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 net charge and isoelectric pH (pI) of a protein depend on the content of ionizable groups and their pK values. Ribonuclease Sa (RNase Sa) is an acidic protein with a pI = 3.5 that contains no Lys residues. By replacing Asp and Glu residues on the surface of RNase Sa with Lys residues, we have created a 3K variant (D1K, D17K, E41K) with a pI = 6.4 and a 5K variant (3K + D25K, E74K) with a pI = 10.2. We show that pI values estimated using pK values based on model compound data can be in error by >1 pH unit, and suggest how the estimation can be improved. For RNase Sa and the 3K and 5K variants, the solubility, activity, and stability have been measured as a function of pH. We find that the pH of minimum solubility varies with the pI of the protein, but that the pH of maximum activity and the pH of maximum stability do not.  相似文献   

2.
K(Ca) channels are involved in control of cell proliferation and differentiation. Here we have revealed their role in overcoming the RNase-induced cytotoxicity. Toxic effects of Streptomyces aureofaciens RNases Sa, Sa2, Sa3, and of RNase Sa charge reversal mutants on the human embryonic kidney cell lines differing only by the presence of K(Ca) channels were characterized. In contrast to other RNases, a basic variant of RNase Sa and RNase Sa3 exhibit significant cytotoxic activity of the same order of magnitude as onconase. Our data indicate the absence of a correlation between catalytic activity and stability of RNases and cytotoxicity. On the other hand, cationization enhances toxic effect of an RNase indicating the major role of a positive charge. Essentially lower sensitivity to cytotoxic microbial RNases of cells expressing K(Ca) channels was found. These results suggest that cells without the K(Ca) channel activity cannot counteract toxic effect of RNases.  相似文献   

3.
The pK values of the titratable groups in ribonuclease Sa (RNase Sa) (pI=3.5), and a charge-reversed variant with five carboxyl to lysine substitutions, 5K RNase Sa (pI=10.2), have been determined by NMR at 20 degrees C in 0.1M NaCl. In RNase Sa, 18 pK values and in 5K, 11 pK values were measured. The carboxyl group of Asp33, which is buried and forms three intramolecular hydrogen bonds in RNase Sa, has the lowest pK (2.4), whereas Asp79, which is also buried but does not form hydrogen bonds, has the most elevated pK (7.4). These results highlight the importance of desolvation and charge-dipole interactions in perturbing pK values of buried groups. Alkaline titration revealed that the terminal amine of RNase Sa and all eight tyrosine residues have significantly increased pK values relative to model compounds.A primary objective in this study was to investigate the influence of charge-charge interactions on the pK values by comparing results from RNase Sa with those from the 5K variant. The solution structures of the two proteins are very similar as revealed by NMR and other spectroscopic data, with only small changes at the N terminus and in the alpha-helix. Consequently, the ionizable groups will have similar environments in the two variants and desolvation and charge-dipole interactions will have comparable effects on the pK values of both. Their pK differences, therefore, are expected to be chiefly due to the different charge-charge interactions. As anticipated from its higher net charge, all measured pK values in 5K RNase are lowered relative to wild-type RNase Sa, with the largest decrease being 2.2 pH units for Glu14. The pK differences (pK(Sa)-pK(5K)) calculated using a simple model based on Coulomb's Law and a dielectric constant of 45 agree well with the experimental values. This demonstrates that the pK differences between wild-type and 5K RNase Sa are mainly due to changes in the electrostatic interactions between the ionizable groups. pK values calculated using Coulomb's Law also showed a good correlation (R=0.83) with experimental values. The more complex model based on a finite-difference solution to the Poisson-Boltzmann equation, which considers desolvation and charge-dipole interactions in addition to charge-charge interactions, was also used to calculate pK values. Surprisingly, these values are more poorly correlated (R=0.65) with the values from experiment. Taken together, the results are evidence that charge-charge interactions are the chief perturbant of the pK values of ionizable groups on the protein surface, which is where the majority of the ionizable groups are positioned in proteins.  相似文献   

