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 31 毫秒
1.
The newly introduced sulfhydryl reductant tris(2-carboxyethyl)phosphine (TCEP) is a potentially attractive alternative to commonly used dithiothreitol (DTT). We compare properties of DTT and TCEP important in protein biochemistry, using the motor enzyme myosin as an example protein. The reductants equally preserve myosin's enzymatic activity, which is sensitive to sulfhydryl oxidation. When labeling with extrinsic probes, DTT inhibits maleimide attachment to myosin and must be removed before labeling. In contrast, maleimide attachment to myosin was achieved in the presence of TCEP, although with less efficiency than no reductant. Surprisingly, iodoacetamide attachment to myosin was nearly unaffected by either reductant at low (0.1 mM) concentrations. In electron paramagnetic resonance (EPR) spectroscopy utilizing nitroxide spin labels, TCEP is highly advantageous: spin labels are two to four times more stable in TCEP than DTT, thereby alleviating a long-standing problem in EPR. During protein purification, Ni(2+) concentrations contaminating proteins eluted from Ni(2+) affinity columns cause rapid oxidation of DTT without affecting TCEP. For long-term storage of proteins, TCEP is significantly more stable than DTT without metal chelates such as EGTA in the buffer, whereas DTT is more stable if metal chelates are present. Thus TCEP has advantages over DTT, although the choice of reductant is application specific.  相似文献   

2.
Vâlcu CM  Schlink K 《Proteomics》2006,6(5):1599-1605
Protein extraction procedure and the reducing agent content (DTT, dithioerythritol, tributyl phosphine and tris (2-carboxyethyl) phosphine (TCEP)) of the sample and rehydration buffers were optimised for European beech leaves and roots and Norway spruce needles. Optimal extraction was achieved with 100 mM DTT for leaves and needles and a mixture of 2 mM TCEP and 50 mM DTT for roots. Performing IEF in buffers containing hydroxyethyldisulphide significantly enhanced the quality of separation for all proteins except for acidic root proteins, which were optimally focused in the same buffer as extracted.  相似文献   

3.
Determination of dehydroascorbic acid in biological samples most commonly involves indirect measurement. The concentration is calculated by subtraction of the measured ascorbic acid concentration from that of total ascorbic acid analyzed after reduction of the dehydroascorbic acid present; a methodology also referred to as subtraction methods. Consequently, successful determination of dehydroascorbic acid is dependent on proper sample handling, quantitative reduction of the compound, and accurate quantification of both ascorbic acid and total ascorbic acid. In this paper, the recently introduced reductant tris[2-carboxyethyl]phosphine (TCEP) is evaluated as a reliable alternative to the commonly used reducing agent dithiothreitol (DTT). The results show that TCEP offers a more efficient reduction of dehydroascorbic acid at low pH compared to that of DTT. Moreover, while DTT maintains a reducing sample environment for less than 24 h, TCEP show complete protection from oxidation of ascorbic acid for at least 96 h following sample preparation. Removal of TCEP prior to analysis is unnecessary. A revised HPLC-EC method incorporating TCEP as reductant as well as the coanalysis of isoascorbic acid and uric acid is presented. The within- and between-day coefficients of variation for the complete assay are less than 1.5 and 3.5% for all analytes. As a whole, the method presented here is simpler and more reliable than existing methods.  相似文献   

4.
G A White  T Wood    W E Hill 《Nucleic acids research》1988,16(22):10817-10831
The cause of 50S ribosomal subunit collapse reportedly triggered by hybridization of a 14-base cDNA probe to the alpha-sarcin region of 23S rRNA was investigated by physical measurement of probe-subunit complexes in varying buffer conditions. The results reported here show that this probe was unable to hybridize to its target site in the intact 50S subunit and the physical characteristics of 50S subunits remained unchanged in its presence. Subunit collapse was induced in buffer containing 20mM Tris-HCl (pH 7.5), 600 mM NH4Cl, 1 mM MgCl2, 1 mM DTT, and 0.1 mM EDTA in the absence of probe. The probe bound specifically to its target site in the collapsed particle, but did not promote further unfolding. The results demonstrate that a DNA probe bound to the alpha-sarcin region cannot cause the 50S subunit to unfold or cause 23S rRNA to degrade. We suggest that the previously reported collapse was most probably the result of the ionic conditions used.  相似文献   

