic transformation using Agrobacterium rhizogenes   总被引:1,自引:0,他引:1  
UDP-glucose pyrophosphorylase (EC 2.7.7.9) has been highly purified from the plant fraction of soybean ( Glycine max L. Merr. cv Williams) nodules. The purified enzyme gave a single polypeptide band following sodium docecyl sulphate polyacryla-mide gel electrophoresis, but was resolved into three bands of activity in non-denaturing gels. The enzyme appeared to be a monomer of molecular weight between 30 and 40 kDa. UDP-glucose pyrophosphorylase had optimum activity at pH 8.5 and displayed typical hyperbolic kinetics. The enzyme had a requirement for divalent metal ions, and was highly specific for the substrates pyrophosphate and UDP-glucose in the pyrophosphorolysis direction, and glucose-1-phosphate and UTP in the direction of UDP-glucose synthesis. The Km values were 0.19 m M and 0.07 m M for pyrophosphate and UDP-glucose, respectively, and 0.23 m M and 0.11 m M for glucose-1-phosphate and UTP. The maximum velocity in the pyrophosphorolysis direction was almost double that for the reverse reaction. UDP-glucose pyrophosphorylase did not appear to be subject to a high degree of fine control, and activity in vivo may be regulated mainly by the availability of the substrates.  相似文献   

10.
A total rate equation was used to calculate the discrimination of valine by the isoleucyl-tRNA synthetase from Escherichia coli. The PPi present in the cell makes the backward reaction or the pyrophosphorolysis of the E.aa-AMP possible. If the E.Ile-AMP has been corrected for wrong aminoacyl adenylation by the pretransfer proofreading, the pyrophosphorolysis rapidly equilibrates the corrected E.Ile-AMP with E.Ile and thus spoils the effect of the proofreading. The loss of the corrected species is avoided if there is a barrier (perhaps conformational) formed by a slow reaction step between the noncorrected E.Ile-AMP and the corrected (*E)tRNA(Ile-AMP). If such a slow conformational change exists, the increase in accuracy from the pretransfer proofreading would be beneficial, and, in addition, the PPi increases the accuracy by optimizing the initial discrimination of the wrong amino acid.  相似文献   

11.
A Srivastava  M J Modak 《Biochemistry》1980,19(14):3270-3275
Terminal deoxynucleotidyltransferase (TdT) has been found to catalyze both pyrophosphate exchange and pyrophosphorolysis reactions. Both reactions are strongly inhibited by antiserum to TdT. The reactions require the presence of a divalent cation, a single- or double-stranded oligomeric or polymeric DNA or RNA, and deoxyribonucleoside triphosphates (for PPi exchange only). Of the three divalent cations tested, Mg2+ and Co2+ are equally effective, while Mn2+ neither is used for catalysis nor inhibits the Mg2+-catalyzed reactions. Ribonucleoside triphosphates have been found to support the PPi exchange reaction to a minor extent and have no inhibitory effect on the catalysis mediated by dNTPs. Inhibition studies, using SH group inhibitors, Zn chelator, and a substrate binding site specific reagent, revealed that PPi exchange and pyrophosphorolysis reactions may be distinguished by differences in their sensitivity to inhibition by various reagents. While the PPi exchange reaction is strongly inhibited by sulfhydryl reagents, o-phenanthroline, and pyridoxal phosphate, the pyrophosphorolysis reaction is insensitive to these reagents. In addition, the pyrophosphorolysis reaction is also found not to require a free 3'-OH terminus of a primer. This difference in the susceptibility of the two reactions indicates that discrete active-site structures exist in TdT which catalyze PPi exchange and pyrophosphorolysis reactions.  相似文献   

12.
Fusarium moniliforme NCIM 1276 isolated from a tropical mangrove ecosystem produces a single extracellular endo-polygalacturonase with an M(r) of 38 kDa and a carbohydrate content of 4%. It has an alkaline pI of 8.1. The K(m) is 0.12 mg.mL(-1), V(max) is 111.1 micromol.min(-1).mg(-1) and the kcat is 4200 min-1. It has a pH optimum of 4.8. Kinetic and fluorescence data show that tryptophan is involved in binding. An arginine residue at or near the active site may be involved in extended binding of the substrate. A carboxylate and a histidine residue are involved in catalysis. These data are discussed with reference to current literature.  相似文献   