4.
The primary goal of this study was to gain a better understanding of the effect of environment and ionic strength on the pK values of histidine residues in proteins. The salt-dependence of pK values for two histidine residues in ribonuclease Sa (RNase Sa) (pI=3.5) and a variant in which five acidic amino acids have been changed to lysine (5K) (pI=10.2) was measured and compared to pK values of model histidine-containing peptides. The pK of His53 is elevated by two pH units (pK=8.61) in RNase Sa and by nearly one pH unit (pK=7.39) in 5K at low salt relative to the pK of histidine in the model peptides (pK=6.6). The pK for His53 remains elevated in 1.5M NaCl (pK=7.89). The elevated pK for His53 is a result of screenable electrostatic interactions, particularly with Glu74, and a non-screenable hydrogen bond interaction with water. The pK of His85 in RNase Sa and 5K is slightly below the model pK at low salt and merges with this value at 1.5M NaCl. The pK of His85 reflects mainly effects of long-range Coulombic interactions that are screenable by salt. The tautomeric states of the neutral histidine residues are changed by charge reversal. The histidine pK values in RNase Sa are always higher than the pK values in the 5K variant. These results emphasize that the net charge of the protein influences the pK values of the histidine residues. Structure-based pK calculations capture the salt-dependence relatively well but are unable to predict absolute histidine pK values.  相似文献   

5.
1. A precursor to small stable RNA, 10Sa RNA, accumulates in large amounts in a temperature sensitive RNase E mutant at non-permissive temperatures, and somewhat in an rnc (RNase III-) mutant, but not in an RNase P- mutant (rnp) or wild type E. coli cells. 2. Since p10Sa RNA was not processed by purified RNase E and III in customary assay conditions, we purified p10Sa RNA processing activity about 700-fold from wild type E. coli cells. 3. Processing of p10Sa RNA by this enzyme shows an absolute requirement for a divalent cation with a strong preference for Mn2+ over Mg2+. Other divalent cations could not replace Mn2+. 4. Monovalent cations (NH+4, Na+, K+) at a concentration of 20 mM stimulated the processing of p10Sa RNA and a temperature of 37 degrees C and pH range of 6.8-8.2 were found to be optimal. 5. The enzyme retained half of its p10Sa RNA processing activity after 30 min incubation at 50 degrees C. 6. Further characterization of this activity indicated that it is RNase III. 7. To further confirm that the p10Sa RNA processing activity is RNase III, we overexpressed the RNase III gene in an E. coli cells that lacks RNase III activity (rnc mutant) and RNase III was purified using one affinity column, agarose.poly(I).poly(C). 8. This RNase III preparation processed p10Sa RNA in a similar way as observed using the p10Sa RNA processing activity purified from wild type E. coli cells, confirming that the first step of p10Sa RNA processing is carried out by RNase III.  相似文献   

6.
The role of electrostatic interactions in the assembly of a native protein structure was studied using fragment complementation. Contributions of salt, pH, or surface charges to the kinetics and equilibrium of calbindin D(9k) reconstitution was measured in the presence of Ca(2+) using surface plasmon resonance and isothermal titration calorimetry. Whereas surface charge substitutions primarily affect the dissociation rate constant, the association rates are correlated with subdomain net charge in a way expected for Coulomb interactions. The affinity is reduced in all mutants, with the largest effect (260-fold) observed for the double mutant K25E+K29E. At low net charge, detailed charge distribution is important, and charges remote from the partner EF-hand have less influence than close ones. The effects of salt and pH on the reconstitution are smaller than mutational effects. The interaction between the wild-type EF-hands occurs with high affinity (K(A) = 1.3 x 10(10) M(-1); K(D) = 80 pM). The enthalpy of association is overall favorable and there appears to be a very large favorable entropic contribution from the desolvation of hydrophobic surfaces that become buried in the complex. Electrostatic interactions contribute significantly to the affinity between the subdomains, but other factors, such as hydrophobic interactions, dominate.  相似文献   