5.
1. The pH optimum of CO2-dependent O2 evolution by barley (Hordeum vulgare L.) chloroplasts was found to be between 7.8 and 8.2. The addition of 1 mM MgCl2 in the dark inhibited O2 evolution over the entire pH range tested and resulted in a much sharper pH profile centered around pH 8.2. 2. The pH optimum for O2 evolution, in the presence and absence of 1 mM MgCl2, was acid-shifted 0.3--0.4 pH units by 2 mM NH4Cl. The pH optimum of O2 evolution, with and without 1 mM MgCl2, was base-shifted by 2 mM sodium acetate, approx. 0.5 pH units relative to the controls. 3. O2 evolution in the presence of bicarbonate plus 3-phosphoglycerate or ribose-5-phosphate was considerably less sensitive to pH than CO2-dependent O2 evolution in the absence of substrate. With these substrates, both in the presence and absence of 1 mM MgCl2, the pH optimum was broad and was centered around pH 7.8. 4. Inhibition of CO2-dependent O2 evolution by inorganic phosphate and magnesium increased as the pH of the reaction mixture was decreased below the optimum. Decreasing the pH from 8.2 to 7.6, reduced over 3-fold the concentration of inorganic phosphate required to inhibit O2 evolution completely. For magnesium, a similar change in pH reduced the concentration required to inhibit O2 evolution 50% approx. 5-fold. At pH 8.2, magnesium inhibition required inorganic phosphate. Magnesium was not required for inhibition of O2 evolution by inorganic phosphate, but incresaed the relative inhibition observed. 5. Illumination of intact barley chloroplasts increased the activity of NADP-glyceraldehyde-3-P dehydrogenase, phosphoribulokinase and fructose-1,6-diphosphatase. MgCl2 and inorganic phosphate prevented this increase in enzyme activity at concentrations that completely inhibited CO2-dependent O2 evolution. 6. The results obtained suggest that magnesium inhibition of O2 evolution may be caused by enhanced phosphate exchange across the chloroplast envelope.  相似文献   

6.
Sopina VA 《Tsitologiia》2002,44(11):1120-1128
Tartrate-resistant acid phosphatase (TRAP) of Amoeba proteus (strain B) was represented by 3 of 6 bands (= electromorphs) revealed after disc-electrophoresis in polyacrylamide gels with the use of 2-naphthyl phosphate as a substrate at pH 4.0. The presence of MgCl2, CaCl2 or ZnCl2 (50 mM) in the incubation mixture used for gel staining stimulated activities of all 3 TRAP electromorphs or of two of them (in the case of ZnCl2). When gels were treated with MgCl2, CaCl2 or ZnCl2 (10 and 100 mM, 30 min) before their staining activity of TRAP electromorphs also increased. But unlike 1 M MgCl2 or 1 M CaCl2, 1 M ZnCl2 partly inactivated two of the three TRAP electromorphs. EDTA and EGTA (5 mM), and H2O2 (10 mM) completely inhibited TRAP electromorphs after gel treatment for 10, 20 and 30 min, resp. Of 5 tested ions (Mg2+, Ca2+, Fe2+, Fe3+ and Zn2+), only the latter reactivated the TRAP electromorphs previously inactivated by EDTA or EGTA treatment. In addition, after EDTA inactivation, TRAP electromorphs were reactivated better than after EGTA. The resistance of TRAP electromorphs to okadaic acid and phosphatase inhibitor cocktail 1 used in different concentrations is indicative of the absence of PP1 and PP2A among these electromorphs. Mg2+, Ca2+ and Zn2+ dependence of TRAP activity, and the resistance of its electromorphs to vanadate and phosphatase inhibitor cocktail 2 prevents these electromorphs from being classified as PTP. It is suggested that the active center of A. proteus TRAP contains zinc ion, which is essential for catalytic activity of the enzyme. Thus, TRAP of these amoebae is metallophosphatase showing phosphomonoesterase activity in acidic medium. This metalloenzyme differs from both mammalian tartrate-resistant PAPs and tartrate-resistant metallophosphatase of Rana esculenta.  相似文献   