13.
Spermine and related polyamines have been reported to substitute for Mg2+ in the aminoacylation of tRNA catalyzed by aminoacyl-tRNA synthetases, but not in the ATP-PP-i exchange reaction. Such observations have led some workers to propose that these reactions proceed via a concerted mechanism rather than the usual two-step mechanism involving an aminoacyladenylate intermediate. In an attempt to elucidate the mechanism of the spermine effect on acylation and exchange, both reactions were re-examined using isoleucyl-tRNA synthetase from Escherichia coli. In the absence of added Mg2+ untreated tRNA was acylated in the presence of spermine, but tRNA from which Mg2+ had been scrupulously removed was not. ATP-PP-i exchange was not observed when spermine was used in place of Mg2+; however, if tRNA possessing sequestered Mg2+ was added, the exchange reaction was observed. These data suggest that a primary effect of spermine is to displace bound Mg2+ from tRNA in quantities sufficient to promote both the ATP-PP-i exchange and esterification of tRNA. The previously reported stimulatory effects of polyamines on these reactions are believed to be artifacts due to Mg2+ contamination of tRNA. Providing trace levels of Mg2+ are present, spermine exerts a secondary stimulation of the rate of aminoacylation, the mechanism of which is unknown. The results presented refute arguments that these enzymes proceed by a concerted mechansim and support the intermediacy of aminoacyladenylates.  相似文献   

14.
The hydrolysis and transphosphatidylation of lysophosphatidylcholine (LPC), with a partially purified preparation of phospholipase D (PL D) from Savoy cabbage, was investigated. These reactions were about 20 times slower than the hydrolysis of phosphatidylcholine (PC) in a micellar system. For the transfer reaction, 2 M glycerol was included in the media, which suppressed the hydrolytic reaction. Both reactions presented similar V(max) values, suggesting that the formation of the phosphatidyl-enzyme intermediate is the rate-limiting step. The enzyme had an absolute requirement for Ca(2+), and the optimum concentration was approximately 40 mM CaCl(2). K(Ca)(app) was calculated to be 8.6+/-0.74 mM for the hydrolytic and 10+/-0.97 mM for the transphosphatidylation reaction. Both activities reached a maximum at pH 5.5, independent of Ca(2+) concentration. Kinetic studies showed that the Km(app) for the glycerol in the transphosphatidylation reaction is 388+/-37 mM. Km(app) for the lysophosphatidylcholine depended on Ca(2+) concentration and fell between 1 and 3 mM at CaCl(2) concentrations from 4 to 40 mM. SDS, TX-100, and CTAB did not activate the enzyme as reported for phosphatidylcholine hydrolysis; on the contrary, reaction rates decreased at detergent concentrations at or above that of lysophosphatidylcholine.  相似文献   

15.
J Pimmer  E Holler 《Biochemistry》1979,18(17):3714-3723
The association of phenylalanylptRNA and Mg2+ follows a biphasic concentration dependence as indicated by the active site directed fluorescent indicator 2-p-toluidinyl-naphthalene-6-sulfonate. The macroscopic dissociation constants are 0.16 +/- 0.03 and 4.1 +/- mM. The effect of Mg2+ on the association of enzyme and MgATP, on the synergistic binding of MgATP and L-phenylalaninol, and on the pre-steady-state synthesis and pyrophosphorolysis of the enzyme-phenylalanyladenylate complex in the absence and the presence of tRNA Phe has been measured by established equilibrium and stopped-flow techniques using 2-p-toluidinylnaphthalene-6-sulfonate. At 10 mM Mg2+, the association of enzyme and MgATP is biphasic with dissociation constants of 0.25 +/- 0.03 and 9.1 +/- 1.7 mM. At 2 mM Mg2+, a single dissociation constant of 5.0 +/- 0.5 mM is indicated. The coupling constant of the synergistic reaction is 15 at 1 mM Mg2+ and 290 at 10 mM Mg2+. The Hill constant of the sigmoidal dependence is 3.6. The strengthening of the synergism is believed to reflect a Mg2+-dependent coupling of the synergistic reactions at the two active sites of the enzyme, the coupling being negligible at 1 mM and maximal at 10 mM Mg2+. The pre-steady-state rate of adenylate synthesis is accelerated by the presence of Mg2+. The effect is to decrease the value of the Michaelis-Menten constant of MgATP. Another effect is to increase the rate constant when tRNA Phe is present. At subsaturating [MgATP], the [Mg2+] dependence of the observed rate constant is hyperbolical in the absence and sigmoidal (Hill constant, 3.5) in the presence of tRNA Phe. The rate of the pyrophosphorolysis is enhanced by a decrease of the Michaelis-Menten constant of MgPPi. The effects on the thermodynamics and kinetics parallel the occupancy of the low-affinity Mg2+-binding sites of the enzyme.  相似文献   