7.
The two most buried carboxyl groups in ribonuclease Sa (RNase Sa) are Asp33 (99% buried; pK 2.4) and Asp79 (85% buried; pK 7.4). Above these pK values, the stability of the D33A variant is 6kcal/mol less than wild-type RNase Sa, and the stability of the D79A variant is 3.3kcal/mol greater than wild-type RNase Sa. The key structural difference between the carboxyl groups is that Asp33 forms three intramolecular hydrogen bonds, and Asp79 forms no intramolecular hydrogen bond. Here, we focus on Asp79 and describe studies of 11 Asp79 variants. Most of the variants were at least 2kcal/mol more stable than wild-type RNase Sa, and the most interesting was D79F. At pH 3, below the pK of Asp79, RNase Sa is 0.3kcal/mol more stable than the D79F variant. At pH 8.5, above the pK of Asp79, RNase Sa is 3.7kcal/mol less stable than the D79F variant. The unfavorable contribution of Asp79 to the stability appears to result from the Born self-energy of burying the charge and, more importantly, from unfavorable charge-charge interactions. To counteract the effect of the negative charge on Asp79, we prepared the Q94K variant and the crystal structure showed that the amino group of the Lys formed a hydrogen-bonded ion pair (distance, 2.71A; angle, 100 degrees ) with the carboxyl group of Asp79. The stability of the Q94K variant was about the same as the wild-type at pH 3, where Asp79 is uncharged, but 1kcal/mol greater than that of wild-type RNase Sa at pH 8.5, where Asp79 is charged. Differences in hydrophobicity, steric strain, Born self-energy, and electrostatic interactions all appear to contribute to the range of stabilities observed in the variants. When it is possible, replacing buried, non-hydrogen bonded, ionizable side-chains with non-polar side-chains is an excellent means of increasing protein stability.  相似文献   

8.
A precursor to 10Sa RNA accumulates in an rne mutant. However, the present studies indicate that RNase III is the enzyme that processes this RNA. Cell extracts prepared from an rne mutant failed to cleave p10Sa RNA, whereas E coli wild type, rne and rnp cell extracts processed p10Sa RNA under specific assay conditions that require the presence of Mn2+ but not under the customary conditions used for assaying RNase III. That the p10Sa cleaving activity is solely RNase III was confirmed by comparing the increase in p10Sa and poly(A).poly(U) cleaving activities in a strain harboring a plasmid carrying an RNase III gene as compared to a normal E coli strain. It is of interest that these 2 substrates are cleaved by RNase III efficiently, but under 2 different assay conditions. In all strains tested, with normal or elevated levels of RNase III, RNase III fractionates predominantly with the membrane. Further characterization of the maturation of 10Sa RNA revealed that the processing of 10Sa RNA is a 2 step reaction involving 2 separate activities, both sensitive to heat and proteinase K treatment. The first step is catalyzed by RNase III, and results in the formation of a molecule, p10Sa', which is larger than the mature 10Sa RNA. The second activity catalyzes the conversion of p10S' to 10Sa RNA, and this step does not require a divalent cation. The second activity is not any of the known processing endoribonucleases, RNase III, E or P, but could be a new enzyme having no obligate requirement for a divalent cation.  相似文献   