7.
The dissociation of apo- and metal-bound human copper-zinc superoxide dismutase (SOD1) dimers induced by the chaotrope guanidine hydrochloride (GdnHCl) or the reductant Tris(2-carboxyethyl)phosphine (TCEP) has been analyzed using analytical ultracentrifugation. Global fitting of sedimentation equilibrium data under native solution conditions (without GdnHCl or TCEP) demonstrate that both the apo- and metal-bound forms of SOD1 are stable dimers. Sedimentation velocity experiments show that apo-SOD1 dimers dissociate cooperatively over the range 0.5-1.0 M GdnHCl. In contrast, metal-bound SOD1 dimers possess a more compact shape and dissociate at significantly higher GdnHCl concentrations (2.0-3.0 M). Reduction of the intrasubunit disulfide bond within each SOD1 subunit by 5-10 mM TCEP promotes dissociation of apo-SOD1 dimers, whereas the metal-bound enzyme remains a stable dimer under these conditions. The Cys-57 --> Ser mutant of SOD1, a protein incapable of forming the intrasubunit disulfide bond, sediments as a monomer in the absence of metal ions and as a dimer when metals are bound. Taken together, these data indicate that the stability imparted to the human SOD1 dimer by metal binding and the formation of the intrasubunit disulfide bond are mediated by independent molecular mechanisms. By combining the sedimentation data with previous crystallographic results, a molecular explanation is provided for the existence of different SOD1 macromolecular shapes and multiple SOD1 dimeric species with different stabilities.  相似文献   

8.
Tris(2-carboxyethyl)phosphine (TCEP) is a widely used substitute for dithiothreitol (DTT) in the reduction of disulfide bonds in biochemical systems. Although TCEP has been recently shown to be a substrate of the flavin-dependent sulfhydryl oxidases, there is little quantitative information concerning the rate by which TCEP reduces other peptidic disulfide bonds. In this study, mono-, di-, and trimethyl ester analogues of TCEP were synthesized to evaluate the role of carboxylate anions in the reduction mechanism, and to expand the range of phosphine reductants. The effectiveness of all four phosphines relative to DTT has been determined using model disulfides, including a fluorescent disulfide-containing peptide (H(3)N(+)-VTWCGACKM-NH(2)), and with protein disulfide bonds in thioredoxin and sulfhydryl oxidase. Mono-, di-, and trimethyl esters exhibit phosphorus pK values of 6.8, 5.8, and 4.7, respectively, extending their reactivity with the model peptide to correspondingly lower pH values relative to that of TCEP (pK = 7.6). At pH 5.0, the order of reactivity is as follows: trimethyl- > dimethyl- > monomethyl- > TCEP > DTT; tmTCEP is 35-fold more reactive than TCEP, and DTT is essentially unreactive. Esterification also increases lipophilicity, allowing tmTCEP to penetrate phospholipid bilayers rapidly (>30-fold faster than DTT), whereas the parent TCEP is impermeant. Although more reactive than DTT toward small-molecule disulfides at pH 7.5, all phosphines are markedly less reactive toward protein disulfides at this pH. Molecular modeling suggests that the nucleophilic phosphorus of TCEP is more sterically crowded than the thiolate of DTT, contributing to the lower reactivity of the phosphine with protein disulfides. In sum, these data suggest that there is considerable scope for the synthesis of phosphine analogues tailored for specific applications in biological systems.  相似文献   

9.
The sequence requirements for self-cleavage of hepatitis delta virus genomic RNA were examined using precursor RNAs which were labeled at either the 5' or 3' ends and progressively deleted from the unlabeled end. In the presence of 50% formamide, which enhances self-cleavage in 2 mM MgCl2 at 37 degrees C, 84 nucleotides (nt) 3' of the break site were required. In the absence of formamide the minimum was reduced to 82 nt. Under both sets of conditions, precursors with 1 nt 5' to the break site cleaved. These results allowed two condition-dependent minimal domains for self-cleavage to be defined. However, in the absence of formamide, sequences flanking the minimal domain inhibited cleavage, possibly through involvement in the formation of non-cleaving structures. These data are consistent with the idea that cleavage in vivo could be regulated by alternative RNA structures.  相似文献   