16.
Composition variation of a complex peptide mixture under enzymatic transformation can be tracked by mass spectrometry (MS). In this report, papain-catalyzed esterification of fibroin peptides was investigated at the individual peptide level using liquid chromatography-mass spectrometry with selected ion monitoring. Optimal conditions for maximizing ester formation were obtained using a water-to-pentanol ratio of 1:9 at pH 2.8 and 40°C; however, the optimum conditions varied for individual peptides. The optimum pH levels were 2.5 and 2.8 for the tetrapeptides with a tyrosine or a valine residue and those with alanine or serine residues, respectively. The optimum pH shifted to 3.4 for dipeptide esters with a tyrosine residue. Tetrapeptides had a relatively higher rate of esterification above 50°C. Alhough, the profiles of peptides and their esters in the esterification reaction were significantly affected by the reaction conditions, alanyl-glycine ester represented the largest fraction in the mixture under most reaction conditions. As demonstrated here, MS analysis of peptide mixtures can be used to elucidate specific reaction conditions for the enrichment of particular peptide products.  相似文献   

17.
Sphingolipid ceramide N-deacylase catalyzes a reversible reaction in which the amide linkages of the ceramides of various sphingolipids are cleaved or synthesized. Hydrolysis of sphingolipids by the enzyme proceeded efficiently at acidic pH in the presence of high concentrations of detergents, whereas the reverse reaction tended to be favored at neutral pH with a decrease in the detergent concentration. Although the catalytic efficiency (V(max)/K(m)) of the hydrolysis and reverse reactions was changed mainly by the concentration of detergents in the reaction mixture, V(max) and K(m) for the reverse reaction were relatively higher than those for the forward reaction, irrespective of the detergent concentration. The reverse reaction proceeded most efficiently when the molar ratio of lyso-sphingolipids and fatty acids was fixed at 1 : 1-2, the yield of the reaction exceeding 70-80%. The reverse and exchange (transacylation) reactions did not require ATP, CoA, metal ions or addition of organic solvents. Studies using inhibitors and chemical modifiers of the enzyme protein suggested that both the hydrolysis and condensation reactions are catalyzed at the same catalytic domain. These results indicate that the reverse hydrolysis reaction of the enzyme is unique, being completely different from those of lipases, proteases and glycosidases reported to date.  相似文献   

18.
C Grubmeyer  H Teng 《Biochemistry》1999,38(22):7355-7362
L-Histidinol dehydrogenase catalyzes the biosynthetic oxidation of L-histidinol to L-histidine with sequential reduction of two molecules of NAD. Previous isotope exchange results had suggested that the oxidation of histidinol to the intermediate histidinaldehyde occurred 2-3-fold more rapidly than overall catalysis. In this work, we present kinetic isotope effects (KIE) studies at pH 9.0 and at pH 6.7 with stereospecifically mono- and dideuterated histidinols. The data at pH 9.0 support minimal participation of the first hydride transfer and substantial participation of the second hydride transfer in the overall rate limitation. Stopped-flow experiments with protiated histidinol revealed a small burst of NADH production with stoichiometry of 0.12 per subunit, and 0.25 per subunit with dideuterated histidinol, indicating that the overall first half-reaction was not significantly faster than the second reaction sequence. Results from kcat and kcat/KM titrations with histidinol, NAD, and the alternative substrate imidazolyl propanediol demonstrated an essential base with pKa values between 7.7 and 8.4. In KIE experiments performed at pH 6.7 or with a coenzyme analogue at pH 9. 0, the first hydride transfer became more rate limiting. Kinetic simulations based on rate constants estimated from this work fit well with a mechanism that includes a relatively fast, and thermodynamically unfavorable, hydride transfer from histidinol and a slower, irreversible second hydride transfer from a histidinaldehyde derivative. Thus, although the chemistry of the first hydride transfer is fast, both partial reactions participate in the overall rate limitation.  相似文献   

19.
Histone PTMs play a crucial role in regulating chromatin structure and function, with impact on gene expression. MS is nowadays widely applied to study histone PTMs systematically. Because histones are rich in arginine and lysine, classical shot‐gun approaches based on trypsin digestion are typically not employed for histone modifications mapping. Instead, different protocols of chemical derivatization of lysines in combination with trypsin have been implemented to obtain “Arg‐C like” digestion products that are more suitable for LC‐MS/MS analysis. Although widespread, these strategies have been recently described to cause various side reactions that result in chemical modifications prone to be misinterpreted as native histone marks. These artefacts can also interfere with the quantification process, causing errors in histone PTMs profiling. The work of Paternoster V. et al. 1 is a quantitative assessment of methyl‐esterification and other side reactions occurring on histones after chemical derivatization of lysines with propionic anhydride [Proteomics 2016, 16, 2059–2063]. The authors estimate the effect of different solvents, incubation times, and pH on the extent of these side reactions. The results collected indicate that the replacement of methanol with isopropanol or ACN not only blocks methyl‐esterification, but also significantly reduces other undesired unspecific reactions. Carefully titrating the pH after propionic anhydride addition is another way to keep methyl‐esterification under control. Overall, the authors describe a set of experimental conditions that allow reducing the generation of various artefacts during histone propionylation.  相似文献   

20.
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