9.
In this study, highly purified (HP) CD3-positive N901-negative T lymphocytes could be induced to become natural killer (NK)-like in culture in the presence of recombinant interleukin-2 (rIL-2) and phytohemagglutinin (PHA). Thus, purified CD3+ N901- T cells from fresh human peripheral blood were obtained by negative selection using an indirect panning technique. To ensure that T lymphocyte fractions were completely devoid of any detectable NK cells, two additional purification procedures were employed: incubation of post-pan T cells with the NK-cytotoxic lysomotropic agent L-leucinemethylester, and complement-mediated lysis using the NK cell specific NKH1a monoclonal antibody. Purity of CD3+ N901- cells could be confirmed by surface marker analysis, whereby two NK-associated antigens, N901 and H-25, were undetectable, while 94 +/- 1% of cells expressed the CD3 (Leu-4) antigen. On functional analysis, fresh HP CD3+ N901- cells exhibited no cytotoxic activity against the standard NK target K562. When HP NK-depleted T lymphocytes were cultured for 7 days in the presence of rIL-2 (100 U/ml), neither surface antigen expression nor cytotoxic activity against K562 changed significantly. However, significant cytotoxicity against K562 [18 +/- 5% specific lysis at 25:1 effector:target (E/T) ratio] could be induced when HP CD3+ N901- cells were grown for 7 days in the presence of rIL-2 and PHA (0.5% v/v). Concomitantly, antigens N901 and H-25 were found to be coexpressed on a minor proportion (22 +/- 16 and 22 +/- 6%, respectively) of CD3+ (88 +/- 2% on day 7) cells. Four-week long-term culture of HP NK-depleted T cells in the presence of rIL-2 and PHA yielded a continuous increase in cytotoxicity against K562 cells (0 up to 46% specific lysis at 25:1 E/T ratio). Of particular interest was the emergence of cytotoxicity against the NK-resistant Daudi cell target (15 +/- 8% specific lysis at 25:1 E/T ratio on day 21). Expression of antigens N901 and H-25 as well as CD3 remained essentially unchanged in long-term culture. In sorting experiments, the H-25+ cell fraction was significantly enriched for cytotoxicity against K562, when compared to both H-25- and unseparated cell fractions. In summary, our results suggest that a proportion of HP CD3+ N901- T lymphocytes may give rise to cells that exhibit NK-like functional and phenotypic properties.  相似文献   

10.
The pore region of the majority of K+ channels contains the highly conserved GYGD sequence, known as the K+ channel signature sequence, where the GYG is critical for K+ selectivity (Heginbotham, L., T. Abramson, and R. MacKinnon. 1992. Science. 258:1152-1155). Exchanging the aspartate residue with asparagine in this sequence abolishes ionic conductance of the Shaker K+ channel (D447N) (Hurst, R.S., L. Toro, and E. Stefani. 1996. FEBS Lett. 388:59-65). In contrast, we found that the corresponding mutation (D292N) in the pore forming alpha subunit (hSlo) of the voltage- and Ca(2+)-activated K+ channel (BKCa, MaxiK) did not prevent conduction but reduced single channel conductance. We have investigated the role of outer pore negative charges in ion conduction (this paper) and channel gating (Haug, T., R. Olcese, T. Ligia, and E. Stefani. 2004. J. Gen Physiol. 124:185-197). In symmetrical 120 mM [K+], the D292N mutation reduced the outward single channel conductance by approximately 40% and nearly abolished inward K+ flow (outward rectification). This rectification was partially relieved by increasing the external K+ concentration to 700 mM. Small inward currents were resolved by introducing an additional mutation (R207Q) that greatly increases the open probability of the channel. A four-state multi-ion pore model that incorporates the effects of surface charge was used to simulate the essential properties of channel conduction. The conduction properties of the mutant channel (D292N) could be predicted by a simple approximately 8.5-fold reduction of the surface charge density without altering any other parameter. These results indicate that the aspartate residue in the BKCa pore plays a key role in conduction and suggest that the pore structure is not affected by the mutation. We speculate that the negative charge strongly accumulates K+ in the outer vestibule close to the selectivity filter, thus increasing the rate of ion entry into the pore.  相似文献   