10.
The disulfide reducing agent, dithiothreitol (DTT) and the sulfhydryl-modifying reagents p-chloromercuribenzenesulfonic acid and N-ethylmaleimide (NEM) were employed to assess the role of disulfide and sulfhydryl groups in organic cation transport. The transport of N1-[3H]methylnicotinamide (NMN), a prototypic organic cation, was examined employing brush-border membrane vesicles isolated from the outer cortex of canine kidneys. DTT inhibited NMN transport reversibly with an IC50 of 250 microM/mg of protein. 5 mM NMN protected against DTT inactivation. The specificity of substrate protection was demonstrated by showing that D-glucose had no effect on the DTT inactivation of NMN transport and conversely that NMN had no effect on the DTT inactivation of D-glucose transport. Disulfide bonds reduced by DTT could be reoxidized by washing with excess buffer or by addition of 0.02% H2O2 thereby restoring NMN transport. p-Chloromercuribenzenesulfonic acid reversibly inactivated NMN transport with an IC50 of 25 microM/mg of protein. 5mM NMN protected against inactivation. NEM irreversibly inactivated transport with an IC50 of 250 microM/mg of protein. The rate of NMN inactivation by NEM followed pseudo-first order reaction kinetics. A replot of the data gave a linear relationship between the apparent rate constants and the NEM concentration with a slope of 1.3. The data are consistent with a simple bimolecular reaction mechanism and imply that one molecule of NEM inactivates 1 sulfhydryl group/active transport unit. The presence of 5 mM NMN affected the rate of NEM (2.5 mM) inactivation: the t1/2 values for inactivation in the presence and absence of substrate were 7.3 and 2.0 min, respectively. The results demonstrate an essential requirement for disulfide and sulfhydryl groups.  相似文献   

11.
12.
Cells of Staphylococcus aureus heated at 52 degrees C in magnesium-chelating buffers [pH 7.2, 50 mM potassium phosphate or 50 mM tris(hydroxymethyl)-aminomethane containing 1 mM ethylenediaminetetraacetic acid] leaked 260-nm absorbing material, shown to be RNA, and suffered destruction of their ribosomes. These cells did not regain their salt tolerance when repair was carried out in the presence of actinomycin D (5 microgram/ml). Cells similarly heated in magnesium-conserving buffers [pH 7.2, 50 mM tris(hydroxymethyl)aminomethane containing 10 mM MgCl2 or piperazine buffer] did not leak RNA, suffered no ribosomal damage when heated for 15 min, and recovered, at least partially, in the presence of actinomycin D. Ribosomal damage, is therefore, a consequence of Mg2+ loss and is not an effect of heat per se. Cells suspended in either Mg2+-chelating or Mg2+-conserving buffers lost salt tolerance to about the same extent during heating at 52 degrees C. Therefore, sublethal heat injury can not be attributed to ribosomal damage.  相似文献   

13.
The F1 and F1-inhibitor-protein complex synthesized tightly bound ATP from ADP and Pi when the organic solvents dimethylsulfoxide (20-50% v/v), ethylene glycol (20-60% v/v) or poly(ethylene glycol) 4000 and 8000 (30-50% w/v) were included in the assay media. There was no synthesis of tightly bound ATP in the absence of organic solvents. In the presence of 50% dimethylsulfoxide, maximal synthesis of ATP was obtained at pH values between 6.5 and 7.7. In both F1 and F1-inhibitor-protein there was no synthesis of ATP in the absence of MgCl2. The rate of ATP synthesis became faster as the MgCl2 concentration in the medium was raised from 0.1-10 mM. The Km for Pi of F1 was in the range of 0.8-1.5 mM. The Km for Pi of the F1-inhibitor-protein was much higher than that of F1 and could not be measured. In the presence of 10 mM MgCl2 and 2 mM Pi, the rate constants of ATP synthesis by F1 and F1-inhibitor-protein were 5.2-10.4 h-1 and 3.5-5.9 h-1 respectively. For both enzymes the rate constant of ATP hydrolysis was 0.69 h-1. The tightly bound ATP of F1 and F1-inhibitor-protein were hydrolyzed at a much slower rate when either the Pi concentration or the MgCl2 concentration was suddenly decreased. Both in presence and absence of Mg2+, 40-60% of the radioactive tightly bound ATP synthesized by F1 was hydrolyzed when non-radioactive ATP was added to the assay medium. This was not observed when F1-inhibitor-protein was used.  相似文献   

14.
Similarities in properties of ribulose diphosphate carboxylase and oxygenase activities further substantiate the hypothesis that the same protein catalyzes both reactions. The Km (ribulose diphosphate) is 0.33 mM for the ribulose diphosphate oxygenase, when assayed in air with an oxygen electrode. Maximum activity is obtained with 10 to 35 mM MgCl2. Higher MgCl2 concentrations are inhibitory, but they shift the pH optimum from 9.3 or 9.4 to 8.7 or 9.0. MnCl2 is an effective cofactor of the oxygenase and some activity is obtained with CoCl2. Both the ribulose diphosphate carboxylase and oxygenase activity of the purified protein from spinach leaves are slowly inactivated by storage at 0 degrees and reactivated in 10 min at 50 degrees, provided both 25 mM MgCl2 and 1 mM dithiothreitol are present. The sulfhydryl groups of the enzyme which react rapidly with 5,5'-dithiobis(2-nitrobenzoic acid) are approximately 4 at pH 7.8 and 11 at pH 9.4. At both pH values ribulose diphosphate prevents two of these sulfhydryl groups from reacting with this reagent. About 50% inhibition of the oxygenase activity at pH 9.0 occurs with 50 mM bicarbonate in the presence of 3 mM ribulose diphosphate, and from variations in these parameters the inhibition is attributed to the CO2 species. The purified enzyme of acrylamide gels prevented the reduction of nitroblue tetrazolium in the presence of the superoxide radical, but the enzyme in solution did not react as a superoxide dismutase.  相似文献   