11.
Transposon Tn10-mediated rearrangement was used to isolate a strain of Escherichia coli carrying a deletion in the rnd region which is known to encode the structural gene for the putative 3' tRNA processing nuclease, RNase D. Genetic analysis indicated that about 0.4-0.5 min of the chromosome in the 39.5-40.0 min region was deleted. The mutant strain was devoid of RNase D activity, but other RNase activities were unaffected. The viability of the mutant strain and its normal growth characteristics indicate that RNase D is not essential for E. coli survival. The normal plating efficiency in this mutant host of wild type T4 and a T4 psu1+-amber double mutant indicates that RNase D is also not required for T4 growth or psu1+-tRNA processing. The implications of these findings for the role of RNase D in bacterial and bacteriophage tRNA metabolism, and the possible involvement of alternative enzymes, are discussed.  相似文献   

12.
13.
J Futami  T Maeda  M Kitazoe  E Nukui  H Tada  M Seno  M Kosaka  H Yamada 《Biochemistry》2001,40(25):7518-7524
Carboxyl groups of bovine RNase A were amidated with ethylenediamine (to convert negative charges of carboxylate anions to positive ones), 2-aminoethanol (to eliminate negative charges), and taurine (to keep negative charges), respectively, by a carbodiimide reaction. Human RNase 1 was also modified with ethylenediamine. Surprisingly, the modified RNases were all cytotoxic toward 3T3-SV-40 cells despite their decreased ribonucleolytic activity. However, their enzymatic activity was not completely eliminated by the presence of excess cytosolic RNase inhibitor (RI). As for native RNase A and RNase 1 which were not cytotoxic, they were completely inactivated by RI. More interestingly, within the cytotoxic RNase derivatives, cytotoxicity correlated well with the net positive charge. RNase 1 and RNase A modified with ethylenediamine were more cytotoxic than naturally occurring cytotoxic bovine seminal RNase. An experiment using the fluorescence-labeled RNase derivatives indicated that the more cationic RNases were more efficiently adsorbed to the cells. Thus, it is suggested that the modification of carboxyl groups could change complementarity of RNase to RI and as a result endow RNase cytotoxicity and that cationization enhances the efficiency of cellular uptake of RNase so as to strengthen its cytotoxicity. The finding that an extracellular human enzyme such as RNase 1 could be effectively internalized into the cell by cationization suggests that cationization is a simple strategy for efficient delivery of a protein into cells and may open the way of the development of new therapeutics.  相似文献   

14.
It is difficult to increase protein stability by adding hydrogen bonds or burying nonpolar surface. The results described here show that reversing the charge on a side chain on the surface of a protein is a useful way of increasing stability. Ribonuclease T1 is an acidic protein with a pI approximately 3.5 and a net charge of approximately -6 at pH 7. The side chain of Asp49 is hyperexposed, not hydrogen bonded, and 8 A from the nearest charged group. The stability of Asp49Ala is 0.5 kcal/mol greater than wild-type at pH 7 and 0.4 kcal/mol less at pH 2.5. The stability of Asp49His is 1.1 kcal/mol greater than wild-type at pH 6, where the histidine 49 side chain (pKa = 7.2) is positively charged. Similar results were obtained with ribonuclease Sa where Asp25Lys is 0.9 kcal/mol and Glu74Lys is 1.1 kcal/mol more stable than the wild-type enzyme. These results suggest that protein stability can be increased by improving the coulombic interactions among charged groups on the protein surface. In addition, the stability of RNase T1 decreases as more hydrophobic aromatic residues are substituted for Ala49, indicating a reverse hydrophobic effect.  相似文献   

15.
Laurents DV  Scholtz JM  Rico M  Pace CN  Bruix M 《Biochemistry》2005,44(21):7644-7655
The conformational stability of ribonuclease Sa (RNase Sa) has been measured at the per-residue level by NMR-monitored hydrogen exchange at pH* 5.5 and 30 degrees C. In these conditions, the exchange mechanism was found to be EXII. The conformational stability calculated from the slowest exchanging amide groups was found to be 8.8 kcal/mol, in close agreement with values determined by spectroscopic methods. RNase Sa is curiously rich in acidic residues (pI = 3.5) with most basic residues being concentrated in the active-site cleft. The effects of dissolved salts on the stability of RNase Sa was studied by thermal denaturation experiments in NaCl and GdmCl and by comparing hydrogen exchange rates in 0.25 M NaCl to water. The protein was found to be stabilized by salt, with the magnitude of the stabilization being influenced by the solvent exposure and local charge environment at individual amide groups. Amide hydrogen exchange was also measured in 0.25, 0.50, 0.75, and 1.00 M GdmCl to characterize the unfolding events that permit exchange. In contrast to other microbial ribonucleases studied to date, the most protected, globally exchanging amides in RNase Sa lie not chiefly in the central beta strands but in the 3/10 helix and an exterior beta strand. These structural elements are near the Cys7-Cys96 disulfide bond.  相似文献   

16.
The ionizable groups in proteins with the lowest pKs are the carboxyl groups of aspartic acid side-chains. One of the lowest, pK=0.6, is observed for Asp76 in ribonuclease T1. This low pK appeared to result from hydrogen bonds to a water molecule and to the side-chains of Asn9, Tyr11, and Thr91. The results here confirm this by showing that the pK of Asp76 increases to 1.7 in N9A, to 4.0 in Y11F, to 4.2 in T91V, to 4.4 in N9A+Y11F, to 4.9 in N9A+T91V, to 5.9 in Y11F+T91V, and to 6.4 in the triple mutant: N9A+Y11F+T91V. In ribonuclease Sa, the lowest pK=2.4 for Asp33. This pK increases to 3.9 in T56A, which removes the hydrogen bond to Asp33, and to 4.4 in T56V, which removes the hydrogen bond and replaces the -OH group with a -CH(3) group. It is clear that hydrogen bonds are able to markedly lower the pK values of carboxyl groups in proteins. These same hydrogen bonds make large contributions to the conformational stability of the proteins. At pH 7, the stability of D76A ribonuclease T1 is 3.8 kcal mol(-1) less than wild-type, and the stability of D33A ribonuclease Sa is 4.1 kcal mol(-1) less than wild-type. There is a good correlation between the changes in the pK values and the changes in stability. The results suggest that the pK values for these buried carboxyl groups would be greater than 8 in the absence of hydrogen bonds, and that the hydrogen bonds and other interactions of the carboxyl groups contribute over 8 kcal mol(-1) to the stability.  相似文献   

17.
3-Oxo-Delta(5)-steroid isomerase (KSI) catalyzes the isomerization of a variety of 3-oxo-Delta(5)-steroids to their conjugated Delta(4) isomers. The mechanism involves sequential enolization and ketonization, with Asp-38 acting to transfer a proton from C-4 to C-6 through a dienol(ate) intermediate. We have previously proposed that this intermediate is anionic, with stabilization provided from direct hydrogen bonding from Tyr-14 and Asp-99 to the oxygen of the steroid. In this work, we analyze the binding of substituted 2-naphthols, which are analogues of the intermediate dienol, to the D38E KSI mutant and the corresponding double mutants lacking one of the two electrophilic groups (D38E/Y14F and D38E/D99A). The binding of these naphthols to the mutant KSIs at pH 7 is described by the modified Bronsted equation: log K(D) = alpha(pK(a)) + constant, where K(D) is the dissociation constant of the complex. The high value of alpha for D38E (alpha = 0.87 +/- 0.06) indicates that the negative charge in these D38E-naphthol complexes is localized almost exclusively on the bound ligand. In contrast, values of alpha for the double mutants (alpha = 0.28 +/- 0.02 for D38E/Y14F and alpha = 0.25 +/- 0.02 for D38E/D99A) are consistent with very little negative charge on the oxygen of the bound naphthol. Ultraviolet spectra of 5-nitro-2-naphthol and the fluorescence spectra of equilenin bound to these mutants support this interpretation. Extrapolation of these results to the intermediate in the catalytic reaction suggests that for the reaction with D38E, the intermediate is a negatively charged dienolate with hydrogen bonding from both Tyr-14 and Asp-99. Removal of either one of these H-bond donors (Tyr-14 or Asp-99) causes destabilization of the anion and results in a dienol enzyme-intermediate complex rather than a dienolate.  相似文献   