15.
We have used liposomes with incorporated pig kidney Na+,K(+)-ATPase to study vanadate sensitive K(+)-K+ exchange and net K+ uptake under conditions of acetyl- and p-nitrophenyl phosphatase activities. The experiments were performed at 20 degrees C. Cytoplasmic phosphate contamination was minimized with a phosphate trapping system based on glycogen, phosphorylase a and glucose-6-phosphate dehydrogenase. In the absence of Mg2+ (no phosphatase activity) 5-10 mM p-nitrophenyl phosphate slightly stimulated K(+)-K+ exchange whereas 5-10 mM acetyl phosphate did not. In the presence of 3 mM MgCl2 (high rate of phosphatase activity) acetyl phosphate did not affect K(+)-K+ exchange whereas p-nitrophenyl phosphate induced a greater stimulation than in the absence of Mg2+; a further addition of 1 mM ADP resulted in a 35-65% inhibition of phosphatase activity with an increase in K(+)-K+ exchange, which sometimes reached the levels seen with 5 mM phosphate and 1 mM ADP. The net K+ uptake in the presence of 3 mM MgCl2 was not affected by acetyl phosphate or p-nitrophenyl phosphate, whereas it was inhibited by 5 mM phosphate (with and without 1 mM ADP). The results of this work suggest that the phosphatase reaction is not by itself associated to K+ translocation. The ADP-dependent stimulation of K(+)-K+ exchange in the presence of phosphatase activity could be explained by the overlapping of one or more step/s of the reversible phosphorylation from phosphate with the phosphatase cycle.  相似文献   

16.
S J Li  A G Marshall 《Biochemistry》1985,24(15):4047-4052
Unfolding of purified wheat germ ribosomal RNA has been studied by differential scanning calorimetry (DSC) from 15 to 95 degrees C, in the presence and absence of 100 mM NaCl and/or 10 mM MgCl2. The total enthalpy of melting varies from about 290 (no sodium or magnesium) to 480 kcal/mol (Mg2+ only) and precisely accounts for the number and types of base pairs deduced from prior Fourier-transform infrared experiments. The composite DSC curves are analyzed into four individual two-state melting stages. Both Na+ and Mg2+ shift the melting transitions to higher temperature; in addition, Mg2+ causes the individual transitions to merge (i.e., melt cooperatively) and leads to irreversible chain cleavage at high temperature. The results are analyzed according to three general secondary base-pairing models for eukaryotic 5S RNA.  相似文献   

17.
1. The higher relative molecular mass (M(r)) forms of larval honeybee haemolymph alpha-glucosidases are dissociated by dithiothreitol (DTT) into lower M(r) electrophoretic forms, without any important loss of activity. 2. The maximum velocity remains unchanged and the apparent dissociation constant is slightly increased, with Ki approximately equal to 247 mM and I50 approximately equal to 730 mM. 3. By contrast, the major changes affect the Hill coefficient which decreases from 1.0 in controls to 0.7 in presence of 600 mM DTT. 4. In the absence of both DTT and substrate, the major native enzyme form, isolated by gel filtration, spontaneously rearranges to give three additional minor forms, one of lower M(r) and two of higher M(r). 5. These data are consistent with the hypothesis of substrate-directed aggregation of enzyme protomers into functional complexes.  相似文献   

18.
Adsorption of DNA to sand and variable degradation rates of adsorbed DNA   总被引:11,自引:0,他引:11  
Adsorption and desorption of DNA and degradation of adsorbed DNA by DNase I were studied by using a flowthrough system of sand-filled glass columns. Maximum adsorption at 23 degrees C occurred within 2 h. The amounts of DNA which adsorbed to sand increased with the salt concentration (0.1 to 4 M NaCl and 1 mM to 0.2 M MgCl2), salt valency (Na+ less than Mg2+ and Ca2+), and pH (5 to 9). Maximum desorption of DNA from sand (43 to 59%) was achieved when columns were eluted with NaPO4 and NaCl for 6 h or with EDTA for 1 h. DNA did not desorb in the presence of detergents. It is concluded that adsorption proceeded by physical and chemical (Mg2+ bridging) interaction between the DNA and sand surfaces. Degradability by DNase I decreased upon adsorption of transforming DNA. When DNA adsorbed in the presence of 50 mM MgCl2, the degradation rate was higher than when it adsorbed in the presence of 20 mM MgCl2. The sensitivity to degradation of DNA adsorbed to sand at 50 mM MgCl2 decreased when the columns were eluted with 0.1 mM MgCl2 or 100 mM EDTA before application of DNase I. This indicates that at least two types of DNA-sand complexes with different accessibilities of adsorbed DNA to DNase I existed. The degradability of DNA adsorbed to minor mineral fractions (feldspar and heavy minerals) of the sand differed from that of quartz-adsorbed DNA.  相似文献   

19.
Live Trypanosoma cruzi amastigotes hydrolyzed p-nitrophenylphosphate (PNPP), phospho-amino-acids and 32P-casein under physiologically appropriate conditions. PNPP was hydrolysed at a rate of 80 nmol.mg-1.h-1 in the presence of 5 mM MgCl2, pH 7.2 at 30 degrees C. In the absence of Mg2+ the activity was reduced 40% and we call this basal activity. At saturating concentration of PNPP, half-maximal PNPP hydrolysis was obtained with 0.22 mM MgCl2. Ca2+ had no effect on the basal activity, could not substitute Mg2+ as an activator and in contrast inhibited the PNPP hydrolysis stimulated by Mg2+ (I50 = 0.43 mM). In the absence of Mg2+ (basal activity) the stimulating half concentration (S0.5) for PNPP was 1.57 mM, while at saturating MgCl2 concentrations the corresponding S0.5 for PNPP for Mg(2+)-stimulated phosphatase activity (difference between total minus basal phosphatase activity) was 0.99 mM. The Mg-dependent PNPP hydrolysis was strongly inhibited by sodium fluoride (NaF), vanadate and Zn2+ but not by tartrate and levamizole. The Mg-independent basal phosphatase activity was insensitive to tartrate, levamizole as well NaF and less inhibited by vanadate and Zn2+. Intact amastigotes were also able to hydrolyse phosphoserine, phosphothreonine and phosphotyrosine but only the phosphotyrosine hydrolysis was stimulated by MgCl2 and inhibited by CaCl2 and phosphotyrosine was a competitive inhibitor of the PNPP hydrolysis stimulated by Mg2+. The cells were also able to hydrolyse 32P-casein phosphorylated on serine and threonine residues but only in the presence of MgCl2. These results indicate that in the amastigote form of T. cruzi there are at least two ectophosphatase activities, one of which is Mg2+ dependent and can dephosphorylate phospho-amino acids and phosphoproteins under physiological conditions.  相似文献   

20.
Spin labels attached to rabbit muscle actin became more immobilized upon conversion of actin from the G state to the F state with 50 mM KCl. Titration of G-actin with MgCl2 produced F-actin-like EPR spectra between 2 and 5 mM-actin filaments by electron microscopy. Higher concentrations of MgCl2 produced bundles of actin and eventually paracrystals, accompanied by further immobilization of spin labels. The effects of MgCl2 and KCl were competitive: addition of MgCl2 to 50 mM could convert F-actin (50 mM KCl) to paracrystalline (P) actin; the reverse titration (0 to 200 mM KCl in the presence of 20 mM MgCl2) was less complete. Addition of DNase I to G- or F-actin gave the expected amorphous electron micrographic pattern, and the actin was not sedimentable at (400,000 x g x h). EPR showed that the actin was in the G conformation. Addition of DNase I to paracrystalline actin gave the F conformation (EPR) but the actin was "G" by electron microscopy. Phalloidin converted G-actin to F-actin, had no effect on F-actin, and converted P-actin to the F state by electron microscopy but maintained the P conformation by EPR. Cytochalasin B produced no effects observable by EPR or centrifugation but "untwisted" paracrystals into nets. Since actin retained its P conformation by EPR in two states which were morphologically not P, we conclude that the P state is a distinct conformation of the actin molecule and that actin filaments aggregate to form bundles (and eventually paracrystals) when actin monomers are able to enter the P conformation.  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