18.
Gaining a better understanding of the denatured state ensemble of proteins is important for understanding protein stability and the mechanism of protein folding. We studied the folding kinetics of ribonuclease Sa (RNase Sa) and a charge-reversal variant (D17R). The refolding kinetics are similar, but the unfolding rate constant is 10-fold greater for the variant. This suggests that charge-charge interactions in the denatured state and the transition state ensembles are more favorable in the variant than in RNase Sa, and shows that charge-charge interactions can influence the kinetics and mechanism of protein folding.  相似文献   

19.
Voltage-dependent inactivation of CaV2.3 channels was investigated using point mutations in the beta-subunit-binding site (AID) of the I-II linker. The quintuple mutant alpha1E N381K + R384L + A385D + D388T + K389Q (NRADK-KLDTQ) inactivated like the wild-type alpha1E. In contrast, mutations of alpha1E at position R378 (position 5 of AID) into negatively charged residues Glu (E) or Asp (D) significantly slowed inactivation kinetics and shifted the voltage dependence of inactivation to more positive voltages. When co-injected with beta3, R378E inactivated with tau(inact) = 538 +/- 54 ms (n = 14) as compared with 74 +/- 4 ms (n = 21) for alpha1E (p < 0.001) with a mid-potential of inactivation E(0.5) = -44 +/- 2 mV (n = 10) for R378E as compared with E(0.5) = -64 +/- 3 mV (n = 9) for alpha1E. A series of mutations at position R378 suggest that positively charged residues could promote voltage-dependent inactivation. R378K behaved like the wild-type alpha1E whereas R378Q displayed intermediate inactivation kinetics. The reverse mutation E462R in the L-type alpha1C (CaV1.2) produced channels with inactivation properties comparable to alpha1E R378E. Hence, position 5 of the AID motif in the I-II linker could play a significant role in the inactivation of Ca(V)1.2 and CaV2.3 channels.  相似文献   

20.
This article utilized “protein charge ladders”—chemical derivatives of proteins with similar structure, but systematically altered net charge—to quantify how missense mutations that cause amyotrophic lateral sclerosis (ALS) affect the net negative charge (Z) of superoxide dismutase-1 (SOD1) as a function of subcellular pH and Zn2+ stoichiometry. Capillary electrophoresis revealed that the net charge of ALS-variant SOD1 can be different in sign and in magnitude—by up to 7.4 units per dimer at lysosomal pH—than values predicted from standard pKa values of amino acids and formal oxidation states of metal ions. At pH 7.4, the G85R, D90A, and G93R substitutions diminished the net negative charge of dimeric SOD1 by up to +2.29 units more than predicted; E100K lowered net charge by less than predicted. The binding of a single Zn2+ to mutant SOD1 lowered its net charge by an additional +2.33 ± 0.01 to +3.18 ± 0.02 units, however, each protein regulated net charge when binding a second, third, or fourth Zn2+Z < 0.44 ± 0.07 per additional Zn2+). Both metalated and apo-SOD1 regulated net charge across subcellular pH, without inverting from negative to positive at the theoretical pI. Differential scanning calorimetry, hydrogen-deuterium exchange, and inductively coupled plasma mass spectrometry confirmed that the structure, stability, and metal content of mutant proteins were not significantly affected by lysine acetylation. Measured values of net charge should be used when correlating the biophysical properties of a specific ALS-variant SOD1 protein with its observed aggregation propensity or clinical phenotype.